61
Copyright © 2020 the authors Research Articles: Development/Plasticity/Repair Activity dependent and independent determinants of synaptic size diversity https://doi.org/10.1523/JNEUROSCI.2181-19.2020 Cite as: J. Neurosci 2020; 10.1523/JNEUROSCI.2181-19.2020 Received: 10 September 2019 Revised: 4 February 2020 Accepted: 13 February 2020 This Early Release article has been peer-reviewed and accepted, but has not been through the composition and copyediting processes. The final version may differ slightly in style or formatting and will contain links to any extended data. Alerts: Sign up at www.jneurosci.org/alerts to receive customized email alerts when the fully formatted version of this article is published.

Activity dependent and independent determinants of synaptic size … · 2020. 2. 27. · 1 JN-RM-2181-19R2 Activity dependent and independent determinants of synaptic size diversity

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • Copyright © 2020 the authors

    Research Articles: Development/Plasticity/Repair

    Activity dependent and independentdeterminants of synaptic size diversity

    https://doi.org/10.1523/JNEUROSCI.2181-19.2020

    Cite as: J. Neurosci 2020; 10.1523/JNEUROSCI.2181-19.2020

    Received: 10 September 2019Revised: 4 February 2020Accepted: 13 February 2020

    This Early Release article has been peer-reviewed and accepted, but has not been throughthe composition and copyediting processes. The final version may differ slightly in style orformatting and will contain links to any extended data.

    Alerts: Sign up at www.jneurosci.org/alerts to receive customized email alerts when the fullyformatted version of this article is published.

  • 1

    JN-RM-2181-19R2 Activity dependent and independent determinants of synaptic size diversity Liran Hazan1 and Noam E. Ziv1,2 1Technion Faculty of Medicine, Rappaport Institute and Network Biology Research Laboratories, Fishbach Building, Technion City, Haifa, 32000, Israel. 2 Corresponding author: Noam E. Ziv Technion Faculty of Medicine and Network Biology Research Laboratories, Fishbach Building Technion city Haifa 32000, Israel [email protected] Abbreviated title: Determinants of synaptic size diversity Number of Pages: 48 Number of Figures: 12 Number of words:

    Abstract: 249 Introduction: 648 Discussion: 1498

    The authors declare no competing financial interests Acknowledgements We are grateful to Tamar Galateanu, Leonid Odesski, Ayub Bolous, Tamar Ziv and the Smoler Proteomics Center, as well as members of the Ciechanover lab for their invaluable assistance. We are also grateful to Naama Brenner, Omri Barak and Aseel Shomar for many helpful discussions. We are particularly grateful to an anonymous reviewer for suggesting the formulation of the Kesten process as a non-linear Langevin process. This work was supported by funding from the Israel Science Foundation (1175/14; 1470/18), The Rappaport Institute, the Allen and Jewel Prince Center for Neurodegenerative Disorders of the Brain and the state of Lower-Saxony and the Volkswagen Foundation, Hannover, Germany. 1 2

  • 2

    Abstract 3

    The extraordinary diversity of excitatory synapse sizes is commonly attributed to activity- 4

    dependent processes that drive synaptic growth and diminution. Recent studies also point to 5

    activity-independent size fluctuations, possibly driven by innate synaptic molecule dynamics, 6

    as important generators of size diversity. To examine the contributions of activity-dependent 7

    and independent processes to excitatory synapse size diversity, we studied glutamatergic 8

    synapse size dynamics and diversification in cultured rat cortical neurons (both sexes), 9

    silenced from plating. We found that in networks with no history of activity whatsoever, 10

    synaptic size diversity was no less extensive than that observed in spontaneously active 11

    networks. Synapses in silenced networks were larger, size distributions were broader, yet 12

    these were rightward-skewed and similar in shape when scaled by mean synaptic size. 13

    Silencing reduced the magnitude of size fluctuations and weakened constraints on size 14

    distributions, yet these were sufficient to explain synaptic size diversity in silenced networks. 15

    Model-based exploration followed by experimental testing indicated that silencing- 16

    associated changes in innate molecular dynamics and fluctuation characteristics might 17

    negatively impact synaptic persistence, resulting in reduced synaptic numbers. This, in turn, 18

    would increase synaptic molecule availability, promote synaptic enlargement, and ultimately 19

    alter fluctuation characteristics. These findings suggest that activity-independent size 20

    fluctuations are sufficient to fully diversify glutamatergic synaptic sizes, with activity- 21

    dependent processes primarily setting the scale rather than the shape of size distributions. 22

    Moreover, they point to reciprocal relationships between synaptic size fluctuations, size 23

    distributions and synaptic numbers mediated by the innate dynamics of synaptic molecules 24

    as they move in, out and between synapses. 25

    26

    27

    28

    29

    30

    31

    32

    33

    34

  • 3

    Significance Statement 35

    Sizes of glutamatergic synapses vary tremendously, even when formed on the same neuron. 36

    This diversity is commonly thought to reflect the outcome of activity-dependent forms of 37

    synaptic plasticity, yet activity-independent processes might also play some part. Here we 38

    show that in neurons with no history of activity whatsoever, synaptic sizes are no less 39

    diverse. We show that this diversity is the product of activity-independent size fluctuations, 40

    which are sufficient to generate a full repertoire of synaptic sizes at correct proportions. By 41

    combining modeling and experimentation we expose reciprocal relationships between size 42

    fluctuations, synaptic sizes and synaptic counts, and show how these phenomena might be 43

    connected through the dynamics of synaptic molecules as they move in, out and between 44

    synapses. 45

    46

  • 4

    Introduction 47

    Properties of mammalian glutamatergic synapses can be extremely diverse. This diversity 48

    is manifested in the broad distributions of many functional and morphological properties, 49

    such as postsynaptic current amplitude, dendritic spine volume and postsynaptic density 50

    (PSD) area. Such distributions are not only broad but also rightward skewed and heavy- 51

    tailed, and are often described as log-normal (e.g. Murthy et al., 1997; Harms and Craig, 52

    2005; Harms et al., 2005; Song et al., 2005; Arellano et al., 2007; Minerbi et al., 2009; Lefort 53

    et al., 2009; Loewenstein et al., 2011; Ikegaya et al., 2013; Keck et al., 2013; Statman et al. 54

    2014; Zhang et al., 2015; Cossell et al., 2015; Santuy et at., 2018; Ishii et al., 2018; Sammons 55

    et al., 2018; Sakamoto et al., 2018; Masch et al., 2018; Wegner et al., 2018; Hobbiss et al., 56

    2018; reviewed in Barbour et al., 2007; Buzsáki and Mizuseki, 2014; Scheler, 2017). Such 57

    distributions, reflecting of a majority of weak/small synapses and a diminishing tail of 58

    increasingly stronger/larger synapses, were suggested to optimize storage capacity, 59

    neuronal firing rates and long-distance information transfer and thus impart important 60

    properties to neuronal networks (Song et al., 2005; Barbour et al., 2007; Lefort et al., 2009; 61

    Ikegaya et al., 2013; Buzsáki and Mizuseki, 2014; Scheler, 2017; Humble et al., 2019). 62

    The diversity of synaptic sizes reflected in these distributions is commonly assumed to 63

    result from activity-dependent synaptic plasticity that drives the growth of some synapses 64

    and the downsizing of others. Moreover, the skewed shape of these distributions is assumed 65

    to reflect the cumulative outcome of such processes (van Rossum et al., 2000; Song et al., 66

    2005; Lefort et al., 2009; Gilson and Fukai, 2011; Zheng et al., 2013; Buzsáki and Mizuseki, 67

    2014; Effenberger et al., 2015; Scheler, 2017; Uzan et al., 2018). Somewhat unexpectedly, 68

    however, synaptic size diversity does not seem to be markedly reduced in animals that 69

    develop in the complete absence of synaptic transmission (Sando et al., 2017; see also Sigler 70

    et al., 2017; Lu et al., 2013), echoing findings of earlier cell-culture studies (e.g. Harms and 71

    Craig, 2005; Harms et al., 2005; Yasumatsu et al., 2008). 72

    Longitudinal in-vitro and in-vivo imaging reveals that sizes of individual glutamatergic 73

    synapses fluctuate considerably over time scales of hours and days (e.g. Yasumatsu et al., 74

    2008; Minerbi et al., 2009; Loewenstein et al., 2011; Kaufman et al., 2012; Fisher-Lavie and 75

    Ziv 2013; Cane et al., 2014; Ishii et al., 2018; reviewed in Ziv and Brenner, 2018). Importantly, 76

    such fluctuations persist following abrupt suppressions of network activity or synaptic 77

    transmission (Yasumatsu et al., 2008; Minerbi et al., 2009; Dvorkin and Ziv, 2016). These 78

  • 5

    intrinsic fluctuations, probably driven by the innate dynamics of synaptic molecules – 79

    binding, unbinding and turnover (Kasai, 2010; Fisher-Lavie and Ziv, 2014; Shomar et al., 80

    2017; Triesch et al., 2018) – were suggested to drive synaptic size diversification and 81

    determine the shape and scale of size distributions (Yasumatsu et al., 2008; Minerbi et al., 82

    2009; Kasai, 2010, Loewenstein et al., 2011; Kaufman et al., 2012; Statman et al., 2014; 83

    Shomar et al., 2017; Ishii et al., 2018; Ziv and Brenner 2018; Humble et al., 2019). It remains 84

    unclear, however, if intrinsic size fluctuations are indeed sufficient to give rise to a full 85

    repertoire of synaptic sizes and at the right proportions. 86

    Given the significance attributed to synaptic ‘weights’, understanding the fundamental 87

    forces that drive synaptic size diversification and define proportions of differently sized 88

    synapses would seem to be important. We therefore set out to examine the contributions of 89

    activity-dependent and -independent processes to synaptic size diversification. We first 90

    asked whether intrinsic, activity-independent size fluctuations are sufficient to give rise to a 91

    full repertoire of synaptic sizes. We then examined how innate processes that drive intrinsic 92

    fluctuations might set the shape and scale of synaptic size distributions and define the sizes 93

    of synaptic populations. Finally, we studied interrelationships between these processes and 94

    how these are affected by network activity. 95

    96

  • 6

    Materials and Methods 97

    Experimental design and statistical analyses 98

    Due to the long duration of each experiment (~1 week), data could not be collected as 99

    age-matched pairs from the same cell culture preparations. Thus, individual experiments 100

    typically came from separate cell culture preparations, resulting in extensive sampling of 101

    preparations, reducing sensitivity to this source of variability. Comparisons between 102

    treatments were typically based on tens (neurons) or thousands (synapses) of data points 103

    except for biochemical and proteomic studies (Fig. 8) which were based on 2(7) and 2(4) 104

    separate experiments (replicates), respectively. 105

    Statistical tests used were based on minimal assumptions. Specific tests used and 106

    significance values are provided in the main text and figure legends. Error bars are either 107

    Standard Deviation or Standard Error of the Mean (SEM) as indicated in legends. 108

    All software used for simulations of Fig. 5 and 6 (Visual Basic for Applications) and Figs. 109

    7,9,10 (C code), data used to create all plots and the full proteomic dataset is available upon 110

    request. 111

    112

    Cell culture 113

    Primary cultures of rat cortical neurons were prepared as described previously (Minerbi 114

    et al., 2009) using a protocol approved by the Technion committee for the supervision of 115

    animal experiments (IL-116-08-71). Briefly, cortices of newborn (1 day-old) Wistar rats 116

    (either sex; Charles River, UK) were dissected, dissociated by trypsin treatment followed by 117

    trituration using a siliconized Pasteur pipette. A total of 1–1.5x106 cells were then plated on 118

    thin-glass multielectrode array (MEA) dishes (MultiChannelSystems—MCS, Germany), pre- 119

    coated with polyethylenimine (Sigma) to facilitate cell adherence. The preparations were 120

    then transferred to a humidified tissue culture incubator and maintained at 37°C in a gas 121

    mixture of 5% CO2, 95% air, and grown in medium containing minimal essential medium 122

    (MEM, Sigma), 25 mg/l insulin (Sigma), 20 mM glucose (Sigma), 2 mM L-glutamine (Sigma), 5 123

    mg/ml gentamycin sulfate (Sigma) and 10% NuSerum (Becton Dickinson Labware). 7 days 124

    after plating, half of the culture medium was replaced with feeding medium similar to the 125

    medium described above, but devoid of NuSerum, containing a lower L-glutamine 126

    concentration (0.5 mM) and 2% B-27 supplement (Invitrogen). About half of the medium 127

    was then replaced three to four times a week. 128

  • 7

    129

    DNA constructs, lentivirus production and transduction 130

    A third generation lentiviral expression system was used to introduce exogenous DNA 131

    into rat cortical neurons. The vector used here {FU(PSD-95:EGFP)W} was described in detail 132

    in Minerbi et al., (2009). Lentiviral particles were produced using a mixture of the expression 133

    vector and the packaging vector mix of the ViraPower plasmid lentiviral expression system 134

    (Invitrogen). HEK293T cells were co-transfected with a mixture of FU(PSD-95:EGFP)W and 135

    the three packaging plasmids: pLP1, pLP2, and pLP\VSVG. Transfection was performed in T75 136

    flasks when the cells had reached 80% confluence, using 3 μg of the vector, 9 μg of the 137

    packaging mixture, and 36 μl of Lipofectamine 2000 (Invitrogen). Supernatant was collected 138

    after 48 and 72h, filtered through 0.45-μm filters, aliquoted, and stored at −80°C. 139

    Transduction of cortical cultures was performed on day 4-5 in vitro by adding 20μl of the 140

    filtered supernatant to each MEA dish. 141

    Pharmacological manipulations 142

    To chronically silence network activity, a mixture of three pharmacological agents was 143

    used: TTX (Alomone Labs), APV (Sigma-Aldrich) and CNQX (Sigma-Aldrich). The agents were 144

    first applied on day one in vitro, and additionally applied to feeding media every three 145

    consecutive days to maintain the same concentration in media. Final concentrations in the 146

    MEA dish were 1 μM (TTX), 10 μM (CNQX) and 50 μM (APV). These agents were also added 147

    to the perfusion media during long term imaging sessions. 148

    Electrophysiological recordings 149

    Network activity was recorded continuously from MEA electrodes (59 electrodes, 30μm 150

    diameter, arranged in an 8x8 array, spaced 200μm apart). A submerged platinum wire loop 151

    connected to a custom designed cap covering the MEA dish was used as a common 152

    reference (ground). Recordings from MEA dishes were performed using a commercial 60- 153

    channel headstage (inverted MEA-1060-BC, MCS) with a gain of 53x and frequency limits of 154

    0.02 to 8,500 Hz. This signal was further filtered with frequency limits of 150 to 3,000 Hz and 155

    amplified (20x) using a filter/amplifier (FA60S-BC, MCS). The 60 channels of amplified and 156

    filtered data were connected to 60 of the 64 analog to digital input channels of a data 157

    acquisition board (PD2-MF-64-3M/12; United Electronic Industries, Walpole, MA, USA) using 158

    a home built connection box. Data acquisition was performed using custom software (Closed 159

  • 8

    Loop Experiment Manager – CLEM; Hazan and Ziv, 2017). Data were collected at 16 160

    kSamples/sec. Action potentials were identified as negative threshold-crossing events, with 161

    the threshold calculated as 5x root-mean-square of traces recorded at the beginning of each 162

    experiment. Data were imported, converted and analyzed using custom scripts in Matlab 163

    (MathWorks, USA). 164

    Long-term imaging 165

    All fluorescence and bright-field images were obtained from neurons growing on thin 166

    glass MEA dishes, as described above. These particular dishes are fabricated of very thin 167

    glass (180 μm), which allows for the use of high numerical aperture, oil immersion objectives 168

    and are thus ideally suited for high-resolution imaging. Images were acquired using a 169

    custom-built confocal laser scanning (inverted) microscope based on a Zeiss Axio Observer 170

    Z1 using a 40×, 1.3 N.A. Plan-Fluar objective. The system was controlled by custom software 171

    and includes provisions for automated, multisite time-lapse microscopy. MEA dishes were 172

    mounted on the headstage/amplifier which was attached to the microscope’s motorized 173

    stage. The dish was covered with a custom-designed cap containing inlet and outlet ports for 174

    perfusion and air as well as a reference ground electrode as mentioned above. Continuous 175

    perfusion with fresh feeding medium (described above) was carried out at a rate of 4ml per 176

    day using an ultra-low-flow peristaltic pump (Instech Laboratories), and a pair of silicon 177

    tubes. The tubes were connected to the dish through the appropriate ports in the custom- 178

    designed cap. A mixture of 95% air and 5% CO2 was continuously streamed into the dish at 179

    very low rates through a third port, with flow rates regulated by a high-precision flow meter 180

    (Gilmont Instruments). The base of the headstage/amplifier and the objective were heated 181

    to 37°C and 36°C, respectively, using resistive elements, separate temperature sensors, and 182

    controllers, resulting in temperatures of approximately 37°C in the culture medium. Images 183

    of PSD-95:EGFP were obtained by excitation at 488nm using a solid state continuous wave 184

    laser (Coherent) and emissions were read simultaneously through a 500–550-nm bandpass 185

    filter (Semrock, USA) and >570nm (Chroma) after splitting the emission between two 186

    detectors using a 555nm longpass filter (Chroma). Time-lapse recordings were usually 187

    performed by averaging five frames at 10 focal planes spaced 0.8 μm apart. All data were 188

    collected at a resolution of 640×480 pixels, at 12 bits/pixel. Data were collected sequentially 189

    from multiple sites using a motorized stage to cycle automatically through these sites at 60- 190

  • 9

    min intervals. Focal drift was corrected automatically by using the confocal microscope 191

    autofocus system. 192

    Fluorescence recovery after photobleaching 193

    Photobleaching was performed by defining 16x16 pixel (~3.2 x 3.2μm) regions of interest 194

    and scanning them repeatedly at 488nm at high illumination intensity using the imaging 195

    systems’ acousto-optical tunable filter (AOTF) to limit illumination to the defined regions. 196

    Photobleaching was controlled through the confocal microscope’s ActiveX interface from 197

    scripts written in Visual Basic for Applications executed in Microsoft Excel. Fluorescence of 198

    each photobleached synapse was normalized (Ft, norm) according to 199

    , = −− Where Ft is the fluorescence at time t, Fmin is fluorescence at the end of the photobleaching 200

    procedure and F0 is the fluorescence just before the photobleaching procedure. 201

    202

    Image analysis 203

    All image analysis was performed using custom written software ("OpenView") which 204

    allows for automated or manual tracking of individual fluorescent puncta and measuring 205

    their fluorescence intensities over time (see Kaufman et al., 2012 for further details). 9 × 9 206

    pixel (~1.8 x 1.8μm) areas were centered on fluorescent puncta and mean pixel intensities 207

    within these areas were obtained from maximal intensity projections of Z section stacks. For 208

    measuring distributions of puncta intensities, areas were placed programmatically on 209

    fluorescent puncta at each time step using identical parameters. For tracking identified 210

    puncta, areas were placed initially over all puncta and then a smaller subset (typically 200 211

    per site) was tracked thereafter. As the reliability of automatic tracking was not absolute, all 212

    tracking was verified and, whenever necessary, corrected manually. Puncta for which 213

    tracking was ambiguous were excluded. 214

    To correct for some neuron to neuron variability in PSD-95:EGFP expression levels, raw 215

    puncta fluorescence measurements were normalized to mean PSD-95:EGFP puncta 216

    fluorescence of each neuron at the first time point (determined by placing areas 217

    programmatically on fluorescent puncta in the field of view), allowing us to pool data from 218

    different neurons and experiments. 219

  • 10

    Images for figures were processed by uniform contrast enhancement and low pass 220

    filtering using Adobe Photoshop and prepared for presentation using Microsoft PowerPoint. 221

    222

    Western blots 223

    Cortical cell preparations were grown in 12-well plates whose surface had been 224

    pretreated with polyethylenimine (Sigma) to facilitate cell adherence. Cells were washed 225

    using Tyrode’s solution (119mM NaCl, 2.5mM KCl, 2mM CaCl2, 25mM HEPES, 30mM glucose, 226

    buffered to pH 7.4) and lysed in RIPA buffer, 8M urea, 100 mM Tris–HCl. Protein 227

    concentrations were measured by the Bradford assay, using BSA as the standard. Equal 228

    protein amounts (25μm) were separated by SDS gel electrophoresis and transferred to 229

    nitrocellulose membranes. Membranes were blocked by non-fat milk, and then staining was 230

    performed using anti PSD-95 (Clone 108E10; Synaptic Systems; 1:1000) and anti-Actin 231

    (Merck; 1:10,000) as primary antibodies. As a secondary antibody, peroxidase-conjugated, 232

    anti-mouse (ImmunoResearch Laboratories, 1:10,000) was used. Prior to exposure, ECL 233

    (Enhanced Chemiluminescence; Pierce) was used for immunodetection. 234

    Multiplexed SILAC and Mass Spectrometry 235

    For multiplexed SILAC experiments, cells were prepared, raised in, and fed with lysine and 236

    arginine-free MEM (Biological Industries) to which 'heavy' (H) variants (Lys8, [13C6, 15N2]; 237

    Arg10, [13C6, 15N4]) or ‘medium’ (M) variants (Lys6 ,[13C6]; Arg6,[13C6]), were added to match 238

    nominal lysine and arginine concentrations in standard cell culture media (0.4mM and 239

    0.6mM respectively). Cells were harvested after 21-22 days in culture by scraping in lysis 240

    buffer containing 10% SDS, mixed together (as pairs of silenced and control sets) and run on 241

    preparative gels as follows: 20% of protein mixtures with additional concentrated Laemmli 242

    buffer were sonicated, boiled and separated on 4-15% SDS-PAGE (Polyacrylamide Gel 243

    Electrophoresis). Each lane was sliced into 5 sections (one being the stacking gel section) 244

    which were analyzed separately. Proteins in each slice were reduced with 3 mM DTT (60°C 245

    for 30 min), modified with 10 mM iodoacetamide in 100 mM ammonium bicarbonate (in the 246

    dark, room temperature for 30 min) and digested in 10% acetonitrile and 10 mM ammonium 247

    bicarbonate with modified trypsin (Promega) at a 1:10 enzyme-to-substrate ratio, overnight 248

    at 37°C. An additional second trypsinization was done for 4 hours. The resulting tryptic 249

    peptides were desalted using C18 tips (Harvard) dried and re-suspended in 0.1% Formic acid. 250

  • 11

    Peptides were analyzed by LC-MS/MS using a Q Exactive HF mass spectrometer (Thermo) 251

    fitted with a capillary HPLC (Easy nLC 1000, Thermo). The peptides were loaded onto a 252

    homemade capillary column (25 cm, 75 micron ID) packed with Reprosil C18-Aqua (Dr 253

    Maisch GmbH, Germany) in solvent A (0.1% formic acid in water). Peptide mixtures were 254

    resolved with a 5 to 28% linear gradient of solvent B (95% acetonitrile with 0.1% formic 255

    acid) for 105 minutes followed by gradient of 15 minutes gradient of 28 to 95% and 15 256

    minutes at 95% acetonitrile with 0.1% formic acid in water at flow rates of 0.15 μl/min. MS 257

    was performed in positive mode (m/z 300–1800, resolution 120,000) using repetitive full MS 258

    scans followed by collision induced dissociation (HCD, at 27 normalized collision energy) of 259

    the 20 most dominant ions (>1 charges) selected from the first MS scan. A dynamic exclusion 260

    list was enabled with exclusion duration of 20s. 261

    MS data was analyzed using MaxQuant 1.5.2.8. (www.maxquant.org) searching against 262

    the rat Uniprot database with mass tolerance of 20 ppm for the precursor masses and 20 263

    ppm for the fragment ions and 4.5ppm after calibration. Oxidation on methionine, 264

    phosphorylation on STY, gly-gly on K and protein N-terminus acetylation were accepted as 265

    variable modifications and carbamidomethyl on cysteine was accepted as static 266

    modifications. Minimal peptide length was set to six amino acids and a maximum of two 267

    miscleavages was allowed. Peptide- and protein-level false discovery rates (FDRs) were 268

    filtered to 1% using the target-decoy strategy. Protein tables were filtered to eliminate 269

    identifications from the reverse database, common contaminants and single peptide 270

    identifications. SILAC analysis was performed using the same software. H/M ratios for all 271

    peptides belonging to a particular protein species were pooled by the software, providing an 272

    average ratio for each protein. Data used in subsequent analyses were filtered according to 273

    the following criteria: 1) H/M ratios were quantified for at least 2 peptides in 3 out of 4 274

    experiments, and 2) no less than 8 peptides were quantified in total. For the set of 226 275

    synaptic proteins, total peptide numbers per protein were ~38±25 and ~32 (average ± 276

    standard deviation and median, respectively). All ratios were normalized to median H/M 277

    ratios in each sample (2,630 stringent proteins only). A median H/M ratio of 1.06 was 278

    obtained for 85 ribosomal proteins after this normalization, indicating that normalization 279

    was acceptable. 280

    Simulation of synaptic dynamics as Kesten processes 281

  • 12

    Simulation of synapse size dynamics as stochastic Kesten processes was done as 282

    described in Statman et al., 2014. At the beginning of each simulation, simulated synapses 283

    were set to initial values (see below). Their sizes were then evolved as follows: At each step 284

    and for each synapse, random values for ϵ and η were obtained from Gaussian distributions 285

    with means of and and standard deviations as indicated in Fig. 5 A,D,K,L. The 286

    random ϵ and η values were then used to calculate the new synapse size xt+1 from the prior 287

    size xt such that 288 = + . Synapses whose ‘sizes’ fell below zero were eliminated (set to zero) and 289 not evolved further. Distributions of synaptic sizes were calculated only for synapses with 290

    non-zero sizes. 291

    for each condition (silenced, control) was obtained from experimental data using 292

    multilinear regression fits to scatter plots such as those shown in Figs. 4D-I (see Statman et 293

    al., 2014 for a detailed explanation of this fitting process). For stationary size distributions 294

    and for normalized fluorescence data, the value of is (1 - ) (Statman et al., 2014) and 295

    thus was set to (1 - ). For the plots in Fig. 5A-J, initial synapse sizes were taken from 296

    the t=24h time point of the ~2,000 synapses tracked in each condition, and these were 297

    evolved for 320 time steps. For the plots in Fig. 5K-M, 4,000 synapses were initialized to an 298

    identical value of 0.1 and thereafter evolved for 480 time steps. 299

    Simulations were carried out using Visual Basic for Applications within Microsoft Excel. 300

    Simulation of synaptic dynamics as a Langevin process 301

    During the review of this manuscript, it was pointed out by one of the reviewers that the 302

    aforementioned Kesten process can be also formulated as a non-linear Langevin process, if 303

    the noise terms of ϵ and η are assumed to be normally distributed random variables. 304

    Specifically, given that the Kesten process is expressed as 305 = + then 306 Δx = ( − 1) + (note that in this discrete mapping, the actual values of ϵ and η depend on the time interval 307

    Δt. For simplicity, we assume Δt = 1 and that ϵ and η values are for this specific time 308

    interval). 309

    Assuming that ( − 1) and are normally distributed random variables then 310

  • 13

    Δx = ( , ) + ( , ) 311 where N1 and N2 are the random variables with means and variances of a,b (for N1) and c,d 312

    (for N2) respectively such that = 〈 − 1〉 (the mean of − 1), = (the variance of 313 − 1), = 〈 〉 (the mean of ) and = (the variance of ). 314 As ϵ and η are assumed to be independent (but see below) 315 Δx = [( + ) , ( + )] 316 Thus, the Kesten process can be expressed as a non-linear Langevin process: 317 Δx = ( + ) + + (0,1) 318 5) After substituting a,b,c,d with the equivalent Kesten process terms we arrive at 319 Δx = (〈 − 1〉 + 〈 〉) + + (0,1) which is a form of a non-linear Langevin process. 320

    Although fluctuations in momentary values of ϵ and η are assumed to occur 321

    independently (as this is the simplest assumption), the validity of this assumption is 322

    unknown. The formulation of the process as a non-linear Langevin process sidesteps this 323

    matter by employing a single noise term which is assumed to be a normally distributed 324

    random variable. Note, however, that in the most general case, the Kesten process makes no 325

    assumptions on the independence of ϵ and η or the shape of their distributions. 326

    Using this formulation, values for 〈 − 1〉, 〈 〉, , for Δt = 8h were obtained from 327 linear regression fits to binned synaptic size changes as shown in Fig. 6A,B (Control: -0.0913, 328

    0.1024, 0.2294, 0.0828; Silenced -0.0481, 0.0675, 0.1220, 0.1138, respectively). Then, 329

    starting with the experimentally observed distributions in control and silenced networks, we 330

    evolved the size of each synapse iteratively for 40, 8-hour steps using the Langevin process 331

    described, specifically 332 = + (〈 − 1〉 + 〈 〉) + + (0,1) 333 Here too, synapses whose sizes fell below zero were eliminated and not evolved further. 334

    Distributions of synaptic sizes were calculated only for synapses with non-zero sizes. 335

    Simulations were carried out using Visual Basic for Applications within Microsoft Excel. 336

    337

    Mesoscopic model of size dynamics 338

  • 14

    The mesoscopic model used to explore relationships between binding and unbinding 339

    kinetics of synaptic molecules, size fluctuations and distributions was based on the model 340

    described in Shomar et al., 2017. Here, each synapse was modeled as a 50x50 square matrix 341

    of sites/slots to which scaffold molecules can bind. Scaffold molecules could bind 342

    nonspecifically directly to the matrix with a low but non-zero probability α and to scaffold 343

    molecules in adjacent slots (see Fig. 7A) such that the probability of binding to a particular 344

    slot increased linearly with the number of occupied neighboring slots. At each time step and 345

    for each unoccupied slot, the fraction of occupied neighboring slots χ, was determined. 346

    Then, the probability Pon for a free scaffold molecule to bind to that slot was determined 347

    according to a) λon, the maximal binding probability (a constant); b) χ, the fraction of 348

    neighboring occupied slots, and c) Nfree, the amount of free (unbound) scaffold molecules, as 349

    well as α, such that 350

    Pon = Nfree · λon · χ + α. 351

    A random number was then sampled from a uniform distribution between 0 and 1. Binding 352

    ‘occurred’ if this number was smaller than Pon, in which case, Nfree was decremented. 353

    Similarly, for each step and each occupied site, the chances of unbinding were calculated 354

    according to χ, the fraction of occupied neighboring sites and λoff, the maximal unbinding 355

    probability (a constant) such that 356

    Poff = λoff · (1-χ). 357

    Here too, a random number between 0 and 1 was sampled from a uniform distribution and 358

    unbinding occurred if this number was smaller than Poff, in which case, Nfree was 359

    incremented. 360

    At the beginning of each simulation, all scaffold molecules were placed in the free pool, 361

    and matrices were set to be empty. Synaptic size at any time step was defined as the 362

    momentary number of molecules bound to its matrix. The procedure described above was 363

    run for 800 steps for 4,000 synapses (matrices), all of which shared (and competed over) a 364

    common pool of scaffold molecules. All presented data was taken from the last 72 365

    simulation steps (steps 727 to 799). 366

    FRAP was simulated by marking the bound molecules of 200 synapses as ‘bleached’ at 367

    simulation step 600, and then following their exchange with ‘unbleached’ molecules from 368

    the pool of free molecules over the subsequent 200 steps. As typical FRAP data in 369

    experiments were obtained from medium to large–sized synapses, FRAP curves in 370

  • 15

    simulations were prepared only from synapses whose average size in the 24 time steps 371

    preceding the simulated bleach procedure was equal to or exceeded average synaptic size 372

    during this period. 373

    Unless stated otherwise, the following parameters were used: λon=1.25·10-6; λoff=0.5; α= 374

    1·10-9; Total scaffold molecules=640,000. 375

    To attain good performance, simulations were written in C, using the fast cryptographic 376

    random number generator ISAAC (Indirection, Shift, Accumulate, Add, and Count; 377

    http://burtleburtle.net/bob/rand/isaacafa.html) to generate streams of pseudorandom 378

    numbers. Size trajectories and FRAP data were saved as text files and thereafter imported 379

    into Excel for further analysis and presentation. 380

    Code accessibility 381

    Code for simulations of synaptic size fluctuations, distributions, and loss modeled as 382

    Kesten and non-linear Langevin processes (Figs. 5, 6; Visual Basic for Applications within 383

    Microsoft Excel) can be found on Model DB (http://modeldb.yale.edu/262059). 384

    Code for mesoscopic simulations of synaptic size fluctuations and distributions (Figs. 385

    7,9,10; C code) can be found on Model DB (http://modeldb.yale.edu/262060). 386

    387

    388

  • 16

    Results 389

    Distributions of synaptic sizes in chronically silenced networks are broad and rightward 390

    skewed 391

    As described in the Introduction, much of synaptic size diversity is attributed to myriad 392

    synaptic plasticity processes, which depend, in turn, on network activity. It thus might be 393

    expected that in neurons with no history of network activity or synaptic transmission, size 394

    diversity would be less extensive, and this difference would be manifested in distributions of 395

    synaptic sizes. To examine this expectation, we raised networks of cultured rat cortical 396

    neurons from day one in culture in Tetrodotoxin (TTX, 1 μM), 6-cyano-7-nitroquinoxaline- 397

    2,3-dione (CNQX 10μM), and (2R)-amino-5 phosphono pentanoate (APV, 50μM), potent 398

    inhibitors of voltage gated sodium channels, AMPA-type and NMDA-type glutamate 399

    receptors, respectively. No overt effects on cell viability were observed, in agreement with 400

    many early studies (e.g. van Huizen et al., 1985, Ramakers et al., 1993; Verderio et al., 1994; 401

    Craig et al., 1994; Benson and Cohen, 1996; Murthy et al., 2001) as well as more recent ones 402

    (Wrosch et al., 2017; Hobbiss et al., 2018). 403

    To verify the elimination of all spiking activity, the networks were grown on thin-glass 404

    multi-electrode array (MEA) dishes, which allow for chronic, non-invasive recordings of 405

    network activity from 59 electrodes (Fig. 1A). As shown in Fig. 1B-D, these pharmacological 406

    agents fully suppressed the vigorous spontaneous activity typical of such networks. 407

    The effects of chronic silencing on excitatory synapse sizes were determined by 408

    expressing an EGFP-tagged variant of the PSD protein PSD-95 (PSD-95:EGFP) in a small 409

    number of neurons (Fig. 1E). PSD-95 is a major postsynaptic scaffold protein that regulates 410

    the number of AMPA and NMDA receptors at the postsynaptic membrane (Won et al., 411

    2017). PSD-95:EGFP fluorescence is correlated with PSD area (Cane et al., 2014) and thus 412

    represents a good proxy of synaptic size. Expression was carried out using lentiviral vectors, 413

    resulting in very low PSD-95:EGFP overexpression (Minerbi et al., 2009). 414

    Experiments were carried out on networks raised in culture for about three weeks. At this 415

    stage, the developmental phases of rapid dendritic growth, axonal arborization and synapse 416

    formation are over for the most part, and neuronal structure becomes relatively stable, 417

    allowing individual synapses to be followed reliably for 24-48 hours and beyond. Silenced (or 418

    control) networks growing on MEA dishes were mounted at day 19-21 in culture on a 419

    combined MEA recording / imaging system used in prior studies from our lab (Minerbi, et al., 420

  • 17

    2009; Kaufman et al., 2012, Rubinski et al., 2015; Dvorkin et al., 2016). The MEA dishes were 421

    maintained at 37°C in an atmospheric environment of 5% CO2 / 95% air and perfused at very 422

    slow rates (2 volumes / day) with fresh cell culture media containing (or free of) the 423

    aforementioned pharmacological agents. After 24 hour adjustment periods, automated 424

    multisite confocal microscopy was initiated, during which images of four to ten fields of view 425

    (portions of dendritic arbors of different neurons expressing PSD-95:EGFP) were obtained at 426

    60-min intervals at ten focal planes, using the microscopes ‘autofocus’ system to correct for 427

    focal drift. 428

    Following the experiments, PSD-95:EGFP puncta were identified anew at each time point 429

    (programmatically, using a puncta detection algorithm as described in Materials and 430

    Methods) and intensities and numbers of all puncta were determined (Silenced networks: 23 431

    neurons, from 5 separate experiments, ~6,800 synapses; Control networks: 20 neurons from 432

    6 separate experiments, ~9,000 synapses). As shown in Fig. 2A, distributions of synaptic PSD- 433

    95:EGFP fluorescence in chronically silenced networks were broad and rightward skewed 434

    (skewness ≈ +1.4 and +2.2, silenced and control networks, respectively; note that 435

    skewness=0 for normal distributions). In fact, distributions in the silenced networks were 436

    much broader than those observed in active networks, with mean PSD-95:EGFP fluorescence 437

    being ~1.5 times greater than mean fluorescence measured in active networks (Fig. 2B; 438

    p=4.6 ּ10-6 by neuron; p=0.019 by experiment; t-test, assuming unequal variances; see also 439

    Kim et al., 2007; Noritake et al., 2009; Sun and Turrigiano, 2011; Shin et al., 2012) suggesting 440

    that chronic silencing was associated with significant synaptic growth (Murthy et al., 2001; 441

    Sando et al., 2017; but see Harms et al., 2005; Yasumatsu et al., 2008). Distributions at early 442

    and late time points (1 and 24h, respectively) were very similar (Fig. 2A). Mean synaptic PSD- 443

    95:EGFP fluorescence was also stable (Fig. 2B). Plotting distributions of PSD-95:EGFP 444

    fluorescence in control networks in scaled units (that is, multiplying the fluorescence of each 445

    synapse by ~1.5), suggested that shapes of synaptic size distributions in chronically silenced 446

    and active networks were similar (Fig. 2C), in excellent agreement with prior findings in 447

    acutely silenced networks in vitro (e.g. Turrigiano et al., 1998; Hobbiss et al., 2018) and in 448

    vivo (Keck et al., 2013). Finally, distributions of synaptic sizes in both silenced and active 449

    networks were well approximated by log-normal distributions (Fig. 2D). 450

    These findings thus confirm prior reports that extensive synaptic size diversification can 451

    occur in the absence of activity-dependent synaptic plasticity processes (e.g. Van Huizen et 452

  • 18

    al., 1985; Harms and Craig, 2005; Harms et al., 2005; Yasumatsu et al., 2008; Sigler et al., 453

    2017; Sando et al., 2017). Moreover, they reveal that the emergence of broad, rightward 454

    skewed and stable size distributions, remarkably similar to those observed in active 455

    networks, can arise de novo, rather than through the scaling of distributions initially 456

    established in active networks. Thus, activity-independent processes can play decisive roles 457

    in synaptic size diversification and in establishing appropriate proportions of differently sized 458

    synapses. 459

    460

    Intrinsic size fluctuations are sufficient to produce differently sized synapses at appropriate 461

    proportions 462

    As mentioned in the introduction, prior studies suggest that synaptic sizes are affected by 463

    intrinsic size fluctuations, as are the shape and scale of synaptic size distributions 464

    (Yasumatsu et al., 2008; Lowenstein et al., 2011; Kaufman et al., 2012; Statman et al, 2014; 465

    Rubinski et al., 2015; Ziv and Brenner, 2018; Ishii et al., 2018; Humble et al., 2019). Yet, as 466

    characteristics of intrinsic size fluctuations in neurons with no prior history of network 467

    activity were not measured to date, it remained unknown if such intrinsic fluctuations are 468

    sufficient to produce the extensive diversity and broad, skewed distributions of synaptic 469

    sizes observed in chronically silenced networks. 470

    To examine this possibility, we followed individual synapses in chronically silenced (and 471

    active) networks for 24-48 hours, measuring PSD-95:EGFP fluorescence of each synapse at 472

    each time point as illustrated in Fig. 3A (see also Minerbi et al., 2009; Fisher-Lavie and Ziv 473

    2013; Rubinski et al., 2015; Dvorkin et al., 2016). Fluorescence measurements were made 474

    from maximal intensity projections of all Z-sections to minimize the effects of focal 475

    positioning errors. Only synapses that could be tracked reliably were included in these 476

    analyses, excluding PSD-95:EGFP puncta that disappeared, split, or merged during the 477

    experiments. To correct for some variability in PSD-95:EGFP expression levels and to allow 478

    for data pooling, the fluorescence of each synapse was normalized to the mean puncta 479

    fluorescence of its respective neuron, measured at the first time point. Additionally, to 480

    minimize the influence of measurement noise, all data were smoothed using a 3 time-point 481

    low pass filter (see Statman et al., 2014). As illustrated for 16 synapses in Fig. 3A, synaptic 482

    sizes changed considerably over time scales of many hours, even in chronically silenced 483

    networks (Fig. 3B). Comparing the initial synaptic ‘configuration’ (the set of inputs to this 484

  • 19

    dendrite in terms of synaptic sizes) to synaptic configurations at later time points suggests 485

    that size fluctuations are associated with a gradual ‘erosion’ of synaptic configurations (Fig. 486

    3C,D). 487

    Magnitudes of temporal fluctuations in synaptic sizes were quantified for 2,032 and 1,922 488

    synapses (23 and 20 neurons, 5 and 6 separate experiments, silenced and control networks, 489

    respectively) by calculating the standard deviation, coefficient of variation, and the 490

    range/mean (Fisher Lavie et al., 2011; Zeidan and Ziv, 2012; Ziv, 2013) of the normalized 491

    fluorescence of each synapse over 24 hour periods. As shown in Fig. 4A-C, all three measures 492

    suggested that magnitudes of size fluctuations were reduced in silenced networks relative to 493

    control (active) networks, but only by 20% to 34%. 494

    To determine if these somewhat subdued fluctuations could give rise to the broad and 495

    skewed size distributions observed in chronically silenced networks, we analyzed these 496

    within the context of a statistical framework we previously developed (Statman et al., 2014). 497

    The basic premise of this framework is that synaptic size dynamics are driven by continuous, 498

    noisy multiplicative downscaling, which is continuously offset by noisy additive growth, 499

    resulting in size fluctuations that have noisy multiplicative and additive components. This 500

    statistical process, known as a Kesten process, faithfully reproduces many experimental 501

    observations concerning synaptic size fluctuations, size distributions, their stability and their 502

    scaling (Statman et al., 2014; Rubinski et al., 2015; Ziv and Brenner, 2018). Importantly, this 503

    framework provides means for parametrically comparing synaptic size fluctuations under 504

    different experimental conditions and determining their effects on synaptic size 505

    distributions. 506

    In more formal terms, this framework stipulates that for a synapse of size x at time t (xt), 507

    its size (xt+1) after some discrete time period will be 508 = + (1) 509 where εt and ηt are the aforementioned multiplicative downscaling and additive growth 510

    parameters. Importantly, εt and ηt are not fixed values but random variables drawn 511

    independently at each time step from some distribution. Iterations of process (1) (i.e. 512 = + ; = + ; etc.) result in fluctuating size 513 ‘trajectories’ similar to those observed experimentally. Note that expressing the change in 514

    synaptic size(Δxt+1) at t=t+1 using equation 1 515 ∆ = − = + −

  • 20

    ∆ = ( − 1) + (2) 516 suggests that the Kesten process can be thought of as a combination of myriad, noisy first 517

    and zero order reactions (for example protein loss/degradation and protein 518

    supply/synthesis, respectively) with 〈 − 1〉 and 〈 〉 (the mean values of these parameters) 519 representing aggregate, effective ‘rate constants’ of such reactions, respectively (Statman et 520

    al., 2014). 521

    The mean downscaling factor () is an important parameter in this framework, and can 522

    be derived in stationary distributions by multiple linear regression analyses of synaptic sizes 523

    as a function of time as illustrated in Fig 4D-J (see Statman et al., 2014 for further details). 524

    Derivation of in this manner revealed that continuous downscaling was substantially 525

    weakened in silenced networks in comparison to active networks, that is was closer to 526

    1.0 (0.995 and 0.985, silenced and control networks, respectively; Fig. 4J). In addition, this 527

    analysis revealed that synaptic configuration ‘erosion’ rates were roughly halved (Fig. 4K); 528

    notably, however, erosion rates were still considerable. Plotting the change in synaptic size 529

    (that is, subtracting, for each synapse, its fluorescence at t=24 from its fluorescence at t=1) 530

    as a function of its size at t=1 illustrates how lessened downscaling weakens the constraints 531

    on synaptic size distributions (Fig. 4L). This weakening might explain the broader 532

    distributions of synaptic sizes in silenced networks (Fig. 2A). On the other hand, it remained 533

    unclear if fluctuations with such weak downscaling would be sufficient to generate and 534

    maintain the broad, skewed and stable synaptic size distributions observed in silenced 535

    networks. 536

    To address this question, we simulated populations of ‘synapses’ whose size fluctuations 537

    were modeled as Kesten processes using the experimentally derived values of and 538

    for silenced and control networks. In the first set of simulations, synaptic sizes were 539

    initialized using the experimentally measured, normalized PSD-95:EGFP fluorescence values 540

    of 2,032 and 1,922 synapses (data of Fig. 4; silenced and control networks respectively). As 541

    shown in Fig. 5, excellent fits to the experimental data were obtained for both silenced and 542

    control conditions not only in distribution shape and stability (Fig. 5C,F) but also in terms of 543

    synaptic configuration erosion rates (compare Fig. 5A,B,D,E with Fig. 4F,I,L) and measures of 544

    size fluctuations (Fig. 5G-I and 4A-C). 545

    In these simulations, synapses whose ‘sizes’ were reduced momentarily to zero were 546

    treated as ‘eliminated’ and not considered further. Interestingly, rates of synapse 547

  • 21

    ‘elimination’ were greater in simulations based on parameters obtained in silenced networks 548

    (Fig. 5J; ~0.63±0.05% per 24 simulation cycles, silenced networks; 0.06±0.02%, control 549

    networks; mean ± standard deviation, 5 runs per condition) indicating that the weaker 550

    constraints on intrinsic fluctuations observed in silenced networks might reduce the chances 551

    of (small) synapses to escape elimination (see also Holtmaat et al., 2006; Yasumatsu et al., 552

    2008; Minerbi et al., 2009). 553

    In a second set of simulations, we simulated 4,000 synapses for each condition, setting 554

    the initial ‘size’ of all synapses to 0.1 (the dimmest synapses identified in our experimental 555

    data sets, in normalized units), and followed the evolution of synaptic size distributions using 556

    the experimentally derived values of and . As shown in Fig. 5K,L, in both conditions, 557

    size distributions gradually converged to the experimentally measured skewed and stable 558

    distributions. Interestingly, however, convergence was much slower in silenced networks, 559

    and even after 480 simulated ‘hours’ (20 ‘days’) convergence was incomplete. Here too, the 560

    negative effects of silencing on (small) synapse survival were very evident (Fig. 5M). 561

    Collectively, these findings suggest that fluctuations in synaptic sizes measured in chronically 562

    silenced networks are sufficient to drive the emergence of the broad, rightward skewed and 563

    stable synaptic size distributions observed in these networks. 564

    The Kesten process described above is based on the assumption that , and their 565

    noise terms do not depend on momentary synaptic size (as might be expected for simple 566

    first and zero order reactions). In a prior study (Yasumatsu et al., 2008), size fluctuations 567

    were modeled using a generic framework that does not necessitate this assumption. Here 568

    fluctuations were modeled as a non-linear Langevin process in which fluctuations were 569

    grouped into deterministic ( ( )) and stochastic terms ( ( )), both formulated as functions 570 of momentary synaptic size (xt). 571

    Assuming that the two noise terms in the Kesten process (2) are distributed normally with 572

    standard deviations of and , it can be shown that the Kesten process (2) can also be 573

    formulated as a non-linear Langevin process (see Materials and Methods). Here, the change 574

    in momentary synaptic size Δ after some time interval (Δt) is expressed as 575 Δ = ( )Δ + ( ) (0,1)Δ (3) 576 with 577 ( ) = (〈 − 1〉 + 〈 〉) (4) 578

  • 22

    ( ) = −12 2 + 2 or ( ) = −12 2 + 2 (5) 579 where (0,1) is a random variable taken from a normal distribution with a mean of 0 and a 580 variance of 1. 581

    This formulation allowed us to test the aforementioned assumptions: if , are 582

    indeed independent of momentary synaptic size, then plotting the fluctuation magnitude 583 ( ) as a function of synaptic size at that time ( ) should result in a straight line with a 584 slope of 〈 − 1〉 and an intercept of 〈 〉, (both scaled by ∆t; see equation 4). Similarly, if the 585 variances of ϵ and η do not depend on momentary synaptic size, plotting the fluctuation 586

    variance ( ( ) ) as a function of should result in a straight line with a slope of and 587 an intercept of (equation 5). To test these predictions, size changes measured over 8 hour 588

    intervals were divided into 20 equally-sized bins according to synaptic size at the beginning 589

    of each interval. The average and variance of size changes in each bin were then plotted 590

    against average initial synaptic size for that bin. As shown in Fig. 6A,B, excellent linear fits 591

    were observed for both the control and silenced conditions, justifying the aforementioned 592

    assumptions. The only noticeable deviation was for ( ) and for the smallest synapses in 593 the silenced data set, which might hint that in silenced networks, ϵ and/or η might slightly 594

    differ for the smallest synapses. 595

    Linear regression fits of the data in Fig. 6A,B allowed us to obtain estimates of 596 〈 〉, 〈 〉, , (for 8 hour intervals) for control and silenced networks. These estimates were 597 then used to examine if synaptic size fluctuations modeled as the Langevin process 598

    described above give rise to the size distributions measured experimentally. To that end, 599

    experimentally measured synaptic sizes (as in Fig. 5C,F) were evolved for 320 simulated 600

    hours (40, 8-hour intervals) using equations 3 to 5 (see Materials and Methods for further 601

    details). As shown in Fig. 6C,D, distributions remained faithful to the experimentally 602

    measured distributions for both control and silenced networks. As expected, identical results 603

    were obtained when synaptic sizes were evolved as a Kesten process using the same 604

    parameters (data not shown). 605

    Collectively these findings suggest that synaptic size distributions, in both active and 606

    chronically silenced networks, can arise from size fluctuations that effectively behave as 607

    combinations of noisy first and zero order processes. Interestingly, the Langevin 608

    transformation of the Kesten process (equations 3 to 5) is very similar to the non-linear 609

  • 23

    Langevin process previously formulated by Yasumatsu and colleagues as an effective 610

    description of size fluctuations in cultured slices of hippocampal neurons (Yasumatsu et al., 611

    2008). Thus analytical approaches coming from different directions and theoretical 612

    backgrounds, applied to data obtained in different experimental systems, converged to a 613

    very similar quantitative description of synaptic size dynamics. 614

    615

    Although synaptic dynamics in both active and silenced networks were well-described by 616

    stochastic Kesten and Langevin processes (Figs. 4-6), we noted one qualitative deviation in 617

    active but not in silenced (or simulated) networks, that is, a minor but conspicuous 618

    population of small synapses that exhibited rapid growth over the imaging period (Fig. 4F, 619

    shaded area; compare with Fig. 4I). We return to this population later. 620

    621

    Relationships between activity levels, innate molecular dynamics, intrinsic fluctuations, and 622

    size distributions 623

    The data described so far suggests that activity-independent, intrinsic size fluctuations are 624

    sufficient to generate a full range of synaptic sizes at correct proportions as reflected in the 625

    breadth and shape of the resulting size distributions. The source of these fluctuations, 626

    however, is not clear. As mentioned in the Introduction, these have been suggested to stem 627

    from the innate dynamics of synaptic molecules at synaptic sites. Moreover, network activity 628

    levels have been shown to affect these dynamics (reviewed in Fisher-Lavie and Ziv, 2014). It 629

    is thus plausible that the altered size fluctuations (and consequential changes in synaptic size 630

    distributions) observed in silenced networks reflect, at least in part, changes in the 631

    underlying dynamics of synaptic molecules associated with low network activity levels. 632

    To obtain a better understanding of the relationships between activity levels, innate 633

    molecular dynamics, intrinsic fluctuations, and size distributions, we used a mesoscopic 634

    model developed previously to study such relationships (Shomar et al., 2017). Specifically, 635

    we used this model to generate hypotheses on the manners by which activity might affect 636

    relationships between innate molecular dynamics, intrinsic fluctuations and size 637

    distributions, and then tested these hypotheses experimentally. It should be emphasized 638

    that the model was used to explore potential explanations, not to generate precise fits to 639

    experimental data. 640

  • 24

    The aforementioned model (illustrated in Fig. 7A) consists of a neuron with a fixed 641

    number of ‘synapses’ (S) each of which is composed of two components: a postsynaptic 642

    ‘membrane’, modeled as a 50x50 matrix of potential binding sites (‘slots’) for synaptic 643

    scaffold molecules; and synaptic scaffold molecules which can bind to, or unbind from these 644

    slots. The ‘size’ of a given synapse at any time is defined as the momentary number of 645

    occupied slots, that is, the number of scaffold molecules bound to its matrix. In the variant 646

    of the model used here, scaffold molecules come from a common (global) pool (Ntotal) shared 647

    and competed over by all synapses. The global amount of free molecules (Nfree) at any 648

    moment is equal to the total amount of molecules in the cell (Ntotal) after subtracting all 649

    molecules presently bound to synaptic membranes (matrices). Binding and unbinding are 650

    modeled as stochastic events characterized by probabilities per unit time. Consequently, the 651

    number of molecules binding to a matrix per unit time depends on the binding probability, 652

    the number of free molecules (Nfree, serving as a proxy of free molecule concentration) and 653

    on the number of vacant slots in that matrix. Similarly, the number of molecules dissociating 654

    from each matrix per unit time depends on the unbinding probability per unit time and on 655

    the number of bound molecules (=occupied slots). In this stochastic description, the binding 656

    and unbinding of scaffold molecules result in temporal fluctuations in synaptic sizes, i.e. in 657

    the momentary numbers of molecules bound to the matrices, while occupancies at all S 658

    matrices give rise to momentary synaptic size distributions. The model as described so far is 659

    insufficient to explain the rightward skewed, experimentally observed distributions of 660

    synaptic sizes. However, when the probabilities of binding to (and unbinding from) each slot 661

    depend positively (and negatively) on the number of its immediately neighboring occupied 662

    slots (gray area in Fig. 7A, right hand side), synaptic size dynamics and distributions become 663

    remarkably similar to those observed experimentally (Shomar et al., 2017). This dependence, 664

    justified by the multiplicity of binding sites typical of most synaptic molecules (e.g. Won et 665

    al., 2017), is essentially a form of cooperativity, and thus both binding and unbinding in this 666

    model are cooperative (see Materials and Methods for a more detailed description of the 667

    model). 668

    Prior studies suggest that chronic suppression of network activity can slow the binding 669

    and unbinding (exchange) kinetics of synaptic molecules (reviewed in Fisher-Lavie and Ziv, 670

    2014). We thus used this model to examine the possibility that reduced synaptic size 671

    fluctuations and broader size distributions in silenced networks might stem from slower 672

  • 25

    exchange kinetics. Specifically, we explored the expected consequences of reducing the 673

    unbinding probability of molecules bound to ‘synaptic’ matrices. As shown in Fig. 7, lower 674

    unbinding probabilities would be expected to drive the broadening of synaptic size 675

    distributions (Fig. 7B) increase mean synaptic size (Fig. 7C), reduce the rate at which the 676

    slope declines in plots such as Fig. 7D (that is, drive to values closer to 1.0; Fig. 7D-E,G), 677

    slow changes in synaptic size configurations (Fig. 7F), and reduce size fluctuation magnitudes 678

    (Fig. 7H-J), all in good agreement with observations made in real neurons (Figs. 2, 4). 679

    Lower unbinding probabilities also predict slower exchange rates of scaffold molecules at 680

    synapses (Fig. 7K). If this hypothesis is correct, experimentally measured exchange rates of 681

    PSD-95 might be expected to be slower in chronically silenced networks {due to, for example 682

    Ser-295 phosphorylation (Kim et al., 2007) or palmitoylation (Sturgill et al., 2009; Noritake et 683

    al., 2009; Fukata et al., 2013)}. To test this prediction, we measured PSD-95:EGFP exchange 684

    rates using fluorescence recovery after photobleaching (FRAP) in chronically silenced and 685

    control networks. To that end, a small number of well separated PSD-95:EGFP puncta were 686

    photobleached by intense laser illumination and subsequently followed by time lapse 687

    imaging, initially at 10 min intervals (for the first hour) and then at one hour intervals, 688

    chosen to match to the slow exchange rates of PSD-95 (Sturgill et al., 2009; Zeidan and Ziv, 689

    2012; Fukata et al., 2013). One example is shown in Fig. 8A-C. Here, three photobleached 690

    PSD-95:EGFP puncta in a chronically silenced network were followed for 90 hours after the 691

    bleaching procedure. The fluorescence traces obtained here (Fig. 8C) illustrate that over 692

    these long time scales, fluorescence recovery measurements are confounded by ongoing 693

    changes in synaptic sizes, complicating accurate estimations of recovery kinetics. 694

    Nevertheless, pooling measurements over shorter time scales (24h or less) allowed us to 695

    compare mean fluorescence recovery profiles for synapses in chronically silenced (77 696

    synapses from 6 experiments) and control networks (72 synapses from 7 experiments). 697

    Surprisingly, mean fluorescence recovery curves for silenced and control networks did not 698

    differ significantly (Fig. 8D). These experiments thus did not support the possibility that 699

    altered size fluctuations, synaptic sizes and distributions in chronically silenced networks 700

    reflect slower unbinding kinetics of PSD-95. 701

    We thus explored an alternative explanation that relates to PSD-95 abundance. As shown 702

    in (Fig. 2A,B) mean synaptic PSD-95:EGFP fluorescence was, on average ~50% greater in 703

    chronically silenced networks, possibly indicating that chronic silencing is associated with 704

  • 26

    increased cellular PSD-95 levels due to, for example, effects on synaptic protein synthesis 705

    (e.g. Schanzenbächer et al., 2016) or degradation (e.g. Jakawich et al., 2010). Using the 706

    aforementioned model to examine the expected effects of increased scaffold molecule 707

    abundance we found that merely increasing Ntotal by 25% was sufficient to qualitatively 708

    recapitulate all of the experimental findings described so far (Fig. 9A-H), including the similar 709

    recovery kinetics in FRAP experiments. 710

    To experimentally test this potential explanation, we measured and compared global 711

    PSD-95 abundance in chronically silenced and control networks by means of Western blots. 712

    Here too, however, the prediction was not supported by the experimental data, as no 713

    consistent increases in global PSD-95 abundance were observed in silenced networks (Fig. 714

    9I,J; 7 replicates from 2 separate experiments; see also Kim et al., 2007; Shin et al., 2012; 715

    Lazarevic et al., 2011). 716

    To examine if this observation applies to synaptic proteins in general, we used 717

    multiplexed SILAC (Stable Isotope Labeling with Amino acids in Cell culture) combined with 718

    MS (Mass Spectrometry; reviewed in Hoedt et al., 2019) to compare synaptic protein 719

    quantities in silenced and control preparations. To that end, neurons were prepared and 720

    grown in lysine and arginine-free media supplemented with lysine and arginine containing 721

    stable, heavy isotopes of carbon and nitrogen. Silenced preparations were labeled with 722

    ‘Heavy’ amino acids (Lys8 - 13C6, 15N2 and Arg10 - 13C6, 15N4) whereas control preparations 723

    were labeled with ‘Medium’ variants (Lys6 - 13C6 and Arg6 - 13C6), which are isotopically 724

    separable from both Heavy and unlabeled lysine and arginine. After 21-22 days in culture, 725

    during which most of the proteome becomes labeled (Hakim et al., 2016), the preparations 726

    were lysed, the extracts mixed together and run on preparative gels which were 727

    subsequently sliced and subjected to MS analysis (See Materials and Methods and Hakim et 728

    al., 2016 for further details). This method, based on mixing and analyzing samples 729

    simultaneously, eliminates much of the variability associated with proteomic approaches; 730

    moreover, it provides a Heavy/Medium (H/M) ratio reading for each peptide and protein, 731

    which reflects the ratio of labeled proteins from silenced and control preparations, 732

    respectively. Average H/M ratios were obtained from 4 replicates (2 separate experiments) 733

    using stringent criteria (see Materials and Methods). Stringent H/M ratios obtained for 226 734

    synaptic proteins (categorized as such as in Hakim et al., 2016; 2,630 proteins in total) 735

    resulted in average and median H/M ratios of 0.89 and 0.86, respectively. Data for 15 and 21 736

  • 27

    well-studied post- and presynaptic proteins are shown in Fig. 9K. Evidently, these data do 737

    not support the possibility that chronic silencing increases synaptic protein abundance (if 738

    anything, we noted a slight reduction), although confounds related to differences in labeling 739

    rates (due to differential metabolism) or cell counts cannot be entirely ruled out. 740

    How could synaptic contents of PSD-95 increase by ~50% without detectable changes in 741

    PSD-95 abundance? One possible explanation is that increased synaptic size in silenced 742

    networks was associated with a commensurate decrease in synaptic number (e.g. van 743

    Huizen et al., 1985; Kossel et al., 1997; see also Annis et al., 1994; Barnes et al., 2017) 744

    resulting is no net change in total PSD-95 levels. Indeed, such decreases were predicted by 745

    the simulations shown in Figs. 5J,M. 746

    We thus used the aforementioned mesoscopic model to examine how a reduction in 747

    synaptic numbers might affect intrinsic fluctuations and synaptic size distributions. To that 748

    end, the number of synapses (matrices) was reduced by 40% (from 4,000 to 2,400) without 749

    changing Ntotal. As shown in Fig. 10A-G the model with these parameters recapitulated the 750

    main experimental findings obtained in silenced networks – increased synaptic size, 751

    broadening of synaptic size distributions, values closer to 1.0, reduced magnitudes of 752

    synaptic size fluctuations, slower changes in synaptic configurations and similar FRAP curves. 753

    To examine if the experimental findings were congruent with this prediction, we revisited 754

    the data set of Fig. 2, finding (Fig. 10H) that PSD-95:EGFP puncta counts were indeed 755

    reduced in chronically silenced networks (by ~37%; ~296 vs. ~470 puncta per field of view, 756

    23 fields of view in 5 experiments and 20 fields of view in 6 experiments, silenced and 757

    control networks, respectively). Moreover, the summed (rather than average) fluorescence 758

    of PSD-95:EGFP puncta in each field of view was practically identical in silenced and active 759

    networks (Fig. 10I) - in agreement with the explanation proposed above as well as the data 760

    of Fig. 9I-K. Differences in synaptic numbers were not associated with changes in synaptic 761

    density (4.91±1.18 and 5.02±0.87 synapses per 10μm dendrite length; 11 and 10 fields of 762

    view from 6 and 5 experiments; silenced and control networks respectively) and thus seem 763

    to reflect lessened dendritic arborization in silenced networks (in agreement with van Huizen 764

    et al., 1985; Benson and Cohen, 1996). Indeed, the summed length of dendritic segments 765

    within each field of view was reduced by ~33% (Fig. 10J; 20 fields of view for each condition, 766

    P= 1.3·10-7; two-sample t-test assuming unequal variances). Moreover, Scholl analysis of 767

    reconstructed neurons confirmed that dendritic arborization was substantially reduced (Fig. 768

  • 28

    11; see also Benson and Cohen, 1996). Interestingly, whereas PSD-95:EGFP puncta counts in 769

    silenced networks were ~stable over 24 hour periods (Fig. 10H), counts in active networks 770

    tended to increase slowly over the same time frame (see also Okabe et al., 1999), as did the 771

    summed fluorescence of PSD-95:EGFP puncta in active networks (Fig. 10I). 772

    These findings are thus most congruent with an interpretation suggesting that synaptic 773

    enlargement in chronically silenced networks, as well as broader size distributions, subdued 774

    size fluctuations and values closer to 1.0, might be attributed to the redistribution of 775

    PSD-95 (and probably other synaptic molecules) among the fewer synaptic connections 776

    formed in the absence of network activity (Fig 12A). 777

    778

    Relationships between activity levels and synaptic numbers 779

    Why are less synapses formed in chronically silenced networks? Our data provides 780

    potentially interesting clues, although, as we show below, interpretation is not as 781

    straightforward as it might seem at first sight. 782

    We note that in active (Fig. 4D), but not in silenced networks (Fig. 4I), a minor population 783

    of synapses (~1-2% per day) exhibited rapid growth in manners not predicted by any of the 784

    models explored here. In a prior study (Minerbi et al., 2009) we found that this phenomenon 785

    reflects the rapid formation and enlargement of (post)synaptic sites during periods of 786

    particularly strong, synchronous network activity, which presumably drives strong 787

    presynaptic activation and possibly long-term potentiation associated spine formation and 788

    growth (e.g. Smith and Jahr, 1992; Engert and Bonhoeffer, 1999; Maletic-Savatic et al., 1999; 789

    Matsuzaki et al., 2004; Kwon and Sabatini, 2011; Meyer et al., 2014; Bosch et al., 2014; Sigler 790

    et al., 2017; Hobbiss et al., 2018 reviewed in Andreae and Burrone, 2014). Conversely, acute 791

    suppression of network activity was found to abruptly arrest and even reverse trends of 792

    synaptic proliferation (Minerbi et al., 2009). It thus seems that network silencing might 793

    suppress activity-dependent forms of synapse formation (in line with prior predictions, e.g. 794

    Yasumatsu et al., 2008), ultimately resulting in lower synaptic numbers. This, in turn, would 795

    negatively affect dendritic arborization (reviewed in Cline and Haas, 2008) potentially 796

    explaining the major findings described above. 797

    Our data also indicates, however, that causal relationships between synaptic numbers 798

    and network activity might be more complex. The increase in PSD-95 availability and 799

    synaptic size associated with reduced synaptic numbers (Fig. 12A) is also associated with 800

  • 29

    changes in size fluctuation characteristics, specifically in and (Fig. 4J). When 801

    expected size change is plotted against present synaptic size using experimentally measured, 802

    absolute values of and (Fig. 12B) it becomes apparent that for particularly small 803

    synapses (Fig. 12B gray shading), the bias toward growth is weaker in silenced networks (see 804

    also Fig. 6A). Consequently, the chances of small synapses to persist in silenced networks is 805

    lowered, and thus more synapse elimination (or less stabilization of nascent synapses) is 806

    expected, as predicted by the simulations of Figs. 5J,M. This bias is not affected by 807

    recalculating and for active networks after removing the population of rapidly 808

    growing synapses from the data set, although this does not preclude the possibility that 809

    some of this bias is created by activity-dependent potentiation of nascent synapses. 810

    To test this prediction, we returned to the time-lapse images obtained in control and 811

    silenced networks, focusing this time on small (dim) PSD-95:EGFP puncta. As shown in Fig. 812

    12C, tracking such synapses revealed greater rates of puncta loss in silenced networks (5.7% 813

    vs. 12.6% in 24 hours; 17/299 and 41/324, 4 control and 4 silenced networks, respectively). 814

    In further agreement with this prediction, lost puncta were particularly dim (333±147 and 815

    373±125; arbitrary fluorescence units at first time point, average ± standard deviation, 816

    control and silenced networks, respectively) as compared to the entire synaptic population 817

    (600±357 and 911±558). 818

    These predictions (Fig. 5J,M, Fig. 12B) and findings (Fig. 10H-J, Fig. 12C) suggest that 819

    synaptic loss might be both the cause and the product of altered synaptic size dynamics. 820

    Stated differently, innate molecular dynamics, intrinsic size fluctuations and synaptic sizes 821

    might be related reciprocally, possibly in self-reinforcing fashion as illustrated in Fig. 12D, 822

    potentially obscuring simple cause and effect relationships between these phenomena. 823

    824

    Discussion 825

    Here we set out to determine the contributions of activity dependent and independent 826

    processes to excitatory synapse size diversity. To that end, we pharmacologically silenced 827

    networks of cortical neurons from the time of plating and then examined synaptic size 828

    distributions and remodeling dynamics using PSD-95:EGFP fluorescence as a proxy of 829

    synaptic size. We found that even in networks with no history of activity, size diversity was 830

    extensive and size distributions were broad, stable, and rightward skewed. Comparisons 831

    with spontaneously active networks revealed that silencing was associated with a 832

  • 30

    broadening of synaptic size distributions and a significant (~50%) increase in mean synaptic 833

    size, yet distribution shapes were similar when scaled by mean synaptic size. Silencing was 834

    associated with reductions in size fluctuation magnitudes, as well as considerable weakening 835

    of constraints on size distributions. Nevertheless, these fluctuations and constraints were 836

    still sufficient to generate broad, skewed and stable size distributions. To better understand 837

    relationships between activity levels, size fluctuations, size distributions and innate dynamics 838

    of synaptic molecules, we used a previously published mesoscopic model to derive potential 839

    explanations which were then tested experimentally. Explanations attributing the effects of 840

    chronic silencing to changes in PSD-95 binding/unbinding kinetics or expression levels were 841

    not supported by FRAP experiments, Western blots or quantitative proteomics. Conversely, 842

    the experimental findings fully supported the possibility that changes in synaptic size 843

    dynamics and distributions primarily reflect PSD-95 redistribution among fewer synapses. 844

    These findings thus suggest that intrinsic, activity-independent size fluctuations are sufficient 845

    to give rise to full repertoires of synaptic sizes at appropriate proportions. Moreover, they 846

    are suggestive of reciprocal and possibly self-reinforcing relationships between synaptic size 847

    fluctuations, size distributions and synaptic counts mediated by the innate dynamics of 848

    synaptic molecules as they continuously move in, out and between synapses 849

    850

    The source of broad and rightward skewed synaptic size distributions 851

    As mentioned above, broad, rightward skewed distributions of synaptic sizes are 852

    ubiquitously observed. Explanations have typically fallen into two classes: 853

    The first attributes their emergence to various activity-dependent, synaptic plasticity 854

    processes (van Rossum et al., 2000; Song et al., 2005; Lefort et al., 2009; Gilson and Fukai, 855

    2011; Zheng et al., 2013; Buzsáki and Mizuseki, 2014; Effenberger et al., 2015; Scheler, 2017; 856

    Uzan et al., 2018). Given that synaptic diversity was not reduced or distribution shapes 857

    grossly affected in chronically silenced networks (Fig. 2; see also Harms and Craig, 2005; 858

    Harms et al., 2005), this class of explanations would seem to be somewhat unsatisfactory. 859

    Indeed, recent studies reported that in chronically silenced mouse forebrains (Sando et al., 860

    2017) and in hippocampal organotypic cultures prepared from Munc13-1 and Munc13-2 861

    knockout mice (which are essentially devoid of presynaptic release; Sigler et al., 2017), spine 862

    types (mushroom, thin, stubby) are present at normal proportions. 863

  • 31

    A second explanation class attributes these distributions to intrinsic size fluctuations that 864

    contain multiplicative (and additive) components (Yasumatsu et al., 2008; Lowenstein et al., 865

    2011; Kaufman et al., 2012; Statman et al, 2014; Rubinski et al., 2015; Ziv and Brenner, 2018; 866

    Ishii et al., 2018, Humble et al., 2019). Most of such explanations are based on descriptive 867

    models in which size fluctuations are treated statistically without addressing their sources. 868

    One exception is the mesoscopic model of Shomar et al., 2017 used here (Figs. 7,9,10) which 869

    showed how cooperative, stochastic binding and unbinding of synaptic molecules can drive 870

    intrinsic size fluctuations that shape synaptic size distributions (see also Ranft et al., 2017; 871

    Triesch et al., 2018). Here we extended these findings, showing that changes in synaptic 872

    molecule abundance or synapse numbers can affect the magnitude of intrinsic fluctuations 873

    and ultimately synaptic diversity (Figs. 9,10), in good agreement with the recent modeling 874

    study of Triesch and colleagues (2018). These findings thus suggest that activity- 875

    independent, intrinsic size fluctuations, whose source can be traced to the innate dynamics 876

    of synaptic molecules, contribute enormously to excitatory synapse size diversity (see also 877

    Yasumatsu et al., 2008). In fact, they indicate that synaptic size distributions might be 878

    primarily shaped by activity-independent processes, with activity levels mainly setting the 879

    scale, rather than the shape of these distributions. 880

    881

    Synaptic size distribution scaling in silenced networks 882

    The finding that chronic suppression of network activity increases average synaptic size 883

    and scales up synaptic size distributions resembles the homeostatic scaling-up of synaptic 884

    strengths and sizes often observed following acute suppressions of network activity 885

    (reviewed in Turrigiano, 2008; Pozo and Goda, 2010; Chowdhury and Hell, 2018). In most 886

    such studies – both in culture and in vivo – manipulations of activity levels were carried out 887

    in networks in which numerous synapses had already formed (e.g. Turrigiano et al., 1998; 888

    Murthy et al., 2001; Minerbi et al., 2009; Sun and Turrigiano, 2011; Keck et al., 2013; Barness 889

    et al., 2017; Hobbiss et al., 2018) and thus the observed scaling presumably reflected 890

    changes in preexisting synaptic populations. This was clearly not the case in our 891

    experiments, as activity was suppressed long before synapses had formed and thus the 892

    semblance is somewhat superficial. Nevertheless, the larger synapses and broader size 893

    distributions in chronically silenced networks are in line with prior suggestions that synaptic 894

    sizes are continually constrained by activity-dependent processes, and that silencing- 895

  • 32

    associated relaxation of these constraints results in synaptic enlargement (Minerbi et al., 896

    2009; Kaufman et al., 2012; Statmann et al., 2014; Ziv and Brenner, 2018). Indeed, a recent 897

    study (Sando et al., 2017) reported a 30-40% enlargement of spines (and presynaptic 898

    boutons) in chronically silenced mouse forebrains analyzed by light and electron-microscopy. 899

    Interestingly, synapse enlargement was associated with comparable reductions in synaptic 900

    counts and arbor complexity in some (although not all) forebrain regions. Similarly, spine 901

    enlargement in sensory deprived animals was recently shown to be preceded by, and 902

    correlate with spine loss in the same dendritic branches (Barnes et al., 2017) further 903

    supporting our observations on reciprocal relationships between synaptic sizes and 904

    numbers. Understanding how relaxed constraints might reduce synaptic numbers is less 905

    intuitive, but can be understood by appreciating how the weaker bias toward growth 906

    increases the likelihood of small synapses to become even smaller and ultimately lost (Fig. 907

    12). The putative intermediate – a shared pool of synaptic building blocks (Fig. 12) - is in line 908

    with many reports on synaptic competition over limited resources (e.g. Harms et al, 2005; 909

    Mondin et al., 2011; Ramiro-Cortés et al., 2013; Levy et al., 2015; Ryglewski et al., 2017; 910

    Triesch et al., 2018). Obviously, the process proposed in Fig. 12D is not the only determinant 911

    of synaptic numbers and dendrite arborization, since large numbers of synapses ultimately 912

    form on relatively stable dendritic trees even in silenced networks. This might be expected 913

    given that synaptogenesis is governed by many processes not touched on here. Moreover, 914

    while dendritic extension and arborization are influenced by synaptogenesis (Cline and Haas, 915

    2008), these typically precede synaptogenesis and do not strictly depend on it. 916

    In this study we found no evidence that silencing slows PSD-95 exchange kinetics. Other 917

    molecules, however, might be affected differently. For example, prolonged silencing was 918

    shown to slow the exchange kinetics of Shank3/ProSAP2 and Munc13-1 (Tsuriel et al., 2006; 919

    Kalla et al., 2006); Thus, relationships between molecular dynamics and size distributions 920

    might differ among synaptic molecules. Finally, many additional mechanisms have been 921

    implicated in synaptic size/strength distribution scaling (Turrigiano, 2008; Pozo and Goda, 922

    2010; Chowdhury and Hell, 2018) including mechanisms directly involving PSD-95 (e.g. Sun 923

    and Turrigiano, 2011; Chowdhury et al., 2018) further highlighting the explanatory 924

    challenges these phenomena pose. 925

    926

    Activity dependent and independent determinants of synaptic sizes 927

  • 33

    In vivo studies consistently report substantial fluctuations in spine volume or PSD size 928

    over hours to day time scales (Grutzendler et al., 2002; Zuo et al., 2005; Holtmaat et al, 929

    2006; Loewenstein et al., 2011; Cane et al., 2014; Ishii et al., 2018). Moreover, a recent study 930

    (using PSD-95:EGFP) suggests that nanoscale PSD organization in the mouse visual cortex 931

    undergoes remarkable ‘morphing’ over these time scales (Wegner et al., 2018). As these 932

    studies were carried out in live animals, it was not possible to separate intrinsic fluctuations 933

    from activity-dependent synaptic remodeling. In fact, it was recently proposed that synaptic 934

    size fluctuations are the product of ongoing potentiation and depression caused by external 935

    stimuli and internal neuronal activity which ride on top of synaptic scaling related to changes 936

    in activity levels (Keck et al., 2017). Our findings suggest, however, that PSD morphing, size 937

    fluctuations, size distributions and their scaling are tightly interconnected phenomena, 938

    whose root causes relate, at least in part, to the innate dynamics of synaptic molecules. 939

    Interestingly, the size dynamics these produce – noisy additive growth offset by noisy 940

    multiplicative downscaling – inherently ‘solve’ one of the thorny issues of Hebbian forms of 941

    synaptic plasticity, that is the requisite for continuous synaptic size normalization (e.g. Zenke 942

    et al., 2017). Of course, these dynamics introduce thorny issues of their own, such the poor 943

    preservation of size relationships during this normalization process (Minerbi et al., 2009; 944

    Kaufman et al., 2012; Statman et al., 2014). For now, reconciliation of these thorny issues 945

    with common notions on synaptic plasticity still awaits resolution (Mongillo et al., 2017; 946

    Chambers and Rumpel, 2017; Ziv and Brenner 2018). 947

    948

    949

    950

  • 34

    References 951

    Andreae LC, Burrone J (2014) The role of neuronal activity and transmitter release on 952 synapse formation. Curr Opin Neurobiol 27:47-52. 953

    Annis CA, Dowd DKO, Robertson RT (1994) Activity-dependent regulation of dendritic spine 954 density on cortical