7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize...

33
1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 1 post-synaptic compartments 2 3 Pragya Goel 1,2 , Samantha Nishimura 1 , Karthik Chetlapalli 1 , Xiling Li 1,3 , Catherine Chen 1 , and 4 Dion Dickman 1, * 5 6 1 Department of Neurobiology, University of Southern California, Los Angeles, CA 90089, USA 7 2 USC Graduate Program in Molecular Biology 8 3 USC Graduate Program in Neuroscience 9 10 Keywords: active zone, homeostasis, synaptic plasticity, Drosophila, neuromuscular junction 11 12 Running title: Target-specific homeostatic plasticity 13 14 *Correspondence: 15 Dion Dickman, PhD 16 Department of Neurobiology 17 University of Southern California 18 Los Angeles, CA 90089 19 Email: [email protected] 20 ORCID: 0000-0003-1884-284X 21 Phone: (213) 740-7533 22 Fax: (877) 518-2393 23 24 25 26 . CC-BY-NC-ND 4.0 International license was not certified by peer review) is the author/funder. It is made available under a The copyright holder for this preprint (which this version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108 doi: bioRxiv preprint . CC-BY-NC-ND 4.0 International license was not certified by peer review) is the author/funder. It is made available under a The copyright holder for this preprint (which this version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108 doi: bioRxiv preprint . CC-BY-NC-ND 4.0 International license was not certified by peer review) is the author/funder. It is made available under a The copyright holder for this preprint (which this version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108 doi: bioRxiv preprint . CC-BY-NC-ND 4.0 International license was not certified by peer review) is the author/funder. It is made available under a The copyright holder for this preprint (which this version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108 doi: bioRxiv preprint

Transcript of 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize...

Page 1: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

1

Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 1

post-synaptic compartments 2

3

Pragya Goel1,2, Samantha Nishimura1, Karthik Chetlapalli1, Xiling Li1,3, Catherine Chen1, and 4

Dion Dickman1,* 5

6

1Department of Neurobiology, University of Southern California, Los Angeles, CA 90089, USA 7

2USC Graduate Program in Molecular Biology 8

3USC Graduate Program in Neuroscience 9

10

Keywords: active zone, homeostasis, synaptic plasticity, Drosophila, neuromuscular junction 11

12

Running title: Target-specific homeostatic plasticity 13

14

*Correspondence: 15

Dion Dickman, PhD 16

Department of Neurobiology 17

University of Southern California 18

Los Angeles, CA 90089 19

Email: [email protected] 20

ORCID: 0000-0003-1884-284X 21

Phone: (213) 740-7533 22

Fax: (877) 518-2393 23

24

25

26

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 2: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

2

ABSTRACT 27

Neurons must establish and stabilize connections made with diverse targets, each with distinct 28

demands and functional characteristics. At Drosophila neuromuscular junctions, synaptic 29

strength remains stable in a manipulation that simultaneously induces hypo-innervation on one 30

target and hyper-innervation on the other. However, the expression mechanisms that achieve 31

this exquisite target-specific homeostatic control remain enigmatic. Here, we identify the distinct 32

target-specific homeostatic expression mechanisms. On the hypo-innervated target, an increase 33

in postsynaptic glutamate receptor (GluR) abundance is sufficient to compensate for reduced 34

innervation, without any apparent presynaptic adaptations. In contrast, a target-specific 35

reduction in presynaptic neurotransmitter release probability is reflected by a decrease in active 36

zone components restricted to terminals of hyper-innervated targets. Finally, loss of 37

postsynaptic GluRs on one target induces a compartmentalized, homeostatic enhancement of 38

presynaptic neurotransmitter release called presynaptic homeostatic potentiation that can be 39

precisely balanced with the adaptations required for both hypo- and hyper-innervation to 40

maintain stable synaptic strength. Thus, distinct anterograde and retrograde signaling systems 41

operate at pre- and post-synaptic compartments to enable target-specific, homeostatic control of 42

neurotransmission. 43

44

45

46

47

48

49

50

51

52

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 3: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

3

INTRODUCTION 53

Synapses are spectacularly diverse in their morphology, physiology, and functional 54

characteristics. These differences are reflected in the molecular composition and abundance of 55

synaptic components at heterogenous synaptic subtypes in central and peripheral nervous 56

systems (Atwood and Karunanithi, 2002; Branco and Staras, 2009; O’Rourke et al., 2012). 57

Interestingly, the structure and function of synapses can also vary substantially across terminals 58

of an individual neuron (Fekete et al., 2019; Grillo et al., 2018; Guerrero et al., 2005) and drive 59

input-specific presynaptic plasticity (Letellier et al., 2019). Both Hebbian and homeostatic 60

plasticity mechanisms can work locally and globally at specific synapses to tune synapse 61

function, enabling stable yet flexible ranges of synaptic strength (Diering and Huganir, 2018; 62

Turrigiano, 2012; Vitureira and Goda, 2013). For example, homeostatic receptor scaling globally 63

adjusts GluR abundance, subtype, and/or functionality at dendrites (Turrigiano and Nelson, 64

2004) yet there is also evidence for synapse specificity (Béïque et al., 2011; Hou et al., 2008; 65

Sutton et al., 2006). Although a number of studies have begun to elucidate the factors that 66

enable both local and global modes of synaptic plasticity at synaptic compartments, it is less 67

appreciated how and why specific synapses undergo plasticity within the context and needs of 68

information transfer in a neural circuit. 69

One major force that sculpts the heterogeneity of synaptic strength is imposed through 70

the specific targets being innervated. For example, studies at neuromuscular synapses in the 71

stomatogastric system of lobsters have demonstrated that presynaptic terminals of the same 72

motor axon can concurrently undergo facilitation and depression due to differences in the 73

synapses made onto two postsynaptic muscle fibers (Katz et al., 1993). Furthermore, at 74

vertebrate neuromuscular junctions (NMJs), secreted factors from muscles can dictate which 75

motor neurons survive during development and in many cases their neurotransmitter phenotype 76

(Calderó et al., 1998; Schotzinger and Landis, 1990), and a parallel target-dependent control of 77

neuropeptide identity has been shown in the Drosophila central nervous system (Allan et al., 78

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 4: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

4

2003; Allan and Thor, 2015). In mammalian central neurons, factors such as BDNF secreted 79

from postsynaptic dendrites not only promote neuronal survival but also can homeostatically 80

enhance presynaptic neurotransmitter release and functional properties of neural circuits 81

(Jakawich et al., 2010; Park and Poo, 2013), while postsynaptic signaling through N-Cadherins 82

and mTORC1 can regulate presynaptic function (Henry et al., 2012; Vitureira et al., 2011). 83

Finally, at the Drosophila NMJ, presynaptic homeostatic plasticity can be expressed at a subset 84

of terminals on a single motor neuron depending on GluR functionality at particular targets (Li et 85

al., 2018a), demonstrating that this form of homeostatic plasticity is target-specific and strongly 86

suggesting it is also synapse-specific. Together, these studies and others have demonstrated 87

that the physiologic, metabolic, and/or structural properties at terminals of a single neuron can 88

be selectively modulated according to the identity and needs of the targets and synapses they 89

innervate. However, the nature of the trans-synaptic dialogue and the molecular mechanisms 90

that achieve target-specific plasticity are not well understood. 91

A seminal study published over 20 years ago found that distinct target-specific 92

modulations in synaptic activity maintain stable neurotransmission following biased innervation 93

at terminals of motor neurons at the Drosophila NMJ (Davis and Goodman, 1998). In this 94

manipulation, biased innervation is achieved by overexpression of the trans-synaptic cell 95

adhesion factor fasciculin II (fasII) on one of the two muscle targets innervated by motor 96

neurons (Davis and Goodman, 1998). This leads to hyper-innervation of the target 97

overexpressing fasII at the expense of the adjacent target, which is hypo-innervated. 98

Remarkably, synaptic strength, as assessed by electrophysiological recordings, was maintained 99

at levels similar in amplitude to normally innervated NMJ targets. Since this pioneering study, 100

however, the molecular and cellular expression mechanisms that achieve this target-specific 101

homeostatic modulation have remained enigmatic. 102

We have investigated how terminals of an individual neuron adapt to simultaneous hypo- 103

and hyper-innervation to maintain stable synaptic strength on two adjacent targets. Our analysis 104

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 5: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

5

reveals that a novel homeostatic signaling system operates in the hypo-innervated target to 105

precisely enhance the abundance of postsynaptic GluRs, offsetting reduced presynaptic 106

neurotransmitter release and stabilizing synaptic strength. In contrast, no apparent adaptations 107

are observed in the hyper-innervated target. Rather, presynaptic release probability is 108

homeostatically reduced, accompanied by a target-specific decrease in the abundance and 109

density of active zone components. Finally, we find that presynaptic homeostatic potentiation 110

can be selectively induced and expressed at synapses on one target and balanced with biased 111

innervation to sustain stable synaptic strength. This work reveals the striking interplay of target-112

specific homeostats modulating the efficacy of neurotransmission across synaptic terminals. 113

114

RESULTS 115

Target-specific mechanisms maintain stable synaptic strength at hypo- and hyper-116

innervated NMJs 117

We first sought to reproduce and confirm the biased innervation and synaptic electrophysiology 118

reported in Davis and Goodman, 1998. At Drosophila larval NMJs, motor neurons distribute 119

their synaptic terminals roughly evenly between two distinct targets – as demonstrated by the 120

NMJs made onto muscles 6 and 7 (Fig. 1A; left). This stereotyped pattern of innervation can be 121

visualized by immunostaining the NMJ with antibodies that recognize the neuronal membrane 122

(HRP) and synaptic vesicles (Synapsin; SYN), which demonstrates ~60% boutons on the larger 123

muscle 6 and ~40% on the smaller muscle 7 (Fig. 1B; left and Table S2). To bias innervation on 124

these targets, we used the H94-Gal4 driver to drive expression of the cell adhesion molecule 125

fasciculin II (fasII) early in development selectively on muscle 6 [(M6>FasII; (Davis and 126

Goodman, 1998)]. Immunostaining of M6>FasII NMJs confirmed biased innervation with ~150% 127

of boutons above controls on muscle 6 (hyper-innervated), and a parallel reduction of ~50% in 128

boutons on muscle 7 (hypo-innervated) (Fig. 1A-E), consistent with the previous study (Davis 129

and Goodman, 1998). However, despite these opposing changes in bouton numbers, 130

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 6: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

6

electrophysiological recordings of M6>FasII found that synaptic strength, measured by the 131

excitatory postsynaptic potential (EPSP) amplitude, was similar on both targets and unchanged 132

from their respective controls (Fig. 1C,D,E). This implies target-specific mechanisms modulate 133

neurotransmission on hypo- and hyper-innervated terminals to maintain stable NMJ strength. 134

To gain insight into how EPSP amplitudes remain similar to baseline values at NMJs 135

with biased innervation, we next examined miniature neurotransmission. On hypo-innervated 136

muscle 7, mEPSP amplitudes were significantly increased by ~40% compared to baseline 137

values (Fig. 1C,D), as previously observed (Davis and Goodman, 1998). Quantal content (QC) 138

was thus decreased by ~40%, a value similar in magnitude to the reduction in bouton number 139

(Fig. 1D). In contrast, mEPSP amplitude was not significantly different on the hyper-innervated 140

muscle 6 NMJ compared to baseline (Fig. 1C,E), with no apparent change in quantal content 141

(Fig. 1E), as previously observed (Davis and Goodman, 1998). Finally, analysis of quantal 142

content normalized per bouton on muscle 6 NMJs revealed an ~30% reduction (Fig. 1E), 143

suggesting a target-specific, homeostatic decrease in presynaptic neurotransmitter release. 144

Together, this data indicates that distinct target-specific mechanisms operate to stabilize 145

neurotransmission at hypo- vs. hyper-innervated NMJs. 146

147

A homeostatic increase in postsynaptic GluR abundance stabilizes synaptic strength on 148

hypo-innervated targets 149

It was previously reported that at hypo-innervated NMJs following M6>FasII, levels of the 150

postsynaptic GluR subunit GluRIIA were increased (Goel and Dickman, 2018). We considered 151

four additional possible mechanisms that could, in principle, contribute to the observed 152

enhancement in quantal size at hypo-innervated NMJs. First, increased presynaptic vesicle size 153

could lead to enhanced glutamate emitted per vesicle, as has been documented in endocytosis 154

mutants and following overexpression of the vesicular glutamate transporter (Daniels et al., 155

2004; Goel et al., 2019a). Second, multivesicular release has been observed in some systems 156

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 7: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

7

(Rudolph et al., 2015) and was raised as a possibility in the original study to potentially explain 157

the increased quantal size (Davis and Goodman, 1998), although there remains no evidence for 158

multi-vesicular release at the fly NMJ. In addition to these two presynaptic mechanisms, other 159

postsynaptic mechanisms are possible that could explain the increased mEPSP amplitude on 160

hypo-innervated NMJs, parallel to the ones that have been documented in mammalian forms of 161

homeostatic receptor scaling (Diering and Huganir, 2018; Turrigiano, 2008). These include 162

increases in the abundance, subtype, and/or functionality of additional postsynaptic GluRs, 163

including GluRIIB-containing receptors, as enhanced levels of GluRIIA-containing GluRs were 164

recently reported at hypo-innervated NMJs in Drosophila (Goel and Dickman, 2018; Goel et al., 165

2019b). 166

We therefore examined postsynaptic GluR levels in hypo-innervated targets induced by 167

M6>FasII. At the Drosophila NMJ, the postsynaptic response to glutamate is mediated by two 168

subtypes of GluRs, GluRIIA- and GluRIIB-containing receptors. Both subtypes are composed of 169

the essential subunits GluRIIC, GluRIID, and GluRIIE but differ in containing either GluRIIA or 170

GluRIIB subunits (DiAntonio, 2006; Qin et al., 2005). We immunostained hypo-innervated NMJs 171

using antibodies against GluRIIA, GluRIIB, and the common GluRIID subunits and observed an 172

~40% decrease in the number of GluR puncta comparted to wild type muscle 7 (Fig. S1A,C), 173

reflecting reduced innervation. However, we found an increase in the intensity of all GluR 174

subunits in hypo-innervated NMJs compared to wild type muscle 7 (Fig. S1B,D). In principle, the 175

40% increase in GluR abundance is sufficient to explain the increased quantal size and to offset 176

the 40% reduction in quantal content to homeostatically maintain stable synaptic strength 177

despite reduced innervation. Consistent with this, we observed no adaptations in the anatomical 178

or functional number of release sites, nor in the size of the readily releasable pool (Fig. S2). 179

These lines of evidence indicate that presynaptic terminals of hypo-innervated NMJs function 180

similarly to wild type, with presynaptic neurotransmitter released onto the muscle 7 NMJ simply 181

reduced by 40%. Thus, a 40% increase in postsynaptic GluR abundance is sufficient to maintain 182

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 8: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

8

synaptic strength at hypo-innervated NMJs without reason to invoke other homeostatic 183

adaptations. 184

185

Hyper-innervation induces a homeostatic decrease in presynaptic release probability 186

We next sought to characterize the expression mechanism that enables stable neurotransmitter 187

output on the hyper-innervated target. It was previously demonstrated that a homeostatic 188

reduction in presynaptic release probability was expressed at hyper-innervated NMJs (Davis 189

and Goodman, 1998). Consistent with this conclusion, and in contrast to the adjacent hypo-190

innervated NMJs, we did not observe any significant changes in postsynaptic GluR levels (Fig. 191

S1E-H). We next performed a series of electrophysiological assays to probe presynaptic 192

function on the hyper-innervated NMJ. First, we used failure analysis to assess presynaptic 193

release independently of miniature transmission by measuring the number of failed release 194

events in very low extracellular Ca2+ concentrations (0.15 mM; see Methods). We observed no 195

significant difference in the failure rates on hyper-innervated NMJs compared to wild type (Fig. 196

2B), consistent with overall quantal content released being unchanged at these NMJs. Next, we 197

probed release probability by determining paired pulse ratios in moderate and high extracellular 198

Ca2+. At 0.4 mM Ca2+, we observed an increase in paired pulse facilitation (PPF) at hyper-199

innervated NMJs compared to wild type (Fig. 2C,D), while at 1.5 mM Ca2+, paired-pulse 200

depression (PPD) was reduced at hyper-innervated NMJs (Fig. 2E,F). Since enhanced PPF and 201

reduced PPD is indicative of reduced release probability (Regehr, 2012), these results suggest 202

that a target-specific, homeostatic decrease in presynaptic release probability stabilizes 203

transmission at hyper-innervated NMJs. Although the PPF/PPD recordings suggested reduced 204

release probability at hyper-innervated terminals, the magnitude of the observed decrease 205

(~25%) was not sufficient to fully compensate for the ~50% increase in bouton numbers. We 206

found no change in the size of the RRP on hyper-innervated NMJs compared to wild type (Fig. 207

2G,H), suggesting that the size of the RRP at individual boutons might be reduced on muscle 6 208

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 9: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

9

of M6>FasII NMJs. Finally, no change in the total number of functional release sites was 209

observed on hyper-innervated targets (Fig. 2I-K), indicating a reduction in the number of release 210

sites participating in neurotransmission per bouton at hyper-innervated NMJs. Thus, a 211

homeostatic adjustment in the release probability per bouton selectively modulates transmission 212

at hyper-innervated NMJs without measurably impacting release at adjacent hypo-innervated 213

terminals. 214

215

A target-specific reduction in both active zone density and intensity is observed at hyper-216

innervated NMJs 217

Our electrophysiological data above suggests a reduction in both release probability and the 218

number of functional release sites at individual boutons of hyper-innervated NMJs. In principle, 219

a target-specific reduction in the number and/or function of anatomical release sites could 220

explain these electrophysiological properties. In addition, recent evidence indicates that bi-221

directional changes in the size and nano-structure of active zone architecture at Drosophila 222

NMJs can adjust release probability at individual active zones (Akbergenova et al., 2018; 223

Bohme et al., 2019; Goel et al., 2019a; Gratz et al., 2019). We therefore characterized the 224

number and intensity of individual active zones on hyper-innervated NMJs by immunostaining 225

the central scaffold BRP and endogenously tagged CaV2.1 calcium channels [CacsfGFP; (Gratz 226

et al., 2019)], defining each BRP punctum to be an active zone. Interestingly, while a ~55% 227

increase in bouton number was observed at hyper-innervated NMJs, the number of active 228

zones was only increased by ~20%, reflected in a concomitant decrease in active zone density 229

(Fig. 3A-E). Thus, hyper-innervated NMJs exhibit a target-specific reduction in the density of 230

release sites that is sufficient in magnitude to limit the increase in active zones to only about 231

20% despite an ~50% increase in innervation. 232

We also quantified the intensity of individual BRP puncta on hyper-innervated NMJs and 233

observed an ~20% decrease in the sum intensity compared to wild type (Fig. 3B,C,F). Similar 234

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 10: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

10

results for puncta density and intensity were found for CacsfGFP (Fig. 3B-F). Finally, given these 235

reductions in the density and intensity of active zone components, the total intensity of both 236

BRP and CacsfGFP per hyper-innervated NMJ was not significantly different from wild type 237

despite the increase in their total number (Fig. 3G). These results parallel recent studies that 238

have shown that while the number and intensity of individual active zones can vary at NMJs, the 239

total abundance of active zone protein remains constant (Goel et al., 2019a; Goel et al., 2019b; 240

Graf et al., 2009). Together, hyper-innervated NMJs express a target-specific reduction in both 241

the number and intensity of release sites and a parallel reduction in presynaptic release 242

probability that stabilizes synaptic strength, while no reciprocal changes are observed at hypo-243

innervated counterparts. 244

245

Distinct target-specific adaptations can homeostatically balance hyper-innervation and 246

GluR perturbation 247

When biased innervation of the NMJ is induced through M6>FasII, the hypo-innervated target 248

responds by homeostatically increasing GluR abundance, while the subset of motor neuron 249

terminals that hyper-innervate the adjacent target selectively reduce the number and apparent 250

abundance of active zone components. In our final set of experiments, we sought to determine 251

whether the target-specific homeostatic adaptations triggered by biased innervation could be 252

balanced with an additional target-specific homeostatic challenge. Presynaptic homeostatic 253

potentiation (PHP) is a well-studied form of homeostatic plasticity at the Drosophila NMJ. Here, 254

rapid pharmacological or chronic genetic manipulations that diminish postsynaptic GluR 255

functionality trigger a trans-synaptic retrograde signaling system that homeostatically increases 256

presynaptic glutamate release to maintain stable synaptic strength (Frank et al., 2020). 257

Recently, it was demonstrated that GluR knockdown specifically on muscle 6 can trigger PHP 258

selectively at the subset of synapses innervating muscle 6 without influencing transmission at 259

the synaptic terminals of the same motor neuron that innervate the adjacent muscle 7 (Li et al., 260

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 11: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

11

2018a), demonstrating a remarkable degree of compartmentalized expression of PHP. We 261

combined these manipulations to induce a simultaneous challenge of biased innervation and 262

GluR loss using fasII overexpression combined with GluRIIA knockdown selectively on muscle 6 263

(M6>FasII+GluRIIARNAi). We first validated if the combined manipulation was successful by 264

assaying synaptic growth and GluRIIA levels. Indeed, we observed the expected increase and 265

decrease in bouton numbers on muscles 6 and 7 respectively, with a near absence of GluRIIA 266

immunostaining selectively on muscle 6 (Fig. 4B-F). Thus, target-specific, homeostatic 267

challenges of biased innervation and GluR loss can be simultaneously induced by 268

overexpressing fasII and GluRIIARNAi selectively on muscle 6. 269

We next performed synaptic electrophysiology at both targets. On the hypo-innervated 270

muscle 7 of M6>FasII+GluRIIARNAi, neurotransmission was indistinguishable from M6>FasII 271

alone, with elevated mEPSP amplitudes, stable EPSP amplitudes, and reduced quantal content 272

observed (Fig. 4D,E). In contrast, on the hyper-innervated muscle 6 of M6>FasII+GluRIIARNAi, 273

mEPSP amplitudes were selectively reduced due to GluR knockdown, but synaptic strength 274

was maintained at baseline levels due to a homeostatic increase in quantal content (Fig. 4D,F). 275

This demonstrates that PHP can be robustly expressed and balanced with the adaptations 276

necessary to adjust release for hyper-innervation in a target-specific manner, without any 277

apparent changes in transmission at adjacent synapses of the hypo-innervated muscle 7. 278

Finally, we tested whether PHP can be acutely induced and balanced at hypo-innervated NMJs 279

after the adjustments made at muscle 6 of M6>FasII+GluRIIARNAi. We applied sub-blocking 280

concentrations of the GluR venom philanthotoxin-433 (PhTx) at NMJs for 10 mins. This acutely 281

induced PHP at wild type NMJs, with reduced mEPSP amplitude but EPSP amplitudes 282

unchanged from baseline due to a rapid, homeostatic increase in quantal content (Fig. S3A-D). 283

Application of PhTx to M6>FasII+ GluRIIARNAi NMJs had no significant change in mEPSP 284

amplitude or quantal content at muscle 6 due to GluRIIA knockdown (Fig. S3A,B,D). However, 285

PhTx application also induced robust PHP at muscle 7 NMJs in M6>FasII+ GluRIIARNAi, with a 286

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 12: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

12

significant reduction in mEPSP amplitude but normal EPSP amplitude due to enhanced quantal 287

content (Fig. S3A,C,E). These results demonstrate that presynaptic release sites at terminals of 288

the same neuron can be selectively modulated with exquisite target specificity to compensate 289

for GluR loss and can be superimposed with the homeostatic plasticity induced by biased 290

innervation. 291

292

DISCUSSION 293

Recent studies have shed light on how neurotransmission is stabilized when synaptic growth 294

and function is challenged (Davis and Muller, 2015; Frank et al., 2020; Goel et al., 2019a; Goel 295

et al., 2019b; Li et al., 2018b). However, less is known about how this stability is maintained 296

when neuronal terminals confront diverse and even opposing challenges in synaptic growth and 297

function. Here, we have combined biased innervation with acute and chronic challenges to 298

postsynaptic GluR function at targets shared by individual neurons. These experiments have 299

revealed two distinct target-specific mechanisms that enable stable transmission despite biased 300

innervation, operating at either pre- or postsynaptic compartments, and that can be balanced 301

with postsynaptic GluR perturbation. Importantly, these processes occur independently, without 302

impacting transmission within the same neuron on neighboring synapses made on the adjacent 303

target. This demonstrates a remarkable degree of compartmentalized autonomy in homeostatic 304

signaling and suggests an independence of local and global homeostats that work in concert to 305

balance synaptic strength. 306

307

Target-specific homeostatic scaling of postsynaptic GluR receptors 308

We took advantage of a previously established manipulation to bias synaptic innervation using 309

target-specific expression of the trans-synaptic cell adhesion protein FasII (Davis and 310

Goodman, 1998). On the hypo-innervated target, a selective upregulation in postsynaptic GluR 311

abundance was elicited that was sufficient in magnitude to offset reduced neurotransmitter 312

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 13: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

13

release and stabilize synaptic strength. This scaling of GluR abundance parallels a well-313

established mechanism of homeostatic synaptic plasticity in mammalian neurons termed 314

homeostatic receptor scaling (Chowdhury and Hell, 2018; Diering and Huganir, 2018; 315

Turrigiano, 2008). Although an enhanced upregulation of GluRs occurs in response to reduced 316

activity in mammalian central neurons, the GluR scaling revealed at the fly NMJ is unique. 317

GluRs at the Drosophila NMJ are quite stable, in contrast to the rapid dynamics of GluR 318

trafficking observed in dendritic spines of mammalian central neurons. This receptor stasis may 319

reflect a fundamental property of NMJs, where postsynaptic receptors have half-lives of ~7 days 320

in rodents (Salpeter and Harris, 1983) and over 24 hours in flies (Rasse et al., 2005). Although 321

NMJ receptors appear to be relatively stable under basal conditions and even in mutants in 322

which synaptic transmission and growth are perturbed (Goel et al., 2019b; Lee et al., 2013; 323

Saitoe et al., 2001), there is emerging evidence that specific challenges, including injury and 324

disease, can provoke relatively rapid remodeling of neurotransmitter receptors at postsynaptic 325

compartments of the NMJ (Goel and Dickman, 2018; Palma et al., 2011; Perry et al., 2017; Rich 326

and Lichtman, 1989). The temporal regulation and dynamics of the GluR plasticity we have 327

characterized at the NMJ in response to hypo-innervation is unclear but likely to be intertwined 328

with the developmental processes of NMJs as well as synaptic growth. 329

The induction mechanism involved in how reduced innervation is sensed to ultimately 330

instruct an adaptive increase in postsynaptic GluR abundance is unclear. It is notable that while 331

hypo-innervation in the M6>FasII manipulation elicits GluR scaling, a variety of mutations that 332

lead to synaptic undergrowth do not consistently change receptor levels (Banovic et al., 2010; 333

Goel et al., 2019b; Kaufmann et al., 2002; Marqués et al., 2002). Further, mutations that 334

severely reduce neurotransmitter release, including synaptotagmin and complexin mutants, do 335

not change GluR levels (Huntwork and Littleton, 2007; Lee et al., 2013; Saitoe et al., 2001). 336

Hence, hypo-innervation and/or reduced neurotransmitter release alone is unlikely to be 337

sufficient to induce postsynaptic GluR scaling. Rather, this form of homeostatic plasticity may be 338

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 14: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

14

dependent on the phenomenon of biased innervation between two targets shared by a single 339

neuron itself, implying some signaling between the motor neuron and/or the adjacent muscles 340

are involved. What is clear is that the postsynaptic signal transduction system that mediates 341

hypo-innervation-dependent GluR scaling is distinct from that which mediates retrograde PHP 342

signaling, as GluR scaling can still be expressed in conditions in which postsynaptic PHP 343

signaling is blocked (Goel and Dickman, 2018). Finally, the capacity for PHP to be robustly 344

expressed at NMJs that have undergone GluR scaling illuminates postsynaptic GluRs to be 345

central targets for both the induction and expression of homeostatic synaptic plasticity. 346

347

Target-specific modulation of active zones 348

In contrast to the exclusively postsynaptic adaptation observed in response to reduced 349

innervation, an entirely presynaptic mechanism stabilizes synaptic strength at hyper-innervated 350

muscles, expressed by a target-specific reduction in the number and intensity of active zone 351

components. Although a similar reduction in the abundance of active zone proteins at individual 352

release sites has recently been found in mutations that cause synaptic overgrowth at the NMJ 353

(Goel et al., 2019a; Goel et al., 2019b), the adaptations observed in the case of hyper-354

innervation is distinct in that they are 1) target-specific and 2) involve a reduction in active zone 355

number in addition to protein abundance. It is remarkable that both the number and abundance 356

of active zone components can be selectively reduced and calibrated at hyper-innervated 357

terminals without any apparent changes at adjacent terminals shared by the same neuron on 358

the hypo-innervated target. One attractive pathway that could be controlled to subserve this 359

target-specific modulation of active zone structure might involve the lysosomal adaptor Arl-8. 360

Arl-8 regulates the delivery of synaptic vesicle and active zone cargo to synapses (Klassen et 361

al., 2010; Vukoja et al., 2018), and was recently shown to promote the delivery of synaptic cargo 362

necessary to remodel active zones during PHP (Goel et al., 2019a). Because the abundance of 363

active zone components are remodeled during PHP (Bohme et al., 2016; Goel et al., 2017; 364

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 15: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

15

Gratz et al., 2019; Weyhersmuller et al., 2011) through an arl-8 dependent mechanism (Goel et 365

al., 2019a), and PHP can be expressed at a subset of terminals with target-specificity (Li et al., 366

2018a), it is tempting to speculate that Arl-8 may also be involved in the target-specific reduction 367

in active zones following hyper-innervation. 368

369

Biased innervation, presynaptic homeostatic plasticity, and information transfer at 370

synapses 371

Global synaptic strength is established during development through intrinsic genetic programs 372

and a dialogue between pre- and post-synaptic compartments. Robustness in this process is 373

ensured by signaling systems that have the capacity to sense and adapt to deviations outside of 374

physiological ranges, such as reductions or enhancements in synaptic growth (Goel et al., 375

2019b; Goel et al., 2019c; Keck et al., 2013; Tripodi et al., 2008; Yuan et al., 2011). 376

Superimposed on this foundation are forms of plasticity such as PHP, which appear to operate 377

as independent homeostats to maintain stable information transfer at synapses and within 378

neural circuits. Presynaptic terminals of a neuron therefore do not function as unitary 379

computational units but are rather compartmentally specialized and flexible according to the 380

physiologic needs of their targets during development and also following homeostatic 381

challenges. In addition to this target-specificity, there is also evidence for input-specificity across 382

dendrites that have the capacity to homeostatically modulate strength in rodent hippocampal 383

neurons (Jia et al., 2010; Katz et al., 2009; Letellier et al., 2019; Stuart and Spruston, 2015). 384

This remarkable control of synaptic activity enables the flexibility to locally adjust synaptic 385

strengths through input- and target-specificity while stabilizing overall network activity and 386

information processing. 387

388

MATERIALS AND METHODS 389

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 16: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

16

Fly stocks: Drosophila stocks were raised at 25°C on standard molasses food. The w1118 strain 390

is used as the wild type control unless otherwise noted as this is the genetic background in 391

which all genotypes are bred. Details of all stocks and their sources are listed in the Reagents 392

and Resource Table S1. 393

394

Immunocytochemistry: Third-instar larvae were dissected in ice cold 0 Ca2+ HL-3 and 395

immunostained using a standard protocol as described (Perry et al., 2017). In brief, larvae were 396

either fixed in Bouin’s fixative for 5 min (Sigma, HT10132-1L), 100% ice-cold ethanol for 5 min, 397

or 4% paraformaldehyde (PFA) for 10 min. Larvae were then washed with PBS containing 0.1% 398

Triton X-100 (PBST) for 30 min, blocked with 5% Normal Donkey Serum followed by overnight 399

incubation in primary antibodies at 4°C. Preparations were then washed 3x in PBST, incubated 400

in secondary antibodies for 2 hours, washed 3x in PBST, and equilibrated in 70% glycerol. Prior 401

to imaging, samples were mounted in VectaShield (Vector Laboratories). Details of all 402

antibodies, their source, dilution, and references are listed in Table S1. 403

404

Confocal imaging and analysis: Samples were imaged using a Nikon A1R Resonant 405

Scanning Confocal microscope equipped with NIS Elements software and a 60x APO 1.4 NA oil 406

immersion objective using separate channels with four laser lines (405, 488, 561, and 637 nm) 407

at room temperature. Boutons were counted using NMJs stained with anti-Synapsin or -vGlut 408

co-stained with anti-HRP on muscle 6/7 of segment A2 and A3, considering each Synapsin or 409

vGlut punctum to be a bouton. For fluorescence quantifications of postsynaptic GluRs and 410

active zone proteins, all genotypes within a data set were immunostained in the same tube with 411

identical reagents, then mounted and imaged in the same session. Z-stacks were obtained 412

using identical settings for all genotypes with z-axis spacing between 0.15-0.2 μm within an 413

experiment and optimized for detection without saturation of the signal. Maximum intensity 414

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 17: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

17

projections were used for quantitative image analysis with the NIS Elements General Analysis 415

toolkit. 416

To quantify sum puncta intensity, the total fluorescence intensity signal of individual 417

puncta were calculated without regard to area as described (Goel et al., 2019a). For each 418

particular sample set, thresholds were optimized to capture the dynamic range of intensity levels 419

within the wild type sample. This same threshold was then used to image all other genotypes in 420

the sample set, and all intensities were normalized to wild type values within an experimental 421

set. Active zones too closely spaced to be resolved (~5% of all analyzed) were excluded from 422

the analysis. Spot detection in the Nikon Elements Software was used to identify individual BRP 423

and Cac puncta as it resolves closely spaced puncta more accurately compared to thresholding. 424

Finally, to quantify total intensity per muscle, the sum fluorescence intensity for each individual 425

punctum was summed across the entire muscle. For calculation of BRP and Cac puncta 426

density, the total number of puncta at a particular muscle was divided by the neuronal 427

membrane area labeled by HRP spanning that muscle (Goel et al., 2019a). For image 428

representation only, the gain and contrast were adjusted identically for all genotypes within a 429

dataset. To show representative images of individual boutons, a particular area was selected 430

from the entire NMJ (denoted with a white box) and rotated and cropped to demonstrate 431

changes at boutons more clearly. 432

433

Electrophysiology: All dissections and recordings were performed in modified HL-3 saline 434

(Kikuma et al., 2017; Stewart et al., 1994) containing (in mM): 70 NaCl, 5 KCl, 10 MgCl2, 10 435

NaHCO3, 115 Sucrose, 5 Trehelose, 5 HEPES, and 0.4 CaCl2 (unless otherwise specified), pH 436

7.2. Neuromuscular junction sharp electrode (electrode resistance between 10-35 MΩ) 437

recordings were performed on muscles 6 and 7 of abdominal segments A2 and A3 in wandering 438

third-instar larvae (Kiragasi et al., 2017). Recordings were performed on an Olympus BX61 WI 439

microscope using a 40x/0.80 NA water-dipping objective. Recordings were acquired using an 440

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 18: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

18

Axoclamp 900A amplifier, Digidata 1440A acquisition system and pClamp 10.5 software 441

(Molecular Devices). Electrophysiological sweeps were digitized at 10 kHz and filtered at 1 kHz. 442

Data were analyzed using Clampfit (Molecular devices), MiniAnalysis (Synaptosoft), and Excel 443

(Microsoft) software. 444

Miniature excitatory postsynaptic potentials (mEPSPs) were recorded in the absence of 445

any stimulation and cut motor axons were stimulated to elicit excitatory postsynaptic potentials 446

(EPSPs). An ISO-Flex stimulus isolator (A.M.P.I.) was used to modulate the amplitude of 447

stimulatory currents. Intensity was adjusted for each cell, set to consistently elicit responses 448

from both neurons innervating the muscle segment, but avoiding overstimulation. Average 449

mEPSP, EPSP, and quantal content were calculated for each genotype. Muscle input 450

resistance (Rin) and resting membrane potential (Vrest) were monitored during each experiment. 451

Recordings were rejected if the Vrest was more depolarized than -60 mV, if the Rin was less than 452

5 MΩ, or if either measurement deviated by more than 10% during the course of the 453

experiment. Larvae were incubated with or without philanthotoxin-433 (PhTx; Sigma; 20 μM) 454

and resuspended in HL-3 for 10 min as described (Dickman and Davis, 2009; Frank et al., 455

2006). 456

The readily releasable pool (RRP) size was estimated by analyzing cumulative EPSC 457

amplitudes while recording using a two-electrode voltage clamp (TEVC) configuration as 458

described (Goel et al., 2019c). Muscles were clamped at -80 mV and EPSCs were evoked with 459

a 60 Hz, 60 stimulus train while recording in 3 mM Ca2+ HL-3. A line fit to the linear phase 460

(stimuli # 18–30) of the cumulative EPSC data was back-extrapolated to time 0. The RRP value 461

was estimated by determining the extrapolated EPSC value at time 0 and dividing this value by 462

the average mEPSC amplitude. 463

Data used in the variance-mean plot was obtained from TEVC recordings using an initial 464

0.5 mM Ca2+ concentration, which was subsequently increased to 1.5, 3.0, and 6.0 mM through 465

saline exchange using a peristaltic pump (Langer Instruments, BT100-2J). EPSC amplitudes 466

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 19: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

19

were monitored during the exchange, and 30 EPSC (0.5 Hz stimulation rate) events were 467

performed in each calcium condition (Li et al., 2018a). To obtain the variance-mean plot, the 468

variance (squared standard deviation) and mean (averaged evoked amplitude) were calculated 469

from the 30 EPSCs at each individual Ca2+ concentration. The variance was then plotted against 470

the mean for each specific calcium condition using MATLAB software (MathWorks, USA). One 471

additional data point, in which variance and mean are both theoretically at 0, was used for Ca2+-472

free saline. Data from these five conditions were fit with a standard parabola (variance = Q*I -473

I 2/N), where Q is the quantal size, I is the mean evoked amplitude (x-axis), and N is the 474

number of functional release sites. N, as a parameter of the standard parabola, was directly 475

calculated for each cell by best parabolic fit. 476

Failure analysis was performed in HL-3 solution containing 0.15 mM CaCl2. At this 477

extracellular Ca2+ concentration, approximately half of the stimulations evoked responses in the 478

muscle in wild type larvae. A total of 40 trials (stimulations) were performed at each NMJ in all 479

genotypes. Failure rate was obtained by dividing the total number of failures by the total number 480

of trials (40). Paired-pulse recordings were performed at a Ca2+ concentration of 0.3 mM 481

to assay facilitation (PPF) and 1.5 mM for depression (PPD). Following the first 482

stimulation, a second EPSC was evoked at an interstimulus interval of 16.66 msec. 483

Paired-pulse ratios were calculated as the difference between the second peak and the 484

maximum value between both peaks (corresponding to the starting point of the second 485

response) divided by the first amplitude. 486

487

Statistical analysis: Data were analyzed using GraphPad Prism (version 7.0) or Microsoft 488

Excel software (version 16.22). Sample values were tested for normality using the D’Agostino & 489

Pearson omnibus normality test which determined that the assumption of normality of the 490

sample distribution was not violated. Data were then compared using either a one-way ANOVA 491

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 20: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

20

and tested for significance using a Tukey’s multiple comparison test or using an unpaired 2-492

tailed Student’s t-test with Welch’s correction. All data are presented as mean +/-SEM; n 493

indicates sample number and p denotes the level of significance assessed as p<0.05 (*), p<0.01 494

(**), p<0.001 (***), p<0.0001 (****); ns=not significant. Statistics of all experiments are 495

summarized in Table S2. 496

497

ACKNOWLEDGEMENTS 498

We acknowledge the Developmental Studies Hybridoma Bank (Iowa, USA) for antibodies used 499

in this study, and the Bloomington Drosophila Stock Center for fly stocks (NIH P40OD018537). 500

We thank Giwoo Kim and Sarah Perry for technical contributions at early stages of this project. 501

P.G. was supported in part by a USC Provost Graduate Research Fellowship. This work was 502

supported by grants from the National Institutes of Health (NS111414 and NS091546) to D.D. 503

504

AUTHOR CONTRIBUTIONS 505

PG and DD designed the research; PG, SN, KC, CC, and XL performed the research; PG 506

analyzed the data; PG and DD wrote the paper. 507

508

DECLARATION OF INTERESTS 509

The authors declare no competing financial interests. 510

511

FIGURE LEGENDS 512

Figure 1: Biased innervation at the NMJ elicits distinct target-specific homeostatic 513

adaptations. (A) Schematic of a motor neuron innervating both muscle 6 and 7 at the 514

Drosophila larval NMJ. Biased innervation is achieved by overexpressing the cell adhesion 515

factor fasII specifically on muscle 6 using H94-Gal4 (M6>FasII: w;UAS-FasII/+;H94-Gal4/+). 516

Red outlines highlight the likely synaptic compartment in which the adaptation occurs. (B) 517

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 21: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

21

Representative images of muscle 6/7 NMJs immunostained with antibodies that recognize the 518

neuronal membrane (HRP) and synaptic vesicles (Synapsin; SYN) in wild type (w1118) and 519

M6>FasII. Note that while boutons labeled by SYN puncta are roughly equally split between 520

muscles 6 and 7 in wild type, M6>FasII causes biased innervation on muscle 6 at the expense 521

of muscle 7. (C) Representative electrophysiological traces of recordings from muscles 7 and 6 522

in wild type and M6>FasII NMJs. Note that while EPSP amplitudes are similar across all 523

muscles, mEPSPs are increased only on muscle 7 of M6>FasII. (D) Quantification of bouton 524

number, EPSP amplitude, mEPSP amplitude, quantal content and quantal content normalized 525

per bouton on muscle 7 in M6>FasII. All values are normalized to the values at wild type muscle 526

7. Enhanced mEPSP amplitude (shaded bar) implies reduced quantal content and no change in 527

quanta released per bouton. (E) Quantification of all values in (D) on muscle 6 of M6>FasII 528

normalized to wild type muscle 6 values. Note that estimated quantal content per bouton 529

(shaded bar) is significantly reduced. Error bars indicate ±SEM (n≥16; one-way ANOVA; Table 530

S2). **p<0.01; ***p<0.001; ****p<0.0001; ns=not significant. 531

532

Figure 2: Hyper-innervation elicits a homeostatic decrease in presynaptic release 533

probability. (A) Schematic illustrating a reduction in RRP size and functional release site 534

number on hyper-innervated muscle 6. (B) Failure analysis reveals no significant change in 535

failure rate on muscle 6 in M6>FasII, consistent with reduced release probability per bouton. (C) 536

Representative paired-pulse EPSC traces at 0.4 mM extracellular Ca2+ with an interstimulus 537

interval of 16.7 ms in the indicated genotypes. Increased paired-pulse facilitation (PPF) was 538

observed on hyper-innervated targets, consistent with reduced release probability. (D) 539

Quantification of PPF ratio (EPSC2/EPSC1). (E) Representative paired-pulse EPSC traces at 540

1.5 mM extracellular Ca2+ with an interstimulus interval of 16.7 ms in the indicated genotypes. 541

Enhanced paired-pulse depression (PPD) was observed on hyper-innervated targets, consistent 542

with reduced probability of release. (F) Quantification of PPD ratio (EPSC2/EPSC1). (G) 543

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 22: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

22

Representative EPSC recordings of 30 stimuli at 3 mM extracellular Ca2+ during a 60 Hz 544

stimulus train in the indicated genotypes. Insets represent the average cumulative EPSC plotted 545

as a function of time. A line fit to the 18th-30th stimuli was back-extrapolated to time 0. (H) 546

Estimated size of the RRP is unchanged on muscle 6 in M6>FasII compared with wild type, 547

suggesting reduced RRP per bouton. (I) Scatter plot EPSC distribution of recordings on muscle 548

6 from wild type and M6>FasII in the indicated extracellular Ca2+ concentrations. (J) Variance-549

mean plots for the indicated genotypes. Dashed lines are the best fit parabolas to the data 550

points. (K) Estimated number of functional release sites (N#) obtained from the variance-mean 551

plots in (J) showing no significant difference between the genotypes. Error bars indicate ±SEM 552

(n≥9; one-way ANOVA; Table S2). **p<0.01; ns=not significant. 553

554

Figure 3: Target-specific reductions in both active zone density and intensity at hyper-555

innervated NMJs. (A) Schematic illustrating a reduction in the number and intensity of active 556

zones at individual boutons on hyper-innervated muscle 6. (B) Representative images of muscle 557

6/7 NMJs in the indicated genotypes (wild type: cacsfGFP-N; M6>FasII: cacsfGFP-N;UAS-fasII/+;H94-558

Gal4/+) immunostained with antibodies against the active zone scaffold bruchpilot (BRP) and 559

GFP to label endogenously tagged Ca2+ channels (CAC). (C) Individual boutons from selected 560

areas (white rectangles) of NMJs stained with BRP and CAC in the indicated genotypes and 561

muscles. Note the reduction in the number and intensity of BRP and CAC puncta specifically on 562

muscle 6 in M6>FasII, while no change is observed on muscle 7 relative to wild type controls. 563

Quantification of BRP and CAC puncta number (D) and density (E) on muscle 6 in M6>FasII 564

normalized as a percentage of wild type muscle 6 values reveals a small but significant increase 565

in BRP puncta number, while BRP and CAC puncta density is significantly reduced on muscle 6 566

in M6>FasII. Quantification of BRP and CAC intensity (F) shows a significant reduction on 567

muscle 6 in M6>FasII, while the total fluorescence intensity of all BRP and CAC puncta 568

summed across the entire muscle 6 NMJ (G) is unchanged compared to wild type muscle 6. 569

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 23: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

23

Error bars indicate ±SEM (n≥13; one-way ANOVA; Table S2). *p<0.05; **p<0.01; ns=not 570

significant. 571

572

Figure 4: Distinct target-specific adaptations balance hyper-innervation and GluR loss. 573

(A) Schematic illustrating the dual manipulation used to both bias innervation and inhibit 574

GluRIIA expression specifically on muscle 6 (M6>FasII+GluRIIARNAi: w;Tub-FRT-STOP-FRT-575

Gal4,UAS-FLP,UAS-CD8-GFP;H94-Gal4,nSyb-Gal80;/UAS-FasII;UAS-GluRIIARNAi). (B) 576

Representative images of muscle 6/7 NMJs from the indicated genotypes immunostained with 577

anti-HRP and -GluRIIA. (C) Individual boutons from selected areas (white rectangles) of NMJs 578

shown in (B). Note the enhanced GluRIIA levels on hypo-innervated muscle 7 with loss on 579

hyper-innervated muscle 6. (D) Electrophysiological traces of recordings from muscles 7 and 6 580

in the indicated genotypes. EPSP amplitudes on muscle 7 and 6 in M6>FasII+GluRIIARNAi are 581

similar to wild-type values. (E) Quantification of bouton numbers, GluRIIA puncta intensity, 582

mEPSP amplitude, EPSP amplitude, and quantal content on muscle 7 in M6>FasII+GluRIIARNAi. 583

All values are normalized to baseline (M6>FasII muscle 7); no significant differences are 584

observed. (F) Quantification of all values in (D) on muscle 6 of M6>FasII+GluRIIARNAi 585

normalized to baseline values. Note that while GluRIIA levels and mEPSP amplitudes are 586

significantly reduced, EPSP amplitude remains unchanged because of a homeostatic increase 587

in quantal content, indicating PHP expression. Error bars indicate ±SEM (n≥8; Student’s t-test; 588

Table S2). ****p<0.0001; ns=not significant. 589

590

Figure S1: Postsynaptic GluR levels are enhanced at hypo-innervated targets but do not 591

change at hyper-innervated targets. (A) Schematic illustrating enhanced levels of both 592

GluRIIA- and GluRIIB-containing receptors on hypo-innervated muscle 7 in M6>FasII. (B) 593

Representative images of individual boutons at NMJs immunostained with antibodies that 594

recognize the postsynaptic GluR subunits GluRIIA, GluRIIB, and GluRIID. Quantification of the 595

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 24: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

24

indicated GluR puncta number (C) and intensity (D) normalized to wild type muscle 7 reveals a 596

significant reduction in GluR puncta number but increase in puncta intensity at hypo-innervated 597

muscle 7. (E) Schematic illustrating that no changes in postsynaptic GluR levels are induced at 598

hyper-innervated NMJs. (F) Representative images of individual boutons in the indicated 599

genotypes at muscle 6 NMJs immunostained with anti-GluRIIA, -GluRIIB, and -GluRIID. (G) 600

Quantification of GluR puncta number normalized to wild type muscle 6 demonstrates a 601

significant increase in M6>FasII. (H) The total sum intensity of individual GluRIIA, GluRIIB, and 602

GluRIID puncta on muscle 6 NMJs is unchanged in M6>FasII compared to wild type. Error bars 603

indicate ±SEM (n≥9; one-way ANOVA; Table S2). **p<0.01; ***p<0.001; ns=not significant. 604

605

Figure S2: No apparent changes in presynaptic function and the size, density or intensity 606

of active zones are observed at hypo-innervated NMJs. 607

(A) Representative images of individual boutons at muscle 7 NMJs in wild type and hypo-608

innervated M6>FasII immunostained with an antibody against the active zone scaffold BRP. (B) 609

Quantification of BRP puncta number, density, and intensity on muscle 7 in M6>FasII 610

normalized to wild-type values. A significant decrease in BRP puncta number but no change in 611

density is observed. No significant change in the total intensity of each individual BRP puncta is 612

observed (puncta sum intensity), while the total fluorescence intensity of all BRP puncta 613

summed across the entire muscle 7 NMJ is reduced in hypo-innervated targets. (C) 614

Representative EPSC traces of 30 stimuli at 3 mM extracellular Ca2+ during a 60 Hz stimulus 615

train recorded in two-electrode voltage clamp (TEVC) configuration in the indicated genotypes. 616

Graphical insets show the average cumulative EPSC plotted as a function of time. A line fit to 617

the 18th-30th stimuli was back-extrapolated to time 0. (D) The size of the RRP is reduced by a 618

proportion similar to the reduction in quantal content on muscle 7 in M6>FasII. (E) Scatter plot 619

EPSC distribution of recordings from muscle 7 in wild type (black circles) and M6>FasII (red 620

triangles) in the indicated extracellular Ca2+ concentrations. (F) Variance-mean plots for the 621

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 25: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

25

indicated genotypes. Variance was plotted against the mean amplitude of 30 EPSCs recorded 622

at 0.2 Hz from the Ca2+ concentrations detailed in (D). Lines are the best fit parabolas to the 623

data points. (G) Estimated number of functional release sites (N #) obtained from the variance-624

mean plots in (H) showing a significant reduction on muscle 7 of M6>FasII compared to wild 625

type muscle 7. Error bars indicate ±SEM (n≥9; one-way ANOVA; Table S2). **p<0.01; 626

***p<0.001; ns=not significant. 627

628

Figure S3: PHP can be rapidly expressed at hypo-innervated targets to balance synaptic 629

strength. (A) Schematics and representative traces illustrating the acute application of PhTx on 630

wild type and M6>FasII+GluRIIARNAi NMJs. Note that PhTx application reduces mEPSP 631

amplitudes on hypo-innervated targets, yet EPSP amplitudes remain similar to baseline values 632

because of a homeostatic increase in presynaptic neurotransmitter release, indicating robust 633

PHP expression. (B,C,D,E) Quantification of mEPSP amplitude, EPSP amplitude and quantal 634

content values normalized to baseline values (wild type muscle 7 in (B) and 6 in (D); 635

M6>FasII+GluRIIARNAi muscle 7 (C) or 6 (E)). Error bars indicate ±SEM (n≥8; Student’s t-test; 636

Table S2). ****p<0.0001; ns=not significant. 637

638 REFERENCES 639 640 Akbergenova, J., K.L. Cunningham, Y.V. Zhang, S. Weiss, and J.T. Littleton. 2018. 641

Characterization of developmental and molecular factors underlying release 642 heterogeneity at Drosophila synapses. Elife. 7:e38268. 643

Allan, D.W., S.E. St Pierre, I. Miguel-Aliaga, and S. Thor. 2003. Specification of neuropeptide 644 cell identity by the integration of retrograde BMP signaling and a combinatorial 645 transcription factor code. Cell. 113:73-86. 646

Allan, D.W., and S. Thor. 2015. Transcriptional selectors, masters, and combinatorial codes: 647 regulatory principles of neural subtype specification. Wiley Interdiscip Rev Dev Biol. 648 4:505-528. 649

Atwood, H.L., and S. Karunanithi. 2002. Diversification of synaptic strength: presynaptic 650 elements. Nat Rev Neurosci. 3:497-516. 651

Banovic, D., O. Khorramshahi, D. Owald, C. Wichmann, T. Riedt, W. Fouquet, R. Tian, S.J. 652 Sigrist, and H. Aberle. 2010. Drosophila neuroligin 1 promotes growth and postsynaptic 653 differentiation at glutamatergic neuromuscular junctions. Neuron. 66:724-738. 654

Béïque, J.C., Y. Na, D. Kuhl, P.F. Worley, and R.L. Huganir. 2011. Arc-dependent synapse-655 specific homeostatic plasticity. PNAS. 108:816-821. 656

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 26: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

26

Bohme, M.A., C. Beis, S. Reddy-Alla, E. Reynolds, M.M. Mampell, A.T. Grasskamp, J. 657 Lutzkendorf, D.D. Bergeron, J.H. Driller, H. Babikir, F. Gottfert, I.M. Robinson, C.J. 658 O'Kane, S.W. Hell, M.C. Wahl, U. Stelzl, B. Loll, A.M. Walter, and S.J. Sigrist. 2016. 659 Active zone scaffolds differentially accumulate Unc13 isoforms to tune Ca(2+) channel-660 vesicle coupling. Nat Neurosci. 19:1311-1320. 661

Bohme, M.A., A.W. McCarthy, A.T. Grasskamp, C.B. Beuschel, P. Goel, B. Jusyte, D. Laber, S. 662 Huang, U. Rey, A.G. Petzold, M. Lehmann, F. Goettfert, P. Haghighi, S.W. Hell, D. 663 Owald, D. Dickman, S.J. Sigrist, and A.M. Walter. 2019. Rapid active zone remodeling 664 consolidates presynaptic potentiation. Nat Commun. 10:1085. 665

Branco, T., and K. Staras. 2009. The probability of neurotransmitter release: variability and 666 feedback control at single synapses. . Nat Rev Neurosci. 10:373-383. 667

Calderó, J., D. Prevette, X. Mei, R.A. Oakley, L. Li, C. Milligan, L. Houenou, M. Burek, and R.W. 668 Oppenheim. 1998. Peripheral target regulation of the development and survival of spinal 669 sensory and motor neurons in the chick embryo. J Neurosci. 18:356-370. 670

Chowdhury, D., and J.W. Hell. 2018. Homeostatic synaptic scaling: molecular regulators of 671 synaptic AMPA-type glutamate receptors. F1000Res. 7:234. 672

Clements, J.D., and R.A. Silver. 2000. Unveiling synaptic plasticity: a new graphical and 673 analytical approach. Trends Neurosci. 23:105-113. 674

Daniels, R.W., C.A. Collins, M.V. Gelfand, J. Dant, E.S. Brooks, D.E. Krantz, and A. DiAntonio. 675 2004. Increased expression of the Drosophila vesicular glutamate transporter leads to 676 excess glutamate release and a compensatory decrease in quantal content. J Neurosci. 677 24:10466-10474. 678

Davis, G.W., and C.S. Goodman. 1998. Synapse-specific control of synaptic efficacy at the 679 terminals of a single neuron. Nature. 392:82-86. 680

Davis, G.W., and M. Muller. 2015. Homeostatic control of presynaptic neurotransmitter release. 681 Annu Rev Physiol. 77:251-270. 682

DiAntonio, A. 2006. Glutamate receptors at the Drosophila neuromuscular junction. Intl. Rev. 683 Neurobiol. 75:165-179. 684

DiAntonio, A., S.A. Petersen, M. Heckmann, and C.S. Goodman. 1999. Glutamate receptor 685 expression regulates quantal size and quantal content at the Drosophila neuromuscular 686 junction. J Neurosci. 19:3023-3032. 687

DiAntonio, A., and T.L. Schwarz. 1994. The effect on synaptic physiology of synaptotagmin 688 mutations in Drosophila. Neuron. 12:909-920. 689

Dickman, D.K., and G.W. Davis. 2009. The schizophrenia susceptibility gene dysbindin controls 690 synaptic homeostasis. Science. 326:1127-1130. 691

Diering, G.H., and R.L. Huganir. 2018. The AMPA Receptor Code of Synaptic Plasticity. 692 Neuron. 100:314-329. 693

Fekete, A., Y. Nakamura, Y.M. Yang, S. Herlitze, M.D. Mark, D.A. DiGregorio, and L.Y. Wang. 694 2019. Underpinning heterogeneity in synaptic transmission by presynaptic ensembles of 695 distinct morphological modules. Nat Commun. 10:826. 696

Frank, C.A., T.D. James, and M. Müller. 2020. Homeostatic control of Drosophila 697 neuromuscular junction function. Synapse. 74:e22133. 698

Frank, C.A., M.J. Kennedy, C.P. Goold, K.W. Marek, and G.W. Davis. 2006. Mechanisms 699 underlying the rapid induction and sustained expression of synaptic homeostasis. 700 Neuron. 52:663-677. 701

Goel, P., D.D. Bergeron, M. Bohme, L. Nunnelly, M. Lehmann, C. Buser, A.M. Walter, S.J. 702 Sigrist, and D.K. Dickman. 2019a. Homeostatic scaling of active zone scaffolds 703 maintains global synaptic strength. J Cell Biol. 218:1706-1724. 704

Goel, P., and D. Dickman. 2018. Distinct homeostatic modulations stabilize reduced 705 postsynaptic receptivity in response to presynaptic DLK signaling. Nat Commun. 9:1856. 706

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 27: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

27

Goel, P., M. Khan, S. Howard, G. Kim, B. Kiragasi, K. Kikuma, and D. Dickman. 2019b. A 707 Screen for Synaptic Growth Mutants Reveals Mechanisms That Stabilize Synaptic 708 Strength. J Neurosci. 39:4051-4065. 709

Goel, P., X. Li, and D. Dickman. 2017. Disparate Postsynaptic Induction Mechanisms Ultimately 710 Converge to Drive the Retrograde Enhancement of Presynaptic Efficacy. Cell Rep. 711 21:2339-2347. 712

Goel, P., X. Li, and D. Dickman. 2019c. Estimation of the Readily Releasable Synaptic Vesicle 713 Pool at the Drosophila Larval Neuromuscular Junction. Bio Protoc. 9:e3127. 714

Graf, E.R., R.W. Daniels, R.W. Burgess, T.L. Schwarz, and A. DiAntonio. 2009. Rab3 715 dynamically controls protein composition at active zones. Neuron. 64:663-677. 716

Gratz, S.J., P. Goel, J.J. Bruckner, R.X. Hernandez, K. Khateeb, G. Macleod, D. Dickman, and 717 K.M. O'Connor-Giles. 2019. Endogenous tagging reveals differential regulation of Ca2+ 718 channels at single AZs during presynaptic homeostatic potentiation and depression. J 719 Neurosci. 39:2416-2429. 720

Grillo, F.W., G. Neves, A. Walker, G. Vizcay-Barrena, R.A. Fleck, T. Branco, and J. Burrone. 721 2018. A Distance-Dependent Distribution of Presynaptic Boutons Tunes Frequency-722 Dependent Dendritic Integration. Neuron. 99:275-282. 723

Guerrero, G., D.F. Reiff, G. Agarwal, R.W. Ball, A. Borst, C.S. Goodman, and E.Y. Isacoff. 724 2005. Heterogeneity in synaptic transmission along a Drosophila larval motor axon. Nat 725 Neurosci. 8:1188-1196. 726

Han, T.H., P. Dharkar, M.L. Mayer, and M. Serpe. 2015. Functional reconstitution of Drosophila 727 melanogaster NMJ glutamate receptors. PNAS. 112:6182-6187. 728

Henry, F.E., A.J. McCartney, R. Neely, A.S. Perez, C.J. Carruthers, E.L. Stuenkel, K. Inoki, and 729 M.A. Sutton. 2012. Retrograde changes in presynaptic function driven by dendritic 730 mTORC1. J Neurosci. 32:17128-17142. 731

Hou, Q., D. Zhang, L. Jarzylo, R.L. Huganir, and H.Y. Man. 2008. Homeostatic regulation of 732 AMPA receptor expression at single hippocampal synapses. PNAS. 105:775-780. 733

Huntwork, S., and J.T. Littleton. 2007. A complexin fusion clamp regulates spontaneous 734 neurotransmitter release and synaptic growth. Nat Neurosci. 10:1235-1237. 735

Jakawich, S.K., H.B. Nasser, M.J. Strong, A.J. McCartney, A.S. Perez, N. Rakesh, C.J. 736 Carruthers, and M.A. Sutton. 2010. Local presynaptic activity gates homeostatic 737 changes in presynaptic function driven by dendritic BDNF synthesis. Neuron. 68:1143-738 1158. 739

Jia, H., N.L. Rochefort, X. Chen, and A. Konnerth. 2010. Dendritic organization of sensory input 740 to cortical neurons in vivo. Nature. 464:1307-1312. 741

Katz, P.S., M.D. Kirk, and C.K. Givind. 1993. Facilitation and Depression at Different Branches 742 of the Same Motor Axon: Evidence for Presynaptic Differences in Release. J Neurosci. 743 13:3075-3089. 744

Katz, Y., V. Menon, D.A. Nicholson, Y. Geinisman, W.L. Kath, and N. Spruston. 2009. Synapse 745 distribution suggests a two-stage model of dendritic integration in CA1 pyramidal 746 neurons. Neuron. 63:171-177. 747

Kaufmann, N., J. DeProto, R. Ranjan, H. Wan, and D. Van Vactor. 2002. Drosophila liprin-alpha 748 and the receptor phosphatase Dlar control synapse morphogenesis. Neuron. 34:27-38. 749

Keck, T., G.B. Keller, R.I. Jacobsen, U.T. Eysel, T. Bonhoeffer, and M. Hubener. 2013. 750 Synaptic scaling and homeostatic plasticity in the mouse visual cortex in vivo. Neuron. 751 80:327-334. 752

Kikuma, K., X. Li, D. Kim, D. Sutter, and D.K. Dickman. 2017. Extended Synaptotagmin 753 Localizes to Presynaptic ER and Promotes Neurotransmission and Synaptic Growth in 754 Drosophila. Genetics. 207:993-1007. 755

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 28: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

28

Kiragasi, B., J. Wondolowski, Y. Li, and D.K. Dickman. 2017. A Presynaptic Glutamate Receptor 756 Subunit Confers Robustness to Neurotransmission and Homeostatic Potentiation. Cell 757 Rep. 19:2694-2706. 758

Klassen, M.P., Y.E. Wu, C.I. Maeder, I. Nakae, J.G. Cueva, E.K. Lehrman, M. Tada, K. Gengyo-759 Ando, G.J. Wang, M. Goodman, S. Mitani, K. Kontani, T. Katada, and K. Shen. 2010. An 760 Arf-like Small G Protein, ARL-8, Promotes the Axonal Transport of Presynaptic Cargoes 761 by Suppressing Vesicle Aggregation. Neuron. 66:710-723. 762

Lee, J., Z. Guan, Y. Akbergenova, and J.T. Littleton. 2013. Genetic analysis of synaptotagmin 763 C2 domain specificity in regulating spontaneous and evoked neurotransmitter release. J 764 Neurosci. 33:187-200. 765

Letellier, M., F. Levet, O. Thoumine, and Y. Goda. 2019. Differential role of pre- and 766 postsynaptic neurons in the activity-dependent control of synaptic strengths across 767 dendrites. PLoS Biol. 17:e2006223. 768

Li, X., P. Goel, C. Chen, V. Angajala, X. Chen, and D. Dickman. 2018a. Synapse-specific and 769 compartmentalized expression of presynaptic homeostatic potentiation. Elife. 7:e34338. 770

Li, X., P. Goel, J. Wondolowski, J. Paluch, and D. Dickman. 2018b. A Glutamate Homeostat 771 Controls the Presynaptic Inhibition of Neurotransmitter Release. Cell Rep. 23:1716-772 1727. 773

Littleton, J.T., H.J. Bellen, and M.S. Perin. 1993. Expression of synaptotagmin in Drosophila 774 reveals transport and localization of synaptic vesicles to the synapse. Development. 775 118:1077-1088. 776

Marqués, G., H. Bao, T.E. Haerry, M.J. Shimell, P. Duchek, B. Zhang, and M.B. O'Connor. 777 2002. The Drosophila BMP type II receptor Wishful Thinking regulates neuromuscular 778 synapse morphology and function. Neuron. 33:529-543. 779

O’Rourke, N.A., N.C. Weiler, K.D. Micheva, and S.J. Smith. 2012. Deep Molecular Diversity of 780 Mammalian Synapses: Why It Matters and How to Measure It. Nat Rev Neurosci. 781 13:365-379. 782

Palma, E., M. Inghilleri, L. Conti, C. Deflorio, V. Frasca, A. Manteca, F. Pichiorric, C. Roseti, G. 783 Torchia, C. Limatola, F. Grassi, and R. Miledi. 2011. Physiological characterization of 784 human muscle acetylcholine receptors from ALS patients. PNAS. 108:20184-20188. 785

Park, H., and M.M. Poo. 2013. Neurotrophin regulation of neural circuit development and 786 function. Nat Rev Neurosci. 14:7-23. 787

Perry, S., Y. Han, A. Das, and D.K. Dickman. 2017. Homeostatic plasticity can be induced and 788 expressed to restore synaptic strength at neuromuscular junctions undergoing ALS-789 related degeneration. Human Mol Genet. 26:4153-4167. 790

Qin, G., T. Schwarz, R.J. Kittel, A. Schmid, T.M. Rasse, D. Kappei, E. Ponimaskin, M. 791 Heckmann, and S.J. Sigrist. 2005. Four different subunits are essential for expressing 792 the synaptic glutamate receptor at neuromuscular junctions of Drosophila. J Neurosci. 793 25:3209-3218. 794

Rasse, T.M., W. Fouquet, A. Schmid, R.J. Kittel, S. Mertel, C.B. Sigrist, M. Schmidt, A. 795 Guzman, C. Merino, G. Qin, C. Quentin, F.F. Madeo, M. Heckmann, and S.J. Sigrist. 796 2005. Glutamate receptor dynamics organizing synapse formation in vivo. Nat Neurosci. 797 8:898-905. 798

Regehr, W.G. 2012. Short-term presynaptic plasticity.Cold Spring Harb Perspect Biol. 799 4:a005702. 800

Rich, M.M., and J.W. Lichtman. 1989. In vivo visualization of pre- and postsynaptic changes 801 during synapse elimination in reinnervated mouse muscle. J Neurosci. 5:1781-1805. 802

Rosenmund, C., and C.F. Stevens. 1996. Definition of the readily releasable pool of vesicles at 803 hippocampal synapses. Neuron. 16:1197-1207. 804

Rudolph, S., M.C. Tsai, H. von Gersdorff, and J.I. Wadiche. 2015. The ubiquitous nature of 805 multivesicular release. Trends Neurosci. 38:428-438. 806

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 29: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

29

Saitoe, M., T.L. Schwarz, J.A. Umbach, C.B. Gundersen, and Y. Kidokoro. 2001. Absence of 807 junctional glutamate receptor clusters in Drosophila mutants lacking spontaneous 808 transmitter release. Science. 293:514-517. 809

Salpeter, M.M., and R. Harris. 1983. Distribution and turnover rate of acetylcholine receptors 810 throughout the junction folds at a vertebrate neuromuscular junction. J Cell Biol. 811 96:1781-1785. 812

Schotzinger, R.J., and S.C. Landis. 1990. Acquisition of cholinergic and peptidergic properties 813 by sympathetic innervation of rat sweat glands requires interaction with normal target. 814 Neuron. 5:91-100. 815

Stewart, B.A., H.L. Atwood, J.J. Renmger, J. Wang, and C.F. Wu. 1994. Improved stability of 816 Drosophila larval neuromuscular prepa- rations in haemolymph-like physiological 817 solutions. J Comp Physiol. 175:179-191. 818

Stuart, G.J., and N. Spruston. 2015. Dendritic integration:60 years of progress. Nat Neurosci. 819 18:1713-1721. 820

Sutton, M.A., H.T. Ito, P. Cressy, C. Kempf, J.C. Woo, and E.M. Schuman. 2006. Miniature 821 neurotransmission stabilizes synaptic function via tonic suppression of local dendritic 822 protein synthesis. Cell. 125:785-799. 823

Tripodi, M., J.F. Evers, A. Mauss, M. Bate, and M. Landgraf. 2008. Structural homeostasis: 824 compensatory adjustments of dendritic arbor geometry in response to variations of 825 synaptic input. PLoS Biol. 6:e260. 826

Turrigiano, G. 2012. Homeostatic Synaptic Plasticity: Local and Global Mechanisms for 827 Stabilizing Neuronal Function. Cold Spring Harb. Perspect. Biol. 4:a005736. 828

Turrigiano, G.G. 2008. The self-tuning neuron: synaptic scaling of excitatory synapses. Cell. 829 135:422-435. 830

Turrigiano, G.G., and S.B. Nelson. 2004. Homeostatic plasticity in the developing nervous 831 system. Nat Rev Neurosci. 5:97-107. 832

Vitureira, N., and Y. Goda. 2013. The interplay between Hebbian and homeostatic synaptic 833 plasticity. J Cell Biol. 203:175. 834

Vitureira, N., M. Letellier, I.J. White, and Y. Goda. 2011. Differential control of presynaptic 835 efficacy by postsynaptic N-cadherin and β-catenin. Nat Neurosci. 15:81-89. 836

Vukoja, A., U. Rey, A.G. Petzoldt, C. Ott, D. Vollweiter, C. Quentin, D. Puchkov, E. Reynolds, 837 M. Lehmann, S. Hohensee, S. Rosa, R. Lipowsky, S.J. Sigrist, and V. Haucke. 2018. 838 Presynaptic Biogenesis Requires Axonal Transport of Lysosome-Related Vesicles. 839 Neuron. 99:1216-1232 e1217. 840

Weyhersmuller, A., S. Hallermann, N. Wagner, and J. Eilers. 2011. Rapid active zone 841 remodeling during synaptic plasticity. J Neurosci. 31:6041-6052. 842

Yuan, Q., Y. Xiang, Z. Yan, C. Han, L.Y. Jan, and Y.N. Jan. 2011. Light-Induced Structural and 843 Functional Plasticity in Drosophila Larval Visual System. Science. 333:1458-1462. 844

845

.CC-BY-NC-ND 4.0 International licensewas not certified by peer review) is the author/funder. It is made available under aThe copyright holder for this preprint (whichthis version posted April 7, 2020. . https://doi.org/10.1101/2020.04.07.029108doi: bioRxiv preprint

Page 30: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

1

B

C

M7 M6

A

M7 M6

M7 M6M7 M6HRP

M7 M6

10 μm

SYN

M6>FasIIwild type

M7 M6 M7 M6

M7 M6HRP SYN

muscle 7 (M6>FasII) muscle 6 (M6>FasII)D E

**

**

ns

mEPSP QC QC/bouton

**

nsns

8 m

V

50 msec

1 m

V

250 msec

***

ns

% w

ild ty

pe M

6

0

50

150

100

bouton # EPSP QC QC/boutonmEPSP

****

ns

% w

ild ty

pe M

7

0

50

150

100

bouton # EPSP

Page 31: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

0.4

mM

Ca2+

B

K

varia

nce

(nA

2 )

0 100 200

10

20

40

30

EPSC (nA)

0

calcium concentration (mM)0.5 1.5 3

50

100

150

0

200

6

EPSC

(nA

)

M6M6

A wild type M6>FasII

estim

ated

RR

P si

ze

1000

1300

500

0

10 n

A 20 msec

50 n

A

20 msec

failu

re ra

te

(%)

0

50

100

N #

0

200

400

600

800 ns

D

0

100

50

% E

PSC

2/EP

SC1

(PPD

)

F

I

C

E

1.5

mM

Ca2+

J

H

100

nA

100 ms

cum

. EPS

C1000

2000

3000

cum

. EPS

C

1000

2000

3000G

ns

**

**

ns

M6>FasII M6

wild type M6

0

100

50

% E

PSC

2/EP

SC1

(PPF

)

150

2

Page 32: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

M7 M6B

A

M7 M6

M6 M6

wild type M6>FasII

BRP

D E

BRP

CAC

M7 M6BRP

M6 M7 M6M7

M7 M6

10 μm

CACCAC

0

100

50

M6

punc

ta d

ensi

ty

(% W

T) ** **

F

0

100

50

M6

punc

ta in

tens

ity

(% W

T)

** *

G

0

100

50

M6

tota

l int

ensi

ty

(% W

T)

ns ns

BRP CAC BRP CAC BRP CAC BRP CAC

C

0

130

100

50

M6

punc

ta #

(% W

T)

*ns

5 μm

3

130 130130

Page 33: 7 8 9 11 Drosophila - bioRxiv.org1 1 Distinct target-specific mechanisms homeostatically stabilize transmission at pre-and 2 post-synaptic compartments 3 4 Pragya Goel1,2, Samantha

M7 M6B M7 M6M7 M6

A

M7 M6

wild type M6>FasII+GluRIIARNAi

E F

8 m

V

1 m

V

250 msec

50 msec

muscle 7 (M6>FasII+GluRIIARNAi) muscle 6 (M6>FasII+GluRIIARNAi)

M7 M6 M7 M6

vGlutHRP

GluRIIA

vGlutHRP

M7 M6 M7 M6

Glu

RIIA

HRP

C

D

150

50

ns ns

% b

asel

ine

M7

(M6>

FasI

I)

0

100

200

M7bouton # GluRIIAintensity

ns nsns

mEPSP QCEPSP

ns

****% b

asel

ine

M6

(M6

> Fa

sII)

0

50

150

100

M6 bouton #GluRIIAintensity

****

****

ns

mEPSP QCEPSP

200

10 μmGluRIIA

2 μm

4