<<

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1 Dual contributions of cerebellar-thalamic networks to learning and offline consolidation of a 2 complex motor task. 3 4 Andres P Varani1, Romain W Sala1, Caroline Mailhes-Hamon1, Jimena L Frontera1, Clément 5 Léna1*, Daniela Popa1* 6 7 (1) Institut de biologie de l'Ecole normale supérieure (IBENS), Ecole normale supérieure, CNRS, 8 INSERM, PSL Research University, 75005 Paris, France 9 * these authors jointly directed the work 10 11 12 SUMMARY 13 14 The contribution of to motor learning is often considered to be limited to 15 adaptation, a short-timescale tuning of reflexes and previous learned skills. Yet, the cerebellum 16 is reciprocally connected to two main players of motor learning, the and the basal 17 ganglia, via the ventral and midline respectively. Here, we evaluated the contribution 18 of cerebellar neurons projecting to these thalamic nuclei in a skilled locomotion task in mice. In 19 the cerebellar nuclei, we found task-specific neuronal activities during the task, and lasting 20 changes after the task suggesting an offline processing of task-related information. Using 21 pathway-specific inhibition, we found that dentate neurons projecting to the midline thalamus 22 contribute to learning and retrieval, while interposed neurons projecting to the ventral thalamus 23 contribute to the offline consolidation of savings. Our results thus show that two parallel 24 cerebello-thalamic pathways perform distinct computations operating on distinct timescales in 25 motor learning. 26

1

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

27 28 INTRODUCTION 29 Learning to execute and automatize certain actions is essential for survival and animals 30 have indeed the ability to learn complex patterns of movement with great accuracy to improve 31 the outcomes of their actions(Krakauer et al., 2019). The acquisition of a motor skill is often 32 divided into at least two components (Krakauer et al., 2019; Seidler, 2010): 1) sequence learning, 33 which is needed when series of distinct actions are required, and 2) adaptation which 34 corresponds to the ability to adapt a previous competence, and typically takes place when 35 motor actions yield unexpected sensory outcomes. The neurobiological substrate of motor skills 36 involves neurons distributed in the cortex, and cerebellum, which each implement 37 a distinct learning algorithm (Doya, 1999). Supervised learning takes place in the cerebellum 38 (Raymond and Medina, 2018), and is central for the adaptation of skills such as oculomotor 39 movements (Herzfeld et al., 2018; Nguyen-Vu et al., 2013; Yang and Lisberger, 2014), reaching 40 (Hewitt et al., 2015), locomotion (Darmohray et al., 2019; Morton and Bastian, 2006), as well as 41 conditioned reflexes (Clopath et al., 2014; Longley and Yeo, 2014). The cerebellum is thought to 42 form associations between actions and predicted sensory outcome at short-time scale (typically 43 under one second), which are seen as internal models (Ito, 2008). The involvement of the 44 cerebellum in learning of complex actions involving sequences of movements is far less well 45 understood and still controversial (Baetens et al., 2020; Bernard and Seidler, 2013; Krakauer et 46 al., 2019; Seidler et al., 2002). 47 Motor skills are generally progressively acquired ((e.g. Karni et al., 1998)). Several phases 48 of motor learning, with distinct behavioral and anatomo-functional hallmarks have been 49 described: in the early phase (Acquisition), fast improvements of performance take place, but 50 they are susceptible to interferences; following a transitional Consolidation phase, the behavior 51 reaches a Maintenance phase, where it becomes less variable, automatic, resistant to 52 interferences, and may rely on different sets of brain structures compared to the initial training 53 ((e.g. Brashers-Krug et al., 1996; Korman et al., 2003; Muellbacher et al., 2002)); some of the 54 consolidation occurs offline during rest, which may be sufficient to change the recruitment of 55 brain regions in the task execution (Shadmehr and Holcomb, 1997). Motor memories may also 56 persist in the form of savings, which facilitate re-learning of the task at a later point in time

2

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

57 (Huang et al., 2011; Mauk et al., 2014). Overall, learning is a process which is distributed in time 58 and space. 59 Understanding the contribution of the cerebellum to learning thus requires taking into 60 account its integration in brain-scale circuits including the cortex and basal ganglia (Caligiore et 61 al., 2017). In the mammalian brain, both cerebellum and basal ganglia receive the majority of 62 their afferences from the cerebral cortex and send projections back to the cortex via 63 anatomically and functionally segregated channels, which are relayed by mostly non-overlapping 64 thalamic regions (Bostan et al., 2013; Hintzen et al., 2018; Proville et al., 2014). Furthermore, 65 anatomical and functional reciprocal di-synaptic connections have been demonstrated between 66 the basal ganglia and the cerebellum (Bostan and Strick, 2010; Carta et al., 2019). The 67 projections from the cerebellum to motor cortex and the are relayed through distinct 68 thalamic regions, respectively the ventral thalamus and intralaminar thalamus (Chen et al., 69 2014; Steriade, 1995), suggesting distinct contributions of these diencephalic projections of the 70 cerebellum. 71 In the present study, we hypothesized that the cerebellum may contribute to some phases 72 of learning in a complex motor task via its projections to the motor cortex and/or the basal 73 ganglia. We thus focused on the dentate and interposed cerebellar nuclei and their projections 74 to the centrolateral (intralaminar) thalamus and ventral anterior lateral complex (motor 75 thalamus), which respectively relay their activity to the striatum and the motor cortex (Chen et 76 al., 2014; Gornati et al., 2018; Proville et al., 2014). We first looked for task-related activities in 77 the cerebellum and thalamus using chronic in vivo extracellular recordings in the cerebellar 78 nuclei, intralaminar and ventral thalamus, along the learning of the motor task. Second, we 79 examined the contribution of cerebellar nuclei and cerebello-thalamic pathways to learning 80 using chemogenetic disruptions either during or after the learning sessions. 81

3

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

82 RESULTS 83 In the present study, we used the paradigm of the accelerating rotarod, where the animals 84 walk on an accelerating rotating horizontal rod; the animals must develop locomotion skills to 85 avoid falling from the rod. To examine the involvement of the cerebellum in the accelerating 86 rotarod task along the different phases of learning, we first evaluated the neuronal activity in 87 the cerebello-cortical and cerebello-striatal pathways which primarily involve the interposed 88 nucleus and ventral anterior lateral thalamus (VAL), as well as the and the 89 centrolateral thalamus (CL), respectively. For this purpose, we implanted C57/Bl6J mice (n=16) 90 with microelectrode arrays that were designed to target either the cerebellar or the thalamic 91 nuclei. Then, mice were trained for seven consecutive days on an accelerating rotarod with 92 seven trials a day, each daily session preceded and followed by 10 minutes of free locomotion in 93 an open-field (Fig 1A,B). The animals showed significant learning during the first day as 94 evidenced by an increase in the latency to fall from the rotating rod between the first and last 95 trial (Fig 1A, bottom, Suppl. Table 1). During the second day, there was still a significant increase 96 between the first and last trial of the session (Fig 1A, bottom, Suppl. Table 1) and also a global 97 increase compared with the first day. The improvement in the later days was more gradual, 98 comparison between first and last trial did not reach significance; the asymptotic values of 99 latency to fall were reached on the 4th day and maintained for the 3 following days. Based on 100 these observations, we defined three phases of learning: Acquisition on day 1, Consolidation on 101 days 2, 3, 4 and Maintenance on days 5, 6, 7 (Durieux et al., 2012; Yin et al., 2009). 102 We first examined to which extent the firing rate of the cerebellar nuclei (Dentate and 103 Interposed) was modified during the running sessions on the rotarod. In most of the cells and 104 for most of the phases of learning, the rate varied between the open-field sessions, the rotarod 105 trials and the inter-trial episodes (Fig 1C, D). We then determined how the firing rate was 106 changing in the rotarod apparatus depending on whether the mice were running or immobile, 107 by comparing the distribution of instantaneous rate during each trial and the immediately 108 preceding inter-trial during which the animal rested at the bottom of the apparatus for 300s (Fig 109 1E). The cells in the two cerebellar nuclei exhibited a significant difference in the distribution of 110 instantaneous firing rate for at least one of the trials (Fig 1E, Suppl. Table 2). Moreover, we found 111 that the number of trials with significant change in the distribution of instantaneous firing rate 4

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

112 of the cells increased from the Acquisition phase to the later phases of learning (Fig 1E, Suppl. 113 Table 2). This indicates that in the rotarod apparatus, the firing of Dentate and Interposed 114 neurons was not affected only during the early phase of learning but instead showed changes 115 along the execution of the rotarod task. 116 To test if the changes of firing rate reported above reflect state differences between 117 locomotion and immobility in the rotarod apparatus or alternatively result from an engagement 118 of the cells in the task, we compared the firing rate during the trials (1st, 4th or 7th trials) with 119 periods of locomotion in the preceding open-field (Fig 1F). We observed that the firing rate in 120 the Dentate nucleus was decreased in all phases, particularly in late trials (Fig 1F top left, Suppl. 121 Table 3), while the reduction of firing in the Interposed nucleus was only limited to the last trials 122 of Consolidation and Maintenance phase (Fig 1F bottom left, Suppl. Table 3). In a separate set of 123 mice, we then examined the evolution of cell firing in the CL and VAL thalamus. We observed a 124 ubiquitous increase of firing rate in CL with the exception of the last trial in the Acquisition 125 phase (Fig 1F top right, Suppl. Table 3). Similarly the VAL thalamus showed increases of firing 126 rate during Consolidation and Maintenance phases in all the trials, but in the Acquisition phase 127 the increase was only observed for the first trial (Fig 1F bottom right, Suppl. Table 3). Thus the 128 main modulation of firing rate in rotarod locomotion (compared to open-field) was a reduction 129 in the cerebellar nuclei which developed as the learning progressed, while increased firing rates 130 were observed for most of the CL and VAL neurons. However, the change in firing rate while 131 walking in the open-field vs rotarod might reflect a difference in the involvement of the cells 132 these two conditions, or more trivially it might reflect differences between the locomotion 133 speed in the rotarod and the open-field.

5

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

134 135 136 Figure 1. The engagement of the cerebellar nuclei during the accelerating rotarod task changes along the 137 learning protocol. a) Scheme of the accelerating rotarod protocol (top) and latency to fall (bottom) during the 138 accelerating rotarod across trials along days in mice recorded over the 7 days (**p<0.01, *p<0.05 t-test 139 comparing trial 1 vs trial 7). b) Example of electrode placements in the Dentate (top) and Interposed nuclei 140 (bottom). c) Firing rate histograms showing the evolution of activity in representative cells for Dentate (left) and 141 Interposed (right) nuclei during Acquisition (top), Consolidation (middle) and Maintenance (bottom). Rotarod 142 trials are shown in colors, while the open-Field before (OF1) and after (OF2) the rotarod session are shown in 143 grey, and the intertrial periods are shown in black. d) Raster plots showing the activity of representative cells for 144 dentate (left) and interposed (right) during trials for acquisition, consolidation and maintenance. e) Distribution 145 of numbers of trials in which cells showed a significiant modulation of firing rate between trial and inter-trial 146 periods during acquisition (Acq), consolidation (Cons) and maintenance (Maint) phase for Dentate (top) and 147 Interposed (bottom) (*p<0.05, ***p<0.001 Mann-Whitney test, Holm-Sidak corrected for multiple comparison). 148 f) Evolution of the average firing rate (mean +/- SEM), normalized by subtracting the average firing rate during 149 the active part of the open-field session before the first rotarod trial, during Acquisition, Consolidation and 150 Maintenance for Dentate (top left), Interposed (bottom left) nuclei, centrolateral thalamus (CL, top right), 151 ventrolateral thalamus (VAL, bottom right) (*p<0.05, **p<0.01, ***p<0.001 Dunett Posthoc test). 152 153 154 To test the link between the locomotion speed and the cells’ firing, we performed a linear 155 regression of the average firing rate as a function of the locomotion speed in the rotarod and in 156 the open-field (Fig 2). Cells in the Dentate and Interposed nuclei exhibited negative correlation 157 with the locomotion speed on the rotarod (Fig 2A,B, Suppl. Table 4). In contrast, on the open- 158 field, much weaker correlations with the locomotion speed were found in the Dentate (Fig 2C, 159 top) and Interposed (Fig 2C, bottom) cells, with significantly more negative slopes on the rotarod 160 compared to the open-field in all learning phases (Fig 2C, Suppl. Table 5). Contrarily to the

6

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

161 preeminence of negative correlation observed in the cerebellar nuclei in the rotarod, both 162 positive and negative correlations were found in CL and VAL neurons (Fig Sup1A). When 163 examining the slope of the correlations in cells with significant correlations with rotarod speed 164 (Fig 2D), we found slightly more negative slopes in the Maintenance phase in the Dentate cells 165 (Fig 2D, top left, Suppl. Table 6) and slightly more negative slopes in the Acquisition phase 166 compared to the later phases in Interposed cells (Fig 2D, bottom left, Suppl. Table 6). Stronger 167 negative slopes in the rotarod were also found in the CL thalamus during Consolidation and 168 Maintenance phases (Fig 2D, top right, Suppl. Table 6), whereas slopes in the VAL thalamus 169 remained stable along learning (Fig 2D, bottom right, Suppl. Table 6). We found monotonous 170 relationships between statistically significant slopes and their associated Pearson’s correlation 171 coefficient for Dentate (Fig 2E, top, Suppl. Table 7) and Interposed (Fig 2E, bottom, Suppl. Table 172 7) cells during all the phases -with the exception of Acquisition phase for the Dentate nucleus- 173 indicating that the stronger modulations of firing rate by the locomotion speed tend to be more 174 consistent. Similarly, we also observed monotonous relationships for CL and VAL thalamus (Fig 175 Sup1B, Suppl. Table 7). In order to analyze the evolution of the modulation by the speed 176 between structures, we then compared their distributions of Pearson’s correlation coefficients 177 (which are dimensionless quantities) in all phases. We observed that Dentate-CL overlap index 178 (ƞ) showed more similar Pearson’s correlation coefficients during Consolidation and 179 Maintenance than in Acquisition (Fig 2F, top, Suppl. Table 8). In contrast, Interposed-VAL overlap 180 index (ƞ) showed similar Pearson’s correlation coefficients along all phases (Fig 2F, bottom, 181 Suppl. Table 8).

7

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

182 183 Fig 2. Dentate and Interposed nuclei display a context-dependent sensitivity to speed, with a negative 184 correlation between cerebellar activity and rotarod speed. a) Scatterplot showing the average firing rate per 185 speed bins during the different trials of rotarod (left) versus open-field (right) session, for Dentate (top) and 186 Interposed (bottom) nucleus (linear regression lines are shown with a 95% confidence interval) (*p<0.05, 187 ***p<0.001 Pearson correlation). b) Scatter plot showing the slope of linear regression explaining the firing 188 rate by the speed, in the open-field versus on the rotarod for each neuron in Dentate (top) and Interposed 189 (bottom) nucleus during Acquisition (Acq.), Consolidation (Cons.) and Maintenance (Maint.). The diagonal 8

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

190 dotted line represents equality between the slope in the open-Field and rotarod. Marginal axes show the 191 histograms of the distributions of slopes, smoothed using a Gaussian kernel density estimate (ς=0.05). c) 192 Boxplots displaying the distributions of linear regression slopes of neurons from Dentate (top) and Interposed 193 (bottom) during open-field (OF) and rotarod (Rtrd) (***p<0.001 Wilcoxon test). d) Significant linear regression 194 slopes for rotarod (mean +/- SEM) for Dentate (top left), centrolateral thalamus (CL, top right), Interposed 195 (bottom left) and ventral anterior lateral thalamus (VAL, bottom right) during Acquisition (A), Consolidation (C) 196 and Maintenance (M) (*p<0.05, **p<0.01, ***p<0.001 Tukey Posthoc test). e) Scatter plot showing the 197 correspondence between the slope of linear regression on rotarod versus the associated Pearson correlation 198 coefficient for each neuron for Dentate (top) and Interposed (bottom) nucleus during Acquisition (Acq.), 199 Consolidation (Cons.) and Maintenance (Maint.). The lines represent the isotonic regression of the Pearson’s r 200 by the slope on rotarod (***p<0.001 Spearman Rank test). f) Density of Pearson’s r coefficients for Dentate 201 and CL, (top) and Interposed and VAL, (bottom) for Acquisition, Consolidation and Maintenance phase, the 202 overlapping area is represented in blue and is associated to the overlapping index η (*p<0.05, **p<0.01, 203 ***p<0.001 Mann Whitney test); a barplot of the value of η is shown on the right. 204

205 206 Fig Sup1. Centrolateral and ventral anterior lateral thalamus display a sensitivity to rotarod speed. a) Scatter 207 plot showing the slope of linear regression explaining the firing rate by the speed, in the open-field versus the 208 rotarod for each neuron in centrolateral thalamus (CL, top) and ventral anterior lateral thalamus (VAL, bottom) 209 during Acquisition (Acq.), Consolidation (Cons.) and Maintenance (Maint.). Marginal axes show the histograms 210 of the distributions of slopes, smoothed using a Gaussian kernel density estimate (ς=0.05). b) Scatter plot 211 showing the correspondence of slope of linear regression on rotarod versus the associated Pearson correlation 212 coefficient for each neuron for CL (top) and VAL (bottom), during Acquisition (Acq.), Consolidation (Cons.) and 213 Maintenance (Maint.). The lines represent the isotonic regression of the Pearson’s r by the slope on rotarod 214 (**p<0.01, ***p<0.001 Spearman Rank test). 215

9

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

216 Overall, these results indicate the presence of a number of cells in the cerebellar nuclei 217 modulated during the rotarod task. This modulation consists mostly in a reduction in firing rate 218 proportional to the locomotion speed on the rotarod, which is not observed in the open-field 219 locomotion, suggesting a specific engagement of these cells in the task execution. Moreover, this 220 modulation is observed in all phases of learning. In contrast, the modulation of units in the CL 221 thalamus exhibit stronger changes in the course of learning, suggesting changes in the encoding 222 of the task in the motor circuits; the observations in the VAL thalamus suggest an earlier 223 engagement of this structure since the modulation was stable along the learning phases. In 224 addition to these observations, we found that the cells in the Dentate and CL thalamus exhibited 225 more similar modulations by the speed in late phases (Consolidation and Maintenance), while 226 the similarity of modulation by speed in Interposed and VAL thalamus cells remained more 227 stable across all phases. This suggests that Dentate-CL and Interposed-VAL pathways have a 228 differential engagement during motor learning, the former exhibiting rather an increased 229 engagement in the course of learning. 230 231 Neuronal activity during rest is modified during and after learning 232 Motor skill learning undergoes consolidation between episodes of learning, and we 233 therefore examined the activity of the motor circuits during periods of immobility before, 234 between and after rotarod trials (Fig 3). For open-field sessions, a reduction of the mean firing 235 rate was found in all phases of learning in the Interposed nucleus and in the Maintenance phase 236 in the Dentate, while little if any change was observed in the thalamus (Fig 3A, Suppl. Table 9). 237 Moreover, we observed that this reduction was more important during Consolidation and 238 Maintenance for Interposed cells (Suppl. Table 10), whereas it remained stable during all the 239 phases for Dentate cells (Suppl. Table 10). In a smaller set of animals, we recorded the firing rate 240 in the open-field quiet periods and in the resting period between rotarod trials (Inter-trials). In 241 these cases, we observed different patterns of discharge during rest (Fig 3B, Suppl. Table 11): in 242 the cerebellar nuclei the firing rate during the quiet period between trials and after learning in 243 the open-field decreased in the Consolidation and Maintenance phases compared to the firing 244 in the open-field before the learning sessions of the day (Suppl. Table 11). In the VAL thalamus 245 the firing rate increased during the inter-trial resting periods and returned progressively to the 10

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

246 baseline in the open-field after learning, while in CL thalamus this evolution was only observed 247 in Acquisition and Maintenance phases (Suppl. Table 11). Overall, these results indicate the 248 presence of lasting changes in the cerebellum “offline” firing during quiet periods during and 249 following the learning session, while thalamic neurons only exhibited changes in firing at rest 250 only at shorter timescale, between trials.

251 252 253 Fig 3. The quiet state activity in the cerebellar nuclei and in the thalamus changes during the learning 254 protocol. a) Boxplots showing the distribution of change in mean firing rate between open-field (OF) before 255 and after rotarod, during acquisition, consolidation and maintenance for Dentate, Interposed, centrolateral 256 thalamus (CL) and ventral anterior lateral thalamus (VAL) (***p<0.001 Tukey Posthoc test). b) Evolution of the 257 mean quiet state firing rate (mean +/- SEM) in open-field sessions and Inter-trial periods during Acquisition, 258 Consolidation and Maintenance for Dentate (top left), Interposed (top right), CL (bottom left) and VAL (bottom 259 right). Mean firing rate was normalized by subtracting the open-field before (OF1) (*p<0.05, **p<0.01, 260 ***p<0.001 Dunett Posthoc test). 261 11

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

262 263 Partial inhibition of the cerebellar nuclei using hM4D(Gi)-DREADD does not impact basic 264 motor abilities. 265 While the analysis above shows that many neurons exhibit some modulation of firing 266 during and after the accelerated rotarod, the contribution of this activity to learning and 267 behavior remains unknown. To evaluate this contribution, we used a chemogenetic approach 268 (Roth, 2016) using the inhibitory DREADD (hM4D-Gi) activated by the synthetic drug Clozapine- 269 N-Oxide (CNO). 270 In order to validate this approach, mice were injected with AAV5-hSyn-hM4D(Gi)-mCherry 271 (DREADD) or AAV5-hSyn-EGFP (Sham) and implanted with microelectrode arrays in the 272 cerebellar nuclei (Fig 4A-H). Post-hoc histology confirmed the position of the electrodes (Fig 4E) 273 and showed that the expression of AAV5-hSyn-hM4D(Gi)-mCherry was confined to the 274 cerebellar nuclei and restricted to the neuronal membrane, with 78% of the cells expressing 275 hM4D(Gi)-DREADD in the cerebellar nuclei (Fig 4B-D). A week after surgery, the neuronal 276 activity was recorded in the cerebellar nuclei in an open-field arena before and after accelerated 277 rotarod sessions. Neither saline (SAL) in DREADD mice nor CNO injection (1 mg/kg) in Sham mice 278 affected the firing rate, while the same dose of CNO induced a significant decrease in the firing 279 rate of DREADD mice during Acquisition (day 1), Consolidation (day 4) and Maintenance (day 7) 280 phases (Fig 4G and H, Suppl. Table 12). 281 We then examined whether the reduction of cerebellar nuclei firing impacted on the 282 spontaneous motor activity, strength and motor coordination. Analysis of the open-field 283 locomotor activity revealed that velocity was not altered by CNO or SAL injection in DREADD or 284 Sham mice (Fig Sup2A, Suppl. Table 26). In addition, no significant differences were observed 285 between the experimental groups in the fixed speed rotarod (Fig Sup2B, Suppl. Table 27), 286 footprint patterns (Fig Sup2C, Suppl. Table 28), grid test (Fig Sup2D, Suppl. Table 28), horizontal 287 bar (Fig Sup2E, Suppl. Table 28) and vertical pole (Fig Sup2F, Suppl. Table 28) indicating that 288 fatigue, coordination and strength are not affected by the reduction of cerebellar nuclei firing 289 induced by 1mg/kg CNO. These results thus indicate that the reduction of the cerebellar nuclei 290 firing induced by the CNO had a limited impact on the basic motor abilities of the mice.

12

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

291 292 Fig Sup2. Cerebellar nuclei inhibition did not affect execution and fatigue, locomotion, motor coordination, 293 balance and strength. a) Locomotor activity (Velocity) in DREADD and non-DREADD (Sham) injected mice 294 after CNO or SAL injection during open-field sessions before (OF1) and after (OF2) rotarod for Acquisition, 295 Consolidation and Maintenance (**p<0.01 t-test OF1 vs OF2). b) Latency to fall during fixed speed rotarod (5, 296 10, 15, 20, 25 r.p.m.) for all experimental groups. One way repeated measure ANOVA was performed on 297 averaged values for all the speed steps in each experimental group followed by a Tukey Posthoc pairwise 298 comparison. c) Footprint patterns were quantitatively assessed for 3 parameters as shown on representative 299 footprint patterns (top) for all experimental groups. Three parameters are represented graphically: linear 300 movement (bottom left), sigma (bottom middle) and alternation coefficient (bottom right). d) Latency 301 reflecting the time before falling from the grid. 30 seconds of cut-off of was established as the maximum 302 latency (dotted line on figure). e) Latency to cross the horizontal bar (balance beam test) for all experimental 303 groups. f) Latency to reach home cage in vertical pole test for all experimental groups. *p<0.05 One Way 304 ANOVA followed by a PostHoc Tukey test. n indicates the number of mice. 305 306 307 Cerebellar inactivation by hM4D(Gi)-DREADD during or after task impact on motor learning. 308 To test the effect of a reduction of cerebellar output on motor learning, we first examined 309 the impact of cerebellar nuclei inhibition by injecting CNO (1 mg/kg) each day before the first 310 trial of an accelerated rotarod session (Fig 4I,K, Suppl. Table 13,14,15). This treatment did not 311 affect significantly the learning on the first day, but reduced the performance on the following 312 days compared to the control groups (Sham+SAL, Sham+CNO: Sham mice which received either 313 SAL or CNO, and DREADD+SAL: DREADD mice which received SAL, Suppl. Table 13). During the 314 Consolidation phase, the performance of the initial trial, but not the last ones, of each day was 315 reduced (Suppl. Table 15), indicating that the DREADD+CNO animals could compensate within 316 each day the low performance of the first trials. These poor performances were associated with

13

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

317 a significant loss of performance between the last trial of one day and the following day (Suppl. 318 Table 14), indicating a defect in the consolidation of motor learning in the DREADD+CNO mice. 319 During the Maintenance phase, the performance of the DREADD+CNO group remained lower 320 than the control groups both for the first and last trials of the task, and no further improvement 321 of the performance of this group was observed during this phase (Suppl. Table 15). These results 322 show that the inhibition of the cerebellar nuclei impacts on learning consolidation in the 323 Consolidation phase, and on the execution of the task when reaching the Maintenance phase. 324 The action of CNO administration lasts longer than the learning trials; therefore the results 325 above do not differentiate the action of cerebellar inhibition during and after the learning trials. 326 We therefore injected another set of mice with CNO after the training sessions (Fig 4J,L, Suppl. 327 Table 13,14,15). Cerebellar nuclei inhibition after learning in the Acquisition and Consolidation 328 phases reduced the performance on the first trials of the next day, but did not prevent learning 329 within days (Suppl. Table 14), indicating a defect in the offline consolidation, which could be 330 overcome by training during the day. The performance on the last day of the Consolidation 331 phase of DREADD+CNO mice was not different from the mice of the control groups (Suppl. Table 332 13). These results indicate that the cerebellum participates to the offline consolidation of the 333 accelerated rotarod learning.

14

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

334 335 336 Fig 4. Cerebellar inactivation impairs the performance in the Consolidation and Maintenance but not 337 Acquisition phases of motor learning task. a) Scheme showing electrode position and viruses injections. b) 338 Coronal section of the cerebellum showing hM4Di-DREADD expression in the three cerebellar nuclei. c) 339 Representative confocal image showing hM4Di-DREADD expression on neuronal membranes. d) Image 340 showing DAPI positive neurons expressing hM4Di-DREADD. e) Coronal section of the cerebellum showing 341 electrode placement (yellow dotted line)(left) and cells expressing hM4Di-DREADD nearby the electrode 342 (right). f) Examples of spike shapes obtained from spike sorting in cerebellar nuclei (mean +/- SD). g) Examples 343 of high-pass filtered traces of cerebellar nuclei recordings before and after CNO injection. h) Boxplots showing 344 the mean firing rate of neurons recorded in DREADD and non-DREADD injected mice after CNO or SAL 345 injection during Acquisition (Acq.), Consolidation (Cons.) and Maintenance (Maint). Cerebellar nuclei firing 346 rate was reduced after 1 mg/kg of CNO injection in DREADD injected mice (***p<0.001 paired t-test). Small 347 variations were observed in Sham+SAL group (*p<0.05, ***p<0.001 paired t-test). i) Daily of injections of CNO 348 before trial 1 in DREADD injected mice induced impairment in motor learning during Consolidation and 349 Maintenance but not Acquisition phase (*p<0.05, ***p<0.001; repeated measure ANOVA Group effect). j) 15

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

350 CNO injections after Acquisition phase (30 min after trial 7) induced impairment in motor learning during 351 Consolidation phase (*p<0.05, **p<0.01; repeated measure ANOVA Group effect). k,l) Latencies to fall in trial 352 1 and 7 were compared for each day (horizontal comparison lines trial 1 vs trial 7; *p<0.05, **p<0.01, 353 ***p<0.001 paired t-test). Latencies to fall in the trial 7 of a day and trial 1 of the next day were compared 354 (horizontal comparison lines trial 7 vs trial 1; *p<0.05, ***p<0.001 paired t-test). Comparisons between 355 groups for trials 1 and trials 7 were also represented (vertical comparison lines, respectively trial 1 difference 356 to controls and trial 7 difference to controls; *p<0.05, **p<0.01, ***p<0.001 t-test). Data represents mean ± 357 S.E.M. n indicates the number of mice. 358 359 360 Selective cerebello-thalamic pathways differentially impact motor learning. 361 The cerebellar nuclei project to a wide array of targets; to examine whether cerebellar 362 neurons with distinct targets differentially contribute to the rotarod learning and execution, an 363 AAV5-hSyn-DIO-hM4D(Gi)-mCherry virus expressing an inhibitory DREADD conditioned to the 364 presence of Cre-recombinase was injected into the Dentate and Interposed cerebellar nuclei, 365 while a retrograde CAV-2 virus expressing the Cre recombinase was injected either in the CL or 366 in the VAL, which respectively relay cerebellar activity to the striatum and cerebral cortex. We 367 could then express the inhibitory DREADD either in the Dentate neurons projecting to the CL 368 (Dentate-CL, Fig 5A) or in Interposed neurons projecting to the VAL thalamus (Interposed-VAL, 369 Fig 5D). Since we did not observe an effect of CNO in Sham mice, we only compared DREADD- 370 injected animals receiving either CNO or SAL. 371 To examine the impact of the inhibition of spontaneous locomotion, motor coordination 372 and strength, we first examined locomotor activity in open-field experiments (Fig Sup3A, Suppl. 373 Table 29). Analysis of the open-field locomotor activity revealed that velocity was generally not 374 systematically by CNO or SAL injection in DREADD or Sham mice (Fig Sup3A, Suppl. Table 29); 375 significant differences in the first open-field (OF1) for Interposed-VAL+CNO and Dentate- 376 CL+CNO indicating respectively slightly higher and lower speed in day 1 and 4, compared to 377 control groups. In addition, no significant differences were observed between Dentate-CL+CNO 378 in the fixed speed rotarod test (Fig Sup3B, Suppl. Table 30). We found a decrease in the latency 379 to fall for 15 and 25 r.p.m. in the Interposed-VAL+CNO group. No significant differences were 380 observed between the experimental groups for footprint patterns (Fig Sup3C, Suppl. Table 31), 381 grid test (Fig Sup3D, Suppl. Table 31), horizontal bar (Fig Sup3E, Suppl. Table 31) and vertical 382 pole (Fig Sup3F, Suppl. Table 31) indicating that coordination and strength are not affected by 383 the inhibition of cerebellar-thalamic pathways induced by 1mg/kg CNO.

16

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

384 385 Fig Sup3. Dentate-centrolateral and Interposed-ventral anterior lateral inhibition did not affect execution 386 and fatigue, locomotion, motor coordination, balance and strength. a) Locomotor activty (Velocity) in 387 DREADD injected mice after CNO or SAL injection during open-fields sessions before (OF1) and after (OF2) 388 rotarod for Acquisition, Consolidation and Maintenance (**p<0.01 t-test OF1 vs OF2). b) Latency to fall during 389 fixed speed rotarod (5, 10, 15, 20 r.p.m.) for all experimental groups. One way repeated measure ANOVA was 390 performed on averaged values for all the speed steps in each experimental group followed by a Tukey Posthoc 391 pairwise comparison. c) Footprint patterns were quantitatively assessed for 3 parameters as shown on 392 representative footprint patterns (top) for all experimental groups. Three parameters are represented 393 graphically: linear movement (bottom left), sigma (bottom middle and alternation coefficient (bottom right). 394 d) Latency reflecting the time before falling from the grid. 30 seconds of cut-off of was established as the 395 maximum latency (dotted line on figure). e) Latency to cross the horizontal bar (balance beam test) for all 396 experimental groups. f) Latency to reach home cage in vertical pole test for all experimental groups. *p<0.05 397 One Way ANOVA followed by a PostHoc Tukey test. DCN, ; CL, centrolateral thalamus; 398 VAL, ventral anterior lateral thalamus. n indicates the number of mice. 399 400 401 We then examined how the rotarod learning was impaired by inhibiting the cerebello- 402 thalamic pathways during and after the task (Fig 5A-N), as for full cerebellar nuclei inhibition 403 (see above). Inhibition of the Dentate-CL pathway during the task (Fig 5B,G, Suppl. Table 404 16,17,18) produced a progressive departure from the performance of the control group, yielding 405 to a strong reduction of performance in the Maintenance phase. In contrast, when the inhibition 406 took place after the task (Fig 5C,H, Suppl. Table 16,17,18), the performances remained similar to 407 the control group (although incidentally, significant Group×Trial interaction or posthoc difference 408 between trial 1 was observed on some days).

17

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

409 In contrast, the pattern induced by the inhibition of the Interposed-VAL pathway differed 410 sensibly. First, when inhibition took place during the task (Fig 5E,K, Suppl. Table 16,17,18), there 411 was a strong loss of performance from the last trial of one day to the first trial of the following 412 day. This overnight loss was compensated by a fast relearning up to similar levels as the control 413 mice during the Consolidation phase. In the Maintenance phase, the learning saturated at lower 414 level than control mice but similar to the Maintenance phase of the Dentate-CL group. Second, 415 when inhibition took place after the task (Fig 5F,L, Suppl. Table 16,17,18), a similar overnight 416 decrease of performance was observed during the Consolidation phase. These results are 417 consistent with a contribution of the offline activity (as observed for full cerebellar nuclei 418 inhibition, Fig 4L Suppl. Table 14,15) to the overnight consolidation of the motor learning. 419 We then shifted the treatment of all mice to SAL in the Maintenance phase (Fig 5I,J,M,N, 420 Suppl. Table 19,20,21). Interestingly, the mice previously treated with CNO continued to exhibit 421 strongly reduced performances on the first trial of each day, but the performances then raised to 422 the control level within each day (associated with significant within-day learning during the 423 Maintenance phase, Fig 5I,M, Suppl. Table 19). This indicates that the mice still failed to 424 consolidate new learning in the Maintenance phase, even in the absence of cerebellar 425 inhibition. 426 Mice which learned the task while either Dentate-CL or Interposed-VAL neurons were 427 inhibited both showed reduced performances compared to controls. To examine to which 428 extent the reduced performance were associated to a deficit of execution versus a reduced 429 learning, we switched the treatment between the two groups (Fig 5I,J,M,N, Suppl. Table 430 19,20,21): mice that received SAL during 7 previous days were then administered with CNO and 431 vice versa. In both groups, mice that received CNO during learning did not exhibit a sudden 432 improvement of performance upon replacement by SAL. In the case of the Dentate-CL (Fig 5I,J, 433 Suppl. Table 19,20,21), the performances only a mildly improved during the two days of reverse 434 treatment. In the case of Interposed-VAL (Fig 5M,N, Suppl. Table 19,20,21), the mice continued 435 to show a pattern of overnight loss of performance and relearning, which yielded similar final 436 values at the end of each day as under CNO. Reciprocally, in both groups, mice that learned 437 under SAL and received CNO exhibited a sudden drop in performance in the first day of reverse 438 treatment. In the case of the Dentate-CL (Fig 5I,J), no improvement within days was then 18

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

439 observed, while the interposed-VAL (Fig 5M,N, Suppl. Table 19,20,21) mice exhibited the same 440 alternation of within-day increases and overnight decrease of performances, as observed in the 441 CNO-treated interposed-VAL mice before the reversal of treatment. These results show that the 442 impact of cerebellar-thalamic inhibition during learning cannot be readily reversed by alleviating 443 the inhibition. Cerebellar-thalamic inhibition after learning reduces the performance and 444 prevents re-learning, but with a different pattern in the two pathways, the Interposed-VAL 445 showing within-day improvement and overnight decline in performances absent in the Dentate- 446 CL pathway.

447 19

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

448 449 Fig 5. Dentate-centrolateral thalamus and Interposed-ventral anterior lateral thalamus pathway 450 inactivation impairs consolidation and maintenance but not acquisition phase of motor learning task. 451 Schemes showing AVV5-hSyn-DIO-hM4D(Gi)-mCherry and CAV2-Cre-GFP virus injections in Dentate and 452 centrolateral thalamus (CL) (a) or Interposed and ventral anterior lateral thalamus (VAL) (d), respectively. 453 Example of Dentate-CL (a, bottom) and Interposed-VAL (d, bottom) neurons expressing GFP (left, 500µm) and 454 m-Cherry (right, 50 µm). b) Daily of injections of CNO before trial 1 in Dentate-CL pathway injected mice 455 induced global impairments during Maintenance but not Acquisition and Consolidation phase (*p<0.05, 456 **p<0.01, ***p<0.001; repeated measure ANOVA Group effect). c) CNO injections after Acquisition phase (30 457 minutes after trial 7) in Dentate-CL pathway did not induce alterations during Consolidation or Maintenance 458 phase (#p<0.05, ##p<0.01, ###p<0.001; repeated measure ANOVA Group effect). e) Daily injections of CNO 459 before trial 1 in Interposed-VAL pathway injected mice induced early impairments during Consolidation and 460 Maintenance but not Acquisition phase (*p<0.05, **p<0.01, ***p<0.001; repeated measure ANOVA Group 461 effect). f) CNO injections after Acquisition phase (30 minutes after trial 7) in Interposed-VAL pathway induced 462 early impairments during Consolidation and Maintenance phase (#p<0.05, ##p<0.01, ###p<0.001; repeated 463 measure ANOVA Group effect). g,h,k,l) Latencies to fall in trial 1 and 7 were compared for each day (horizontal 464 comparison lines trial 1 vs trial 7; *p<0.05, **p<0.01, ***p<0.001 paired t-test). Latencies to fall in the trial 7 465 of a day and trial 1 of the next day were compared (horizontal comparison lines trial 7 vs trial 1; *p<0.05, 466 ***p<0.001 paired t-test). Comparisons between groups for trials 1 and trials 7 were also represented 467 (vertical comparison lines, difference with controls trial 1 and trial 7; *p<0.05, **p<0.01, ***p<0.001 t-test). 468 i,m) Dentate-CL and Interposed-VAL pathway inhibition during Maintenance after a proper Consolidation 469 phase of motor learning task showed early impairments (*p<0.05; repeated measure ANOVA Group effect) 470 (*p<0.05, **p<0.01 and ***p<0.001). j,n) Latencies to fall in trial 1 and 7 were compared for each day 471 (horizontal comparison lines difference trial 1 vs next trial 7; *p<0.05, **p<0.01, ***p<0.001 paired t-test). 472 Latencies to fall in the trial 7 of a day and trial 1 of the next day were compared (horizontal comparison lines 473 trial 7 vs next trial 1; *p<0.05, ***p<0.001 paired t-test). Comparisons between groups for trials 1 and trials 7 474 were also represented (vertical comparison lines, difference from control, trial 1 and trial 7; *p<0.05, 475 **p<0.01, ***p<0.001 t-test). Dent. : Dentate cerebellar nucleus, Int.: interposed cerebellar nucleus. Data 476 represents mean ± S.E.M, n indicates the number of mice. 477 478 Learning is inversely correlated to initial performance in control animals 479 The measure of performance reflects the sum of the learning over all past sessions. To gain 480 more insight into the nature of impairments induced by cerebellar inhibitions, we then 481 reanalyzed the data by focusing on the change of performance across days and nights during the 482 learning (Fig 6). The latency to fall of a single trial is a “noisy” measure of the degree of learning; 483 to obtain a more reliable measure of learning, we simplified the data by performing a linear 484 regression on the performance of each day where the slope indicates the learning rate of the 485 day, and the endpoints provide estimates of the initial and final performances on each day (Fig 486 6A). An initial inspection of average change in latency during days (Fig 6B, warm colors, Suppl. 487 Table 22,23) or change overnight (Fig 6B, cold colors, Suppl. Table 22,23) showed various 488 patterns across the groups (all control groups were pooled here). Interestingly, within-day 489 learning was still significant in the Maintenance phase –when performance does not increase 490 anymore– but was counterbalanced by similar overnight loss of performance. In most cases, the 491 amplitudes of within-day learning and overnight loss were not significantly different from the 492 respective control groups, preventing thus to interpret the nature of difference in learning 20

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

493 between groups. We then examined how the performance changes at each phase of learning 494 evolved as a function of another. 495 In the control group, we unexpectedly found an inverse correlation between the initial 496 performance of each day and the amount of learning: animals with strong initial performance on 497 a given day showed weaker improvements than animals with poor initial performances; this was 498 true for all phases of learning (Fig 6C). The relationship between initial performance and within- 499 day learning evolved across the phases (Fig 6C,D Suppl. Table 24). Notably, the intercept of the 500 linear regression line on the x-axis revealed the existence of “fixed points”, which correspond to 501 the initial performance which is not expected to be followed by an increase or decrease of 502 performance following the training of the day. For the control group, such fixed points were 503 found for the 3 phases of learning and exhibited a progressive increase in value from the 504 Acquisition to the Maintenance phases (92s, 145s and 164s for the three phases, see Suppl. 505 Table 24). Thus, in control mice, the amount of learning on each day depends on the initial 506 performance of the day and is a decreasing function of the initial performance up to a fixed 507 point where no learning is expected, which varies with the phases of learning. 508 The learning not only depends on the process taking place during the multiple trials of a 509 day, but also on the ability of the animal to maintain the same level of performance the next day 510 by consolidating the learning. We then examined h ow the increase of initial performance 511 between two successive days was correlated with the gain of performance within the first of 512 these two days. We found that these values were strongly correlated in the control group with a 513 slope close to 1 at all phases of learning (Fig 6G, Suppl. Table 25); interestingly, the correlation 514 remained high in the Maintenance phase even if the average performance of the population did 515 not increase. This indicates that for each mouse, the improvement of performance during the 516 task (or loss in the Maintenance phase) is fully retrieved the next day (Fig 7A top left). 517 In groups with full cerebellar nuclei inhibition, the learning patterns shall result from the 518 combination of multiple output pathways, so we focused our attention on groups with discrete 519 phenotypes: 1) the Dentate-CL group with inhibition by CNO during the task which shows 520 reduced learning (Fig 5B,G) that contrasts with the lack of effect when inhibition by CNO takes 521 place after the task (Fig 5C,H) indicating that the main effect occurs during learning; 2) the 522 Interposed-VAL group with inhibition by CNO after the task which shows an effect on overnight 21

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

523 change in performance (Fig 5F,L). We did not try to further analyze the effect of CNO inhibition 524 during the task of Interposed-VAL mice, since their performance likely result from a compound 525 effect of online inhibition –which resemble the effect found in Dentate-CL mice, but is difficult to 526 interpret because of the reduced performance in the fixed-speed rotarod– , and of offline 527 inhibition because the action of CNO extends beyond the duration of the task.

528 22

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

529 530 Fig 6. Learning profile is altered by the chemogenetic modulation of Dentate-centrolateral thalamus and 531 Interposed-ventral anterior lateral thalamus pathway. a) Example of evolution of latencies to fall for a control 532 mouse during the accelerated rotarod protocol, showing a high variability along session. Linear regressions 533 estimated values of trial 1 and 7 during each day are shown using red hollow dots, within-day learning (red 534 arrows) and overnight change (dotted black arrows). b) Bar plot showing change in latency associated to 535 within-day learning and overnight change for all experimental groups in all phases (mean +/- SEM). Difference 536 from 0: ° p<0.05, °° p<0.01 °°° p<0.001; Tukeys’ difference between phases: * p<0.05, ** p<0.01, *** p<0.001. 537 Difference from control group: c p<0.05. c) Scatterplots showing the within-day learning as a function of the 538 initial performances in pooled control group in all phases. Ordinary least square linear regressions outcomes 539 are shown for each phase. Bivariate plot showing the within-day learning as a function of the initial 540 performance (d,e,f) and day+night learning as a function of within-day learning (g,h,i) for Pooled controls 541 (left), Dent.-CL task+offline (middle) and Int.-VAL offline (right). The ellipse contains 50% of a bivariate normal 542 distribution fitted to the points and the dot indicates the center of the distribution. Deming linear regression 543 outcomes are represented for each phase (d,e,f,g,h,i). The intercepts on the initial performance axis are 544 represented for each phase (D,E,F). ρ indicates the value of the Pearson’s coefficient for each phase. 545 546 547 Differential impact on learning of Dentate-CL and Interposed-VAL neurons inhibition. 548 The inhibition of the Dentate-CL neurons during the task preserved the inverse correlation 549 between the initial performance and the learning intensity in the Acquisition and Consolidation 550 phases, but the intensity of learning was lower as attested by the lower values of the fixed 551 points (59 s and 115 s respectively, Fig 6E , Suppl. Table 24). This indicates that for a given initial 552 performance, the amount of learning was lower in the CNO-treated group than in the 553 corresponding SAL-treated control group (with similar slopes but non-overlapping confidence 554 intervals for intercept of x axis in CNO vs SAL groups in Consolidation phase, Suppl. Table 24). In 555 the Maintenance phase, the initial performances remained at a similar value as in the 556 Consolidation phase, indicating that no further long-lasting learning occurred. In contrast with 557 the control group, the correlation between initial performance and within-day learning was not 558 significant anymore (Fig 6E, Suppl. Table 24). Moreover, while the learning was preserved 559 overnight in the Acquisition and Consolidation phases, the correlation between within-day and 560 between-days learning was lost in the Maintenance phase (Fig 6H, Suppl. Table 25). Thus, the 561 inhibition of the Dentate-CL reduced the intensity of learning and prevented the consolidation in 562 the late stage of learning (Fig 7A, top right). 563 The inhibition of the Interposed-VAL neurons after the task in the Acquisition and 564 Consolidation phases was associated with a striking pattern of increased intensity of within-day 565 learning and increased overnight forgetting in the Consolidation and Maintenance phases (Fig 566 6B, Suppl. Table 22). However, this came with a de-correlation of initial performance of each day

23

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

567 and learning intensity (Fig 6F, Suppl. Table 24), except in the Acquisition phase, when the CNO 568 has not yet been administered. Despite the large within-day learning, the initial performance of 569 each day only marginally improved consistent with a limited consolidation. Indeed, the amount 570 of learning over each day + following night was not correlated anymore with the learning during 571 the day (Fig 6I, Suppl. Table 25) in the Consolidation and in the Maintenance phase. 572 Interestingly, the increase in learning and reduced day+night learning observed in the 573 Consolidation persisted in the Maintenance phase, once the CNO has been replaced by SAL, 574 indicating that the deficit in learning could not be recovered afterwards (Fig 5F,L). These results 575 suggest that despite a poor consolidation, the mice increased their performance on each day by 576 regaining the learning from previous days and adding some more learning on top, a feature 577 distinctive of the existence of savings (Fig 7A bottom left, see Discussion). 578 Overall these results indicate that the Dentate-CL and Interposed-VAL cerebellar neurons are 579 involved in different functions and that the cerebellum controls multiple processes during and 580 after learning sessions of accelerated rotarod.

24

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

581 582 583 Fig 7. Summary of the behavioral findings. a) Schematic representation of the within-day learning (plain 584 arrow) and overnight loss (dotted arrow) for Pooled controls, Dentate-centrolateral thalamus+CNO 585 administered during the task (Dent.-CL+CNO) and Interposed-ventral anterior lateral thalamus+CNO 586 administred after the task(Int.-VAL+CNO). In the former group, the intensity of learning is reduced but 587 consolidation is intact, while in the latter, the consolidation is reduced but a latent trace remains in the form 588 of savings that are promptly relearned the next day. b) Summary of the controls exerted by Dentate-CL and 589 Interposed-VAL neurons on the skill learning: the Dentate-CL neurons contribute to online-learning and 590 retrieval from learned skill, while the Interposed-VAL neurons contribute to consolidate offline the recent 591 learning into a form of consolidated, readily-available, memory; in the absence of such consolidation, a trace 592 of learning still remains in the form of savings, which will accelerate the online learning the next day; a defect 593 in consolidation in the earlier phases cannot be rescued in the late phase. 594 595

25

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

596 DISCUSSION 597 In this study, we observed task-specific discharge in the cerebellar nuclei consistent with 598 an involvement of the intermediate and lateral parts of the cerebellum. We also found that the 599 transient, mild chemogenetic modulation which reduces but does not suppress the cerebellar 600 nuclei activity, preserves the motor abilities but disrupts motor learning. Moreover, we could 601 distinguish two contributions of the cerebellum to learning; one is carried by neurons projecting 602 toward the midline thalamus and possibly to the motor thalamus, and is needed for learning 603 and recall. The other is carried by neurons projecting toward the motor thalamus and is required 604 to perform an offline consolidation of a latent memory trace into a consolidated, readily 605 available, motor skill (Fig 7). Thus, our results suggest that learning a complex motor task 606 involves the coupling of the cerebellum with the basal ganglia, which is needed during the 607 learning and execution, and the coupling of the cerebellum with the cortex which plays a critical 608 role in the offline consolidation of the task. 609 610 A role for the cerebellum in a multi-nodal network 611 The accelerating rotarod task has proven to be a powerful motor learning paradigm to 612 study the multiple scales of the mechanisms of motor learning in the rodents ((e.g. Costa et al., 613 2004; Rothwell et al., 2014; Yang et al., 2009)). It is also suitable to study the multiple time scales 614 of motor learning: when repeated over multiple days, distinct phases of learning, with different 615 rate of performance improvement and organization of locomotion strategies, can be 616 distinguished (Buitrago et al., 2004) and selectively disrupted (Hirata et al., 2016). The basal 617 ganglia and motor cortex are recruited and required to complete the task (Cao et al., 2015; Costa 618 et al., 2004; Kida et al., 2016). Moreover, the areas involved in the basal ganglia evolves along 619 the phases of learning (Cao et al., 2015; Durieux et al., 2012; Yin et al., 2009). Consistent with 620 the growing evidence of the dynamical interplay between the cerebellum, cortex and basal 621 ganglia (Caligiore et al., 2017), the involvement of the connections between the cerebellum and 622 the other structures in motor learning is thus expected. 623 The impact of cerebellar defects or manipulations on rotarod learning have been 624 examined in too many studies to be listed exhaustively here, but the reported effects range from 625 ataxia and disruption of the ability to run on a rod (Sausbier et al., 2004), to normal learning 26

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

626 (Galliano et al., 2013), defects in learning (Groszer et al., 2008), defects in consolidation (Sano et 627 al., 2018) and even increase in learning (Iscru et al., 2009). However, the cerebellum is critical for 628 interlimb coordination (Machado et al., 2015), and many studies lack proper motor controls to 629 test the ability to walk on a rotating rod: the improvement in performance may be limited by 630 problems of running on the rod rather than problems of learning. Moreover studies often 631 involve genetic mutations which leave room from variable compensations along development 632 and adult life, as exemplified by the diversity of the motor phenotype of mice with degeneration 633 of (Porras-Garcia et al., 2013). Finally the multiphasic nature of rotarod learning is 634 often overlooked. Yet, the targeted suppression of parallel-fiber to Purkinje cell synaptic long- 635 term depression in the cerebellar cortex disrupts rotarod learning after the acquisition phase 636 without altering any other motor ability (Galliano et al., 2013). Consistently, Thyrotropin- 637 releasing hormone (TRH) knock-out mice do not express long-term depression at parallel fiber- 638 Purkinje cell synapses and exhibit impaired performance in late phase of the rotarod learning, 639 while the administration of TRH in the knock-out mice both restores long-term depression and 640 accelerated rotarod learning(Watanave et al., 2018). More generally, studies in mutant mice 641 suggest that cerebellar plasticity is required for adapting skilled locomotion(Vinueza Veloz et al., 642 2015). This raises the possibility that cerebellar plasticity is involved in accelerated rotarod 643 learning. 644 645 A specific impact on learning of DN-CL neurons. 646 In our study, we found that the chemo-genetic inhibition of Dentate-CL neurons during the 647 task reduces the performances of the mice in the late phases of learning. This effect unlikely 648 results from basic motor deficits: we found that the chemogenetic modulation did not 649 significantly alter 1) limb motor coordination in footprint analysis, 2) strength in the grid test, 3) 650 speed in spontaneous locomotion in the open-field test, 4) locomotion speed and balance 651 required to complete the horizontal bar test and 5) body-limb coordination and balance 652 required in the vertical pole test. Since all these motor parameters may be affected by cerebellar 653 lesions, this suggests that Dentate-CL neurons are not necessary to maintain those functions, 654 which might thus be relayed by other cerebellar nuclei neurons; indeed focal lesions in the 655 intermediate cerebellum (thus projecting to the Interposed nuclei) has been reported to induce

27

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

656 ataxia without altering rotarod learning (Stroobants et al., 2013). Alternatively the effect of the 657 partial inhibition induced by CNO (typically ~50-80% reduction in firing rate) may be 658 compensated at other levels in the motor system to ensure normal performances in these tasks; 659 indeed, the selective ablation of Dentate-CL neurons has been reported to yield locomotor 660 deficits in the initial performances on the accelerated rotarod (Sakayori et al., 2019), which 661 contrasts with the lack of significant deficit in the Acquisition phase following Dentate-CL 662 (partial) inhibition in our study. A possible explanation could be that our intervention selectively 663 disrupted the advanced patterns of locomotion only needed at the higher speeds of the rotarod. 664 However, the highest speeds reached on the rotarod correspond to the average locomotion 665 speed in the open-field. Moreover, in our conditions, the inhibition of the Dentate-CL neurons 666 did not produce significant deficits in the fixed speed rotarod; CNO-treated animals ran in 667 average for about two minutes at 20 r.p.m. while they fell in average after the same amount of 668 time on the accelerating rotarod, corresponding to a rotarod speed below 20 r.p.m. at the time 669 of the fall. This rules out a contribution of weakness or fatigue to the latency to fall in the 670 accelerated rotarod, the CNO-treated animals being able to run on the fixed speed rotarod more 671 than twice the distance, at a higher speed, than the distance they run on the accelerated 672 rotarod before falling. Overall, this indicates that the inhibition of the Dentate-CL neurons 673 selectively affects accelerated-rotarod learning and not the elementary motor abilities needed in 674 the task. 675 We observed in control mice and in mice with inhibition of Dentate-CL neurons that the 676 daily gain in latency depended from the initial latency of the day; for a given initial performance, 677 CNO-treated animals in the Consolidation phase showed lower increase in latency to fall 678 indicating weaker learning rate. This effect did not reach significance in the Acquisition phase, 679 possibly because the DN-CL network is more engaged in later phases of learning, as suggested 680 by our electrophysiological experiments. Moreover, the administration of CNO in animals that 681 learned under Saline SAL induced a sudden drop in performance, revealing a deficit in the 682 execution of the learned task. This indicates that the Dentate-CL neurons both contribute to 683 learning and retrieval of motor skills. 684 The inhibition of Interposed-VAL neurons during the task also yield lower levels of 685 performance in the Maintenance stage, suggesting that these neurons contribute also to 28

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

686 learning and retrieval of motor skills, although the mild defect in fixed speed rotarod could 687 indicate the presence of an execution problem. Interestingly, both Dentate and Interposed 688 nuclei contain some neurons with collaterals in both thalamic structures(Aumann and Horne, 689 1996; Sakayori et al., 2019), suggesting that the effect on learning could be mediated in part by a 690 combined action on the learning process of the striatum (via the CL thalamus) and cortex (via 691 the VAL thalamus); however, consistent with (Sakayori et al., 2019), the manipulations of 692 cerebellar neurons retrogradely targeted either from the CL or from the VAL produce different 693 effects. 694 695 Contribution of VAL-projecting cerebellar neurons to offline consolidation. 696 While in the control mice, the final performance at the end of a session could be 697 reproduced at the beginning of the next session, this link was lost when Interposed-VAL neurons 698 were inhibited after the task (‘offline’), suggesting an impairment of consolidation. However, in 699 this group of mice, the daily gain of performance increased across days instead of decreasing; 700 this larger gain of performance reveals the presence of savings. Therefore, if the inhibition of 701 Interposed-VAL neurons alters the consolidation, a latent trace of the learning remains and 702 allows for a faster relearning on the next day. 703 The effect of CNO peaks in less than an hour and lasts for several hours afterwards 704 (Alexander et al., 2009); therefore consolidation is substantially disrupted by altering the 705 cerebellar activity in the few hours that follow the learning session. We found evidence for 706 reduced tonic activity during rest in the open-field immediately after the learning sessions in the 707 Interposed nucleus in all phases of learning ; this was also observed during the rest between 708 trials in a smaller subset of animals (in which the other cerebellar and thalamic nuclei also 709 exhibited significant changes in firing rate). These results suggest that the consolidation starts 710 readily when the animals are resting. This falls in line with a number of evidence indicating that 711 cerebellar-dependent learning is consolidated by the passage of time, even in the awake state 712 (Cohen et al., 2005; Doyon et al., 2009; Muellbacher et al., 2002; Nagai et al., 2017; Shadmehr 713 and Holcomb, 1997), although very few studies in the Human have performed offline cerebellar 714 stimulations (Samaei et al., 2017).

29

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

715 In the case of rotarod, it has been noted that sleep is not required for the overnight 716 preservation of performance (Nagai et al., 2017); however sleep may still be required for the 717 change of cortical (Cao et al., 2015; Li et al., 2017) or striatal neuronal substrate of the 718 accelerated rotarod skill (Yin et al., 2009). Therefore, while the offline activity of cerebellar 719 nuclei could be more specifically associated in converting savings into readily available skills, 720 multiple processes of memory consolidation would co-exist and operate at different timescales. 721 The existence of multiple timescales for consolidation has already been described in 722 Human physiology where the movements could be consolidated without sleep while 723 consolidation of goals (Cohen et al., 2005) or sequences (Doyon et al., 2009) would require 724 sleep. It is indeed difficult, as for most real-life skills, to classify the accelerating rotarod as a 725 pure adaptive learning or sequence learning: on one hand, the shape of the rod and its rotation 726 induce a change in the correspondence between steps and subsequent body posture and thus 727 require some locomotor adaptation. On the other hand, the acceleration of the rod introduces 728 sequential aspects: 1) the same step executed a few seconds or tens of seconds apart in a trial 729 result in different consequences on body position, and 2) asymptotic performances involve to 730 use successively multiple types of gait as the trials progress (Buitrago et al., 2004). Following 731 offline inhibition of Interposed-VAL neurons, which aspect would be maintained and which 732 would be lost? Faster relearning has been proposed to reflect an improved performance at 733 selecting successful strategies (Morehead et al., 2015; Ruitenberg et al., 2018). An attractive 734 possibility could be that novel sensory-motor correspondences encountered on the rotarod 735 would remained learned, possibly leaving a memory trace within the cerebellum, but these 736 elementary ‘strategies’ would not be properly temporally ordered into a sequence over the 737 duration of a trial (2 minutes); the next day learning session would benefit from the existence of 738 these fragments of skill, but learning would still be required to order them properly. A similar 739 idea has indeed been proposed for the contribution of cerebellum to sequence learning 740 (Spencer and Ivry, 2009). Alternatively, recent studies revealed mechanisms which could serve 741 sequence learning (Khilkevich et al., 2018; Ohmae and Medina, 2015) and would be affected by 742 the offline inhibition of Interposed-VAL neurons via their feedback collaterals to the cerebellar 743 cortex (Houck and Person, 2015). However, our study does not allow us to conclude on the 744 nature of savings remaining after the offline inhibition of Interposed-VAL neurons. 30

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

745 746 An internal model in the output of the cerebellum? 747 Improving the rotarod performance requires the mice to match their locomotion speed to 748 the accelerating speed of the rod. In the cerebellar nuclei, we observed that neurons exhibited 749 substantial modulation as a function of speed on the rotarod, with a negative slope in the 750 majority of cases, but this modulation was very different from the modulation of speed in the 751 open-field, which rather exhibited a positive slope if any. This suggests that the purpose of the 752 cerebellar activity in the rotarod is not to specify the locomotion speed. In mice, the speed of 753 locomotion and gait are transmitted to the motoneurons by neurons in the locomotor 754 region, which receives sparse inputs from the forebrain except from the basal ganglia, and little 755 if any inputs from the cerebellar nuclei (Caggiano et al., 2018). Therefore, in the accelerated 756 rotarod task, the profile of locomotion speed determined in a cerebello-cortico-basal ganglia 757 network is more likely propagated to the midbrain locomotor region via the output nuclei of the 758 basal ganglia (Rueda-Orozco and Robbe, 2015). 759 It is well established that a number of neurons in the Interposed nucleus exhibit a 760 modulation along the stride (Armstrong and Edgley, 1984; Sarnaik and Raman, 2018), and a 761 transient optogenetic manipulation of this activity results in an alteration of the gait (Sarnaik 762 and Raman, 2018). Indeed, the optogenetic activation of genetically-defined Interposed nucleus 763 neurons projecting mainly to the ventral lateral thalamus and the produced higher 764 strides (Low et al., 2018). Stride-related signals are also found in the Dentate nucleus, which is 765 more clearly recruited in response to perturbations (Schwartz et al., 1987) or during skilled 766 locomotion such as on a ladder (Marple-Horvat and Criado, 1999); this is consistent with the 767 selective impact of lateral cerebellum lesion to obstacle stepping while preserving overground 768 locomotion (Aoki et al., 2013). The changes of firing rate in the Dentate and Interposed units 769 reported in our study thus likely reflect the integrated representation of the stride (Sauerbrei et 770 al., 2015). An alternate –and compatible- hypothesis, is that the cerebellar output reflects the 771 expected consequences of actions, representing an internal model of locomotion on the rod. In 772 the rotarod, the faster the rod turns, the less individual steps produce the expected forward 773 progress; the decrease of nuclear cell firing as the speed increase could indeed reflect the 774 estimated “undershooting” of the outcome of the steps. 31

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

775 776 In conclusion, our results provide clear evidence for the existence of online contributions 777 of the cerebello-thalamic pathways to the formation and retrieval of motor memories 778 distributed in a cerebello-striato-cortical network. They also show a contribution to the offline 779 consolidation of savings by a distinct pathway. Thus, our work highlights the importance of 780 studying the contribution to learning of single nodes in the motor network from an integrated 781 perspective (Caligiore et al., 2017; Krakauer et al., 2019). 782 783 AUTHOR CONTRIBUTIONS 784 Conceptualization, A.P.V., D.P. and C.L.; Methodology, A.P.V. and D.P.; Software, R.W.S. and C.L.; 785 Formal Analysis, A.P.V., R.W.S. and C.L.; Investigation, A.P.V., R.W.S., C.M.H and J.L.F.; Data 786 Curation, A.P.V., R.W.S. and C.L.; Writing –Original Draft, A.P.V., C.L. and D.P.; Writing –Review & 787 Editing, D.P. and C.L.; Visualization, A.P.V., R.W.S., J.L.F and C.L.; Supervision, D.P. and C.L.; 788 Funding Acquisition, D.P. and C.L. 789 790 ACKNOWLEDGMENTS 791 This work was supported by Fondation pour la Recherche Medicale (FRM, DPP20151033983) to 792 D.P. and Agence Nationale de Recherche to D.P. (ANR-16-CE37-0003-02 Amedyst, ANR-19-CE37- 793 0007-01 Multimod, Labex Memolife) and to C.L. (ANR-17-CE37-0009 Mopla, ANR-17-CE16-0019 794 Synpredict) and by the Institut National de la Santé et de la Recherche Médicale (France). 795 796 MAIN FIGURE TITLES AND LEGENDS 797 798 Figure 1. The engagement of the cerebellar nuclei during the accelerating rotarod task 799 changes along the learning protocol. a) Scheme of the accelerating rotarod protocol (top) and 800 latency to fall (bottom) during the accelerating rotarod across trials along days in mice recorded 801 over the 7 days (**p<0.01, *p<0.05 t-test comparing trial 1 vs trial 7). b) Example of electrode 802 placements in the Dentate (top) and Interposed nuclei (bottom). c) Firing rate histograms 803 showing the evolution of activity in representative cells for Dentate (left) and Interposed (right) 804 nuclei during Acquisition (top), Consolidation (middle) and Maintenance (bottom). Rotarod 805 trials are shown in colors, while the open-Field before (OF1) and after (OF2) the rotarod session

32

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

806 are shown in grey, and the intertrial periods are shown in black. d) Raster plots showing the 807 activity of representative cells for dentate (left) and interposed (right) during trials for 808 acquisition, consolidation and maintenance. e) Distribution of numbers of trials in which cells 809 showed a significiant modulation of firing rate between trial and inter-trial periods during 810 acquisition (Acq), consolidation (Cons) and maintenance (Maint) phase for Dentate (top) and 811 Interposed (bottom) (*p<0.05, ***p<0.001 Mann-Whitney test, Holm-Sidak corrected for 812 multiple comparison). f) Evolution of the average firing rate (mean +/- SEM), normalized by 813 subtracting the average firing rate during the active part of the open-field session before the 814 first rotarod trial, during Acquisition, Consolidation and Maintenance for Dentate (top left), 815 Interposed (bottom left) nuclei, centrolateral thalamus (CL, top right), ventrolateral thalamus 816 (VAL, bottom right) (*p<0.05, **p<0.01, ***p<0.001 Dunett Posthoc test). 817 818 819 Fig 2. Dentate and Interposed nuclei display a context-dependent sensitivity to speed, with a 820 negative correlation between cerebellar activity and rotarod speed. a) Scatterplot showing the 821 average firing rate per speed bins during the different trials of rotarod (left) versus open-field 822 (right) session, for Dentate (top) and Interposed (bottom) nucleus (linear regression lines are 823 shown with a 95% confidence interval) (*p<0.05, ***p<0.001 Pearson correlation). b) Scatter 824 plot showing the slope of linear regression explaining the firing rate by the speed, in the open- 825 field versus on the rotarod for each neuron in Dentate (top) and Interposed (bottom) nucleus 826 during Acquisition (Acq.), Consolidation (Cons.) and Maintenance (Maint.). The diagonal dotted 827 line represents equality between the slope in the open-Field and rotarod. Marginal axes show 828 the histograms of the distributions of slopes, smoothed using a Gaussian kernel density 829 estimate (ς=0.05). c) Boxplots displaying the distributions of linear regression slopes of neurons 830 from Dentate (top) and Interposed (bottom) during open-field (OF) and rotarod (Rtrd) 831 (***p<0.001 Wilcoxon test). d) Significant linear regression slopes for rotarod (mean +/- SEM) 832 for Dentate (top left), centrolateral thalamus (CL, top right), Interposed (bottom left) and 833 ventral anterior lateral thalamus (VAL, bottom right) during Acquisition (A), Consolidation (C) 834 and Maintenance (M) (*p<0.05, **p<0.01, ***p<0.001 Tukey Posthoc test). e) Scatter plot 835 showing the correspondence between the slope of linear regression on rotarod versus the 33

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

836 associated Pearson correlation coefficient for each neuron for Dentate (top) and Interposed 837 (bottom) nucleus during Acquisition (Acq.), Consolidation (Cons.) and Maintenance (Maint.). 838 The lines represent the isotonic regression of the Pearson’s r by the slope on rotarod 839 (***p<0.001 Spearman Rank test). f) Density of Pearson’s r coefficients for Dentate and CL, (top) 840 and Interposed and VAL, (bottom) for Acquisition, Consolidation and Maintenance phase, the 841 overlapping area is represented in blue and is associated to the overlapping index η (*p<0.05, 842 **p<0.01, ***p<0.001 Mann Whitney test); a barplot of the value of η is shown on the right. 843 844 Fig 3. The quiet state activity in the cerebellar nuclei and in the thalamus changes during the 845 learning protocol. a) Boxplots showing the distribution of change in mean firing rate between 846 open-field (OF) before and after rotarod, during acquisition, consolidation and maintenance for 847 Dentate, Interposed, centrolateral thalamus (CL) and ventral anterior lateral thalamus (VAL) 848 (***p<0.001 Tukey Posthoc test). b) Evolution of the mean quiet state firing rate (mean +/- 849 SEM) in open-field sessions and Inter-trial periods during Acquisition, Consolidation and 850 Maintenance for Dentate (top left), Interposed (top right), CL (bottom left) and VAL (bottom 851 right). Mean firing rate was normalized by subtracting the open-field before (OF1) (*p<0.05, 852 **p<0.01, ***p<0.001 Dunett Posthoc test). 853 854 Fig 4. Cerebellar inactivation impairs the performance in the Consolidation and Maintenance 855 but not Acquisition phases of motor learning task. a) Scheme showing electrode position and 856 viruses injections. b) Coronal section of the cerebellum showing hM4Di-DREADD expression in 857 the three cerebellar nuclei. c) Representative confocal image showing hM4Di-DREADD 858 expression on neuronal membranes. d) Image showing DAPI positive neurons expressing 859 hM4Di-DREADD. e) Coronal section of the cerebellum showing electrode placement (yellow 860 dotted line)(left) and cells expressing hM4Di-DREADD nearby the electrode (right). f) Examples 861 of spike shapes obtained from spike sorting in cerebellar nuclei (mean +/- SD). g) Examples of 862 high-pass filtered traces of cerebellar nuclei recordings before and after CNO injection. h) 863 Boxplots showing the mean firing rate of neurons recorded in DREADD and non-DREADD 864 injected mice after CNO or SAL injection during Acquisition (Acq.), Consolidation (Cons.) and 865 Maintenance (Maint). Cerebellar nuclei firing rate was reduced after 1 mg/kg of CNO injection 34

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

866 in DREADD injected mice (***p<0.001 paired t-test). Small variations were observed in 867 Sham+SAL group (*p<0.05, ***p<0.001 paired t-test). i) Daily of injections of CNO before trial 1 868 in DREADD injected mice induced impairment in motor learning during Consolidation and 869 Maintenance but not Acquisition phase (*p<0.05, ***p<0.001; repeated measure ANOVA 870 Group effect). j) CNO injections after Acquisition phase (30 min after trial 7) induced 871 impairment in motor learning during Consolidation phase (*p<0.05, **p<0.01; repeated 872 measure ANOVA Group effect). k,l) Latencies to fall in trial 1 and 7 were compared for each day 873 (horizontal comparison lines trial 1 vs trial 7; *p<0.05, **p<0.01, ***p<0.001 paired t-test). 874 Latencies to fall in the trial 7 of a day and trial 1 of the next day were compared (horizontal 875 comparison lines trial 7 vs trial 1; *p<0.05, ***p<0.001 paired t-test). Comparisons between 876 groups for trials 1 and trials 7 were also represented (vertical comparison lines, respectively 877 trial 1 difference to controls and trial 7 difference to controls; *p<0.05, **p<0.01, ***p<0.001 t- 878 test). Data represents mean ± S.E.M. n indicates the number of mice. 879 880 Fig 5. Dentate-centrolateral thalamus and Interposed-ventral anterior lateral thalamus 881 pathway inactivation impairs consolidation and maintenance but not acquisition phase of 882 motor learning task. Schemes showing AVV5-hSyn-DIO-hM4D(Gi)-mCherry and CAV2-Cre-GFP 883 virus injections in Dentate and centrolateral thalamus (CL) (a) or Interposed and ventral 884 anterior lateral thalamus (VAL) (d), respectively. Example of Dentate-CL (a, bottom) and 885 Interposed-VAL (d, bottom) neurons expressing GFP (left, 500µm) and m-Cherry (right, 50 µm). 886 b) Daily of injections of CNO before trial 1 in Dentate-CL pathway injected mice induced global 887 impairments during Maintenance but not Acquisition and Consolidation phase (*p<0.05, 888 **p<0.01, ***p<0.001; repeated measure ANOVA Group effect). c) CNO injections after 889 Acquisition phase (30 minutes after trial 7) in Dentate-CL pathway did not induce alterations 890 during Consolidation or Maintenance phase (#p<0.05, ##p<0.01, ###p<0.001; repeated 891 measure ANOVA Group effect). e) Daily injections of CNO before trial 1 in Interposed-VAL 892 pathway injected mice induced early impairments during Consolidation and Maintenance but 893 not Acquisition phase (*p<0.05, **p<0.01, ***p<0.001; repeated measure ANOVA Group 894 effect). f) CNO injections after Acquisition phase (30 minutes after trial 7) in Interposed-VAL 895 pathway induced early impairments during Consolidation and Maintenance phase (#p<0.05, 35

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

896 ##p<0.01, ###p<0.001; repeated measure ANOVA Group effect). g,h,k,l) Latencies to fall in trial 897 1 and 7 were compared for each day (horizontal comparison lines trial 1 vs trial 7; *p<0.05, 898 **p<0.01, ***p<0.001 paired t-test). Latencies to fall in the trial 7 of a day and trial 1 of the 899 next day were compared (horizontal comparison lines trial 7 vs trial 1; *p<0.05, ***p<0.001 900 paired t-test). Comparisons between groups for trials 1 and trials 7 were also represented 901 (vertical comparison lines, difference with controls trial 1 and trial 7; *p<0.05, **p<0.01, 902 ***p<0.001 t-test). i,m) Dentate-CL and Interposed-VAL pathway inhibition during Maintenance 903 after a proper Consolidation phase of motor learning task showed early impairments (*p<0.05; 904 repeated measure ANOVA Group effect) (*p<0.05, **p<0.01 and ***p<0.001). j,n) Latencies to 905 fall in trial 1 and 7 were compared for each day (horizontal comparison lines difference trial 1 vs 906 next trial 7; *p<0.05, **p<0.01, ***p<0.001 paired t-test). Latencies to fall in the trial 7 of a day 907 and trial 1 of the next day were compared (horizontal comparison lines trial 7 vs next trial 1; 908 *p<0.05, ***p<0.001 paired t-test). Comparisons between groups for trials 1 and trials 7 were 909 also represented (vertical comparison lines, difference from control, trial 1 and trial 7; *p<0.05, 910 **p<0.01, ***p<0.001 t-test). Dent. : Dentate cerebellar nucleus, Int.: interposed cerebellar 911 nucleus. Data represents mean ± S.E.M, n indicates the number of mice. 912 913 Fig 6. Learning profile is altered by the chemogenetic modulation of Dentate-centrolateral 914 thalamus and Interposed-ventral anterior lateral thalamus pathway. a) Example of evolution 915 of latencies to fall for a control mouse during the accelerated rotarod protocol, showing a high 916 variability along session. Linear regressions estimated values of trial 1 and 7 during each day are 917 shown using red hollow dots, within-day learning (red arrows) and overnight change (dotted 918 black arrows). b) Bar plot showing change in latency associated to within-day learning and 919 overnight change for all experimental groups in all phases (mean +/- SEM). Difference from 0: ° 920 p<0.05, °° p<0.01 °°° p<0.001; Tukeys’ difference between phases: * p<0.05, ** p<0.01, *** 921 p<0.001. Difference from control group: c p<0.05. c) Scatterplots showing the within-day 922 learning as a function of the initial performances in pooled control group in all phases. Ordinary 923 least square linear regressions outcomes are shown for each phase. Bivariate plot showing the 924 within-day learning as a function of the initial performance (d,e,f) and day+night learning as a 925 function of within-day learning (g,h,i) for Pooled controls (left), Dent.-CL task+offline (middle) 36

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

926 and Int.-VAL offline (right). The ellipse contains 50% of a bivariate normal distribution fitted to 927 the points and the dot indicates the center of the distribution. Deming linear regression 928 outcomes are represented for each phase (d,e,f,g,h,i). The intercepts on the initial performance 929 axis are represented for each phase (D,E,F). ρ indicates the value of the Pearson’s coefficient for 930 each phase. 931 932 Fig 7. Summary of the behavioral findings. a) Schematic representation of the within-day 933 learning (plain arrow) and overnight loss (dotted arrow) for Pooled controls, Dentate- 934 centrolateral thalamus+CNO administered during the task (Dent.-CL+CNO) and Interposed- 935 ventral anterior lateral thalamus+CNO administred after the task(Int.-VAL+CNO). In the former 936 group, the intensity of learning is reduced but consolidation is intact, while in the latter, the 937 consolidation is reduced but a latent trace remains in the form of savings that are promptly 938 relearned the next day. b) Summary of the controls exerted by Dentate-CL and Interposed-VAL 939 neurons on the skill learning: the Dentate-CL neurons contribute to online-learning and retrieval 940 from learned skill, while the Interposed-VAL neurons contribute to consolidate offline the 941 recent learning into a form of consolidated, readily-available, memory; in the absence of such 942 consolidation, a trace of learning still remains in the form of savings, which will accelerate the 943 online learning the next day; a defect in consolidation in the earlier phases cannot be rescued in 944 the late phase. 945

37

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

946 STAR METHODS 947 RESOURCE AVAILABILITY 948 Lead contact 949 Further information and requests for resources should be directed to and will be fulfilled by the 950 Lead Contact, Daniela Popa ([email protected]). 951 952 Material availability 953 This study did not generate new unique reagents. 954 955 Data and Code availability 956 The data and source code generated during this study will be made available by the 957 corresponding author upon a reasonable request. 958 959 960 EXPERIMENTAL MODEL AND SUBJECT DETAILS 961 Adult male C57BL/6J mice (Charles River, France, IMSR Cat# JAX:000664, 962 RRID:IMSR_JAX:000664), 8 weeks of age and 24 ± 0.4 g of weight at the beginning of the 963 experiment were used in the study. Mice were fed with a chow diet and housed in a 22 °C 964 animal facility with a 12-hr light/dark cycle (light phase 7am–7pm). The animals had free access 965 to food and water. All animal procedures were performed in accordance with the 966 recommendations contained in the European Community Council Directives. 967 968 969 METHOD DETAILS 970 1. Behavioral experiments 971 1.1. Accelerated rotarod task 972 The rotarod apparatus (mouse rotarod, Ugo Basile) consisted of a plastic roller with small 973 grooves running along its turning axis (Bearzatto et al., 2005). One week after injections, mice 974 were trained with seven trials per day during seven consecutive days. This training protocol was 975 chosen in order to distinguish the three different phases (Acquisition, Consolidation and 38

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

976 Maintenance). During each trial, animals were placed on the rod rotating at a constant speed (4 977 r.p.m.), then the rod started to accelerate continuously from 4 to 40 r.p.m. over 300 s. The 978 latency to fall off the rotarod was recorded. Animals that stayed on the rod for 300 s were 979 removed from the rotarod and recorded as 300 s. Mice that clung to the rod for two complete 980 revolutions were removed from the rod and time was recorded. Between each trial, mice were 981 placed in their home cage for a 5-minutes interval. 982 983 1.2. Open-field activity 984 Mice were placed in a circle arena made of plexiglas with 38 cm diameter and 15 cm height 985 (Noldus, Netherlands) and video recorded from above. Each mouse was placed in the open-field 986 for a period of 10 minutes before and after the accelerated rotarod task with the experimenter 987 out of its view. The position of center of gravity of mice was tracked using an algorithm 988 programmed in Python 3.5 and the OpenCV 4 library. Each frame obtained from the open fields 989 videos were analyzed according to the following process: open-field area was selected and 990 extracted in order to be transformed into a grayscale image. Then, a binary threshold was 991 applied on this grayscale image to differentiate the mouse from the white background. To 992 reduce the noise induced by the recording cable or by particles potentially present in the Open- 993 field, a bilateral filter and a Gaussian blur were sequentially applied, since those components are 994 supposed to have a higher spatial frequency compared to the mouse. Finally, the OpenCV 995 implementation of Canny algorithm was applied to detect the contours of the mouse, the 996 position of the mouse was computed as mouse’s center of mass. The trajectory of the center of 997 mass were interpolated in x and y using scipy’s Univariate Spline function (with smoothing factor 998 s=0.2 x length of the data), allowing the extraction of a smoothed trajectory of the mouse. The 999 distance traveled by the mouse between two consecutive frames was calculated as the variation 1000 of position of the mouse multiplied by a scale factor, to allow the conversion from pixel unit to 1001 centimeters. The total distance traveled was obtained by summing the previously calculated 1002 distances over the course of the entire open-field session. The speed was computed as the 1003 variation of position of center points on two consecutive frames divided by the time between 1004 these frames (the inverse of the number of frames per seconds). This speed was then averaged 1005 by creating sliding windows of 1 second. After each session, fecal boli were removed and the 39

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1006 floor was wiped clean with a damp cloth and dried after the passing of each mouse. Active and 1007 quiet state was determined by using a bi-threshold method in which two consecutive 1008 thresholded and filtered frames were subtracted one from each other. In order to have a proper 1009 recognition of both active and quiet state, lower and upper threshold of the changed pixels were 1010 arbitrarily set to 0.04% and 0.12%, respectively. The both thresholds were based on the video 1011 acquisition conditions (camera resolution, focal distance of the objective and distance from the 1012 objective) and the size of the animal detected. For each time point, every percentage of changed 1013 pixels below the lower or above the upper threshold was considered as quiet or active state, and 1014 percentages placed between both thresholds were considered as a continuity of the previous 1015 state. 1016 1017 1.3. Horizontal bar 1018 Motor coordination and balance were estimated with the balance beam test which consists of a 1019 linear horizontal bar extended between two supports (length: 90 cm, diameter: 1.5 cm, height: 1020 40 cm from a padded surface). The mouse is placed in one of the sides of the bar and released 1021 when all four paws gripped it. The mouse must cross the bar from one side to other and 1022 latencies before falling are measured in a single trial session with a 3-minutes cut-off period. 1023 1024 1.4. Vertical pole 1025 Motor coordination was estimated with the vertical pole test. The vertical pole (51 cm in length 1026 and 1.5 cm in diameter) was wrapped with white masking tape to provide a firm grip. Mice were 1027 placed heads up near the top of the pole and released when all four paws gripped the pole. The 1028 bottom section of the pole was fixated to its home-cage with the bedding present but without 1029 littermates. When placed on the pole, animals naturally tilt downward and climb down the 1030 length of the pole to reach their home cage. The time taken before going down to the home- 1031 cage with all four paws was recorded. A 20 seconds habituation was performed before placing 1032 the mice at the top of the pole. The test was given in a single trial session with a 3-minutes cut- 1033 off period. 1034 1035 1.5. Footprint patterns 40

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1036 Motor coordination was also evaluated by analysing gait patterns. Mouse footprints were used 1037 to estimate foot opening angle and hindbase width, which reflects the extent of muscle 1038 loosening. The mice crossed an illuminated alley, 70 cm in length, 8 cm in width, and 16 cm in 1039 height, before entering a dark box at the end. Their hindpaws were coated with nontoxic water- 1040 soluble ink and the alley floor was covered with sheets of white paper. To obtain clearly visible 1041 footprints, at least 3 trials were conducted. The footprints were then scanned and examined 1042 with the Dvrtk software (Jean-Luc Vonesch, IGBMC). The stride length was measured with 1043 hindbase width formed by the distance between the right and left hindpaws. 1044 1045 1.6. Grid test 1046 The grid test is performed to measure the strength of the animal. It consists of placing the 1047 animal on a grid which tilts from a horizontal position of 0° to 180°. The animal is registered by 1048 the side and the time it drops is measured. The time limit for this experiment is 30 seconds. In 1049 those cases where the mice climbed up to the top of grid, a maximum latency of 30 seconds was 1050 applied. 1051 1052 1.7. Fixed speed rotarod 1053 Motor coordination, postural stability and fatigue were estimated with the rotorod (mouse s view, the mice placed on top of the׳rotarod, Ugo Basile). Facing away from the experimenter 1054 1055 plastic roller were tested at constant speeds (5, 10, 15 and and 20 r.p.m). Latencies before falling 1056 were measured for up to 3 minutes in a single trial session. 1057 1058 2. Chronic in vivo extracellular recordings 1059 Recordings were performed in awake behaving control mice during the open-field as well as the 1060 accelerated rotarod sessions in the day 1, 4 ad 7. Recordings and analysis were performed using 1061 an acquisition system with 32 channels (sampling rate 25kHz; Tucker Davis Technology System 3) 1062 as described in (de Solages et al., 2008; Popa et al., 2013). Cells activity in the Intralaminar 1063 thalamus (centrolateral, CL), motor thalamus (ventral anterior lateral, VAL) and cerebellar nuclei 1064 (Dentate and Interposed) was recorded by using bundles of electrodes consisting in nichrome 1065 wire (0.005 inches diameter, Kanthal RO-800) folded and twisted into six to eight electrode 41

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1066 bundles. To protect these bundles and ensure a good electrodes placement, they were then 1067 placed through metal tubing (8-10mm length, 0.16-0.18mm inner diameter, Coopers Needle 1068 Works Limited, UK) attached to an electrode interface board (EIB-16 or EIB-32; Neuralynx) by 1069 Loctite universal glue. Different configurations were used in order to record simultaneously, CL 1070 (from bregma: AP -1.70 mm, ML ±0.75 mm, DV −3.0 mm), VAL (from bregma: AP -1.4 mm, ML 1071 ±1.0 mm, DV −3.5 mm) and/or the cerebellar nuclei (from Bregma: Interposed: -6.0 AP, +/-1.5 1072 ML, -2.1 depth from dura; Dentate: -6.2 AP, +/-2.3 ML, -2.4 depth from dura). Microwires of each 1073 bundle were connected to the EIB with gold pins (Neuralynx). The entire EIB and its connections 1074 were secured in place by dental cement for protection purpose. Electrodes were cut to the 1075 desired length (extending 0.5mm below tube tip). The impedance of every bundle was 1076 measured and gold-plated electrochemically to lowered and set microwire’s impedance to 200– 1077 500 kΩ. Mice were anesthetized with isoflurane and placed in the stereotaxic apparatus, then 1078 skull and dura were removed above CL, VAL and cerebellar nuclei recording site (see section 2.5. 1079 for a detailed description of the surgical procedure). Electrodes bundles were lowered into the 1080 brain, the ground was placed on the cerebellar cortex and the assembly was secured with dental 1081 cement. One week after the surgery, we started to record cellular activity in CL, VAL and 1082 cerebellar nuclei in freely moving mice placed in the open-field and the accelerated rotarod 1083 sessions. Mice were habituated to the recording cable for 2–3 d before starting the recording. A 1084 custom-made pulley system balanced the weight and torque of the wires during running and 1085 allowed the wires to accompany the mouse during the accelerated rotarod task. Signal was 1086 acquired by headstage and amplifier from TDT (RZ2, RV2, Tucker-Davis Technologies, USA) and 1087 analyzed with Matlab and Python 3.5. The spike sorting was performed with Matlab 1088 (Mathworks, Natick, MA, USA) scripts based on k-means clustering on PCA of the spike 1089 waveforms (Paz et al., 2006). At the end of experiments, the placement of the electrodes (CL, 1090 VAL and cerebellar nuclei) was verified. For each neuron, the spike times were used to compute 1091 the raster plots and firing rate histograms across trial and along phases. Mean firing rate was 1092 calculated for rotarod trials, resting periods and open-field sessions. The algorithm described in 1093 2.2.3 was used to separate from the open-field sessions the active or quiet state in order to 1094 determine their corresponding mean firing rate. In order to assess the effect of the task on the 1095 neurons, a Mann-Whitney test (with α=0.01) was used to compare instantaneous firing rate 42

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1096 during each trial vs the resting period before and after the trial. For each trial, two distributions 1097 were created: one corresponding to the firing rates during the given trial, and the other to the 1098 firing rate during the resting period before and after the trial. In the case of trial 1, only the 1099 resting period after the trial was taken into account, given the fact that trial 1 is preceded by an 1100 open-field session. 1101 1102 2.1. Correlation between neuronal activity and locomotor speed 1103 Linear regression of firing rate (Hz) by locomotor speed (cm/s) was used to compute the 1104 intercept (Hz), slope (Hz*s/cm), R squared and Pearson's correlation coefficient. The slope 1105 reflects the strength of the modulation of the neuronal activity by the locomotor activity and the 1106 Pearson’s correlation coefficient reflects the consistency of this modulation. For each neuron, 1107 linear regressions were computed in two conditions: during trials, and open-Field sessions. For 1108 rotarod trials, speed was estimated as 2πrv/60, where r is the radius of the rotarod axis in 1109 centimeters (1,5cm) and v is the rotation speed in r.p.m., allowing us to estimate the value of 1110 speed in cm/s. Mean firing rate was computed for each speed steps on the rotarod (8s bins). In 1111 order to assess the relationship between neuronal activity and speed while freely moving in the 1112 open-field, the speed was computed as explained in 2.2.3, resulting in a set of speeds expressed 1113 by the mouse during the open-field session. Those speeds were then interpolated in time 1114 considering the previous element in the set, allowing a continuous function that express speed 1115 in cm/s as a function of the time. To allow a valid comparison of the relationship between 1116 neuronal activity and speed on the rotarod and while freely moving in an open-field, the 1117 previously mentioned speed function was then discretized to match equivalent speed step on 1118 the rotarod (speed steps of 1 +/- 0.5 r.p.m.). The mean firing rate associated to each speed step 1119 in the open-field was then calculated as the number of spikes detected while being in the speed 1120 step, divided by the cumulative time spent in this speed step. The linear models were fitted 1121 using the linregress function from the scipy Python package. The modulation of activity by the 1122 speed along phases was compared for the different structures. Pearson’s correlation 1123 distributions for Acquisition, Consolidation and Maintenance phase were compared by using an 1124 overlapping index (non parametric measure of the effect size associated to the difference of the 1125 distributions between the compared structures). A continuous function was calculated from 43

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1126 both distributions and smoothed by using a kernel density estimation (bandwidth = 0.05). Those 1127 functions were normalized in order to have their AUC equal to 1. The overlapping index was 1128 estimated by calculating the integral of the minimum of those functions over the interval of 1129 definition (Pastore and Calcagnì, 2019). This provided the overlapping index, a continuous value 1130 defined between 0 and 1 where the former would mean no overlap between the distributions 1131 and the latter would mean equal distributions. 1132 1133 3. Chemogenetic 1134 3.1. Cerebellar outputs inactivation 1135 We used evolved G-protein-coupled muscarinic receptors (hM4Di) that are selectively activated 1136 by the pharmacologically inert drug Clozapine-N-Oxide (CNO) (Alexander et al., 2009). In our 1137 study, non-cre and cre dependent version of the hM4Di receptor packaged into an AAV were 1138 used in order to facilitate the stereotaxic-based delivery and regionally restricted the expression 1139 of hM4Di. As demonstrated previously (Anaclet et al., 2018; Anaclet et al., 2014; Anaclet et al., 1140 2015; Pedersen et al., 2017; Venner et al., 2016). hM4Di receptor and ligand are biologically 1141 inert in the absence of ligand. Moreover, at the administered dose of 1 mg/kg, CNO injection 1142 induces a maximum effect during the 1–3 h postinjection period (Anaclet et al., 2018; Anaclet et 1143 al., 2014) which enables us to confirm that during the whole duration of our protocols the CNO 1144 was still effective. We are therefore confident that the findings described in our study result 1145 from specific inhibition of the targeted neuronal population and not from a nonspecific effect of 1146 CNO or its metabolite clozapine (Gomez et al., 2017). 1147 In order to globally inactivate the cerebellar outputs, stereotaxic surgeries were used to inject 1148 DREADD viral constructs bilaterally into the Dentate, Interposed and . Mice were 1149 anesthetized with isoflurane for induction (3% in closed chamber during 4-5 minutes) and 1150 placed in the Kopf stereotaxic apparatus (model 942; PHYMEP, Paris, France) with mouse 1151 adapter (926-B, Kopf), and isoflurane vaporizer. Anesthesia was subsequently maintained at 1– 1152 2% isoflurane. A longitudinal skin incision and removal of pericranial connective tissue exposed 1153 the bregma and lambda sutures of the skull. The coordinates for the Dentate nucleus injections 1154 were: 6.2 mm posterior to bregma, +/-2.3 mm lateral to the midline and -2.4 mm from dura 1155 while the Interposed injections were placed anteroposterior (AP) -6.0 mm, mediolateral (ML) = 44

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1156 +/-1.5 mm in respect to bregma and dorsoventral (DV) -2.1 mm depth from dura. Finally, the 1157 Fastigial injections were placed -6.0 AP, +/-0.75 ML in respect to bregma and -2.1 depth from 1158 dura. Small holes were drilled into the skull and DREADD (AAV5-hSyn-hM4D(Gi)-mCherry, 1159 University of North Carolina Viral Core, 7.4 × 1012 vg per ml, 0.2 μl) or control (AAV5-hSyn-EGFP, 1160 UPenn Vector Core, the same concentration and amount) virus were delivered bilaterally via 1161 quartz micropipettes (QF 100-50-7.5 ,Sutter Instrument, Novato, USA) connected to an infusion 1162 pump (Legato 130 single syringe, 788130-KDS, KD Scientific, PHYMEP, Paris, France) at a speed of 1163 100 nl/minutes. The micropipette was left in place for an additional 5 minutes to allow viral 1164 dispersion and prevent backflow of the viral solution into the injection syringe. The scalp wound 1165 was closed with surgical sutures, and the mouse was kept in a warm environment until resuming 1166 normal activity. All animals were given analgesic and fluids before and after the surgery. 1167 In a separate set of mice, non-DREADD or DREADD (Dentate, Fastigial and Interposed) bundles of 1168 electrodes were implanted into the cerebellar nuclei, as described above. Both non-DREADD or 1169 DREADD injections and electrodes implantation were performed the same day. This experiment 1170 was performed in order to evaluate and validate that hM4D(Gi) receptors decrease the activity 1171 within the three cerebellar nuclei. Surgery, virus injections (AAV5-hSyn-hM4D(Gi)-mCherry or 1172 AAV5-hSyn-EGFP), coordinates (Fastigial: -6.0 AP, +/-0.75 ML, -2.1 depth from dura; Interposed: - 1173 6.0 AP, +/-1.5 ML, -2.1 depth from dura; Dentate: -6.2 AP, +/-2.3 ML, -2.4 depth from dura), 1174 chronic in vivo extracellular recordings and analysis were performed as we previously described 1175 above. One week following stereotaxic surgery to allow for virus expression, recordings at open- 1176 field were performed before and after CNO or saline (SAL) injection at the day 1, 4 and 7 of the 1177 accelerated rotarod task protocol. Mice were recorded for a 10 minutes baseline period followed 1178 by intraperitoneal injections of CNO 1mg/kg or SAL which were performed in a random 1179 sequence using a crossover design. After CNO or SAL injection, mice were recorded during 30 1180 minutes before and 15 minutes after the accelerated rotarod task protocol. 1181 1182 3.2. Cerebellar-thalamic outputs inactivation 1183 In order to inhibit specifically cerebellar outputs to the centrolateral (CL) and/or ventral anterior 1184 lateral (VAL) thalamus we applied a chemogenetic pathway-specific approach (Boender et al., 1185 2014). The technique comprises the combined use of a CRE-recombinase expressing canine 45

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1186 adenovirus-2 (CAV-2) and an adeno-associated virus (AAV-hSyn-DIO-hM4D(Gi)-mCherry) that 1187 contains the floxed inverted sequence of the DREADD hM4D(Gi)-mCherry. It entails the infusion 1188 of these two viral vectors into two sites that are connected through direct neuronal projections 1189 and represent a neuronal pathway. AAV-hSyn-DIO-hM4D(Gi)-mCherry is infused in the site 1190 where the cell bodies are located, while CAV-2 is infused in the area that is innervated by the 1191 corresponding axons. After infection of axonal terminals, CAV-2 is transported towards the cell 1192 bodies and expresses CRE-recombinase (Kremer et al., 2000; Hnasko et al., 2006). AAV-hSyn- 1193 DIO-hM4D(Gi)-mCherry contains the floxed inverted sequence of hM4D(Gi)-mCherry, which is 1194 reoriented in the presence of CRE, prompting the expression of hM4D(Gi)-mCherry. This ensures 1195 that hM4D(Gi)-mCherry is not expressed in all AAV-hSyn-DIO-hM4D(Gi)-mCherry infected 1196 neurons, but exclusively in those that are also infected with CAV-2. Using the same procedures 1197 described above, 0.4 μl of the retrograde canine adeno-associated cre virus (CAV-2-cre, titter ≥ 1198 2.5 × 108) (Plateforme de Vectorologie de Montpellier, Montpellier, France) was bilaterally 1199 injected in the CL (from bregma: AP -1.70 mm, ML ±0.75 mm, DV −3.0 mm) and VAL (from 1200 bregma: AP -1.4 mm, ML ±1.0 mm, DV −3.5 mm). In addition, 0.2 μl of AAV-hSyn-DIO-hM4D(Gi)- 1201 mCherry (UNC Vector Core, Chapel Hill, NC, USA) was bilaterally injected one week later into the 1202 cerebellar nuclei, focusing on the Dentate (from bregma: AP −6.2 mm, ML ±2.3 mm, DV −2.4 1203 mm) and Interposed (from bregma: AP −6.0 mm, ML ±1.5 mm, DV −2.1 mm) nucleus. Based on 1204 anatomical and functional evidences (Hintzen et ., 2018; Chen et al., 2014; Teune et al., 2000; 1205 Sakai 2000), in a group of mice we decided to inhibit those neurons that project from Dentate to 1206 CL and in another group we targeted those neurons that project from Interposed to VAL. All the 1207 stereotactic coordinates were determined based on The Mouse Brain Atlas (Paxinos and 1208 Franklin, 2004). 1209 1210 4. Behavioral experiments design 1211 Behavioral tests were performed one week following stereotaxic surgery to allow for virus 1212 expression. Balance beam, vertical pole, footprint patterns, grid test and fixed speed rotarod 1213 experiments were performed 30 minutes after CNO (1 mg/kg, ip) or SAL injections. Two different 1214 strategies were used for the accelerating rotarod motor learning task experiments: 1) CNO (1 1215 mg/kg, ip) or SAL was injected every day 30 minutes before the 1st trial of the accelerated 46

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1216 rotarod task. Four days later to ensure a proper CNO washout, mice were retested by receiving 7 1217 trials for two consecutive daily sessions. Drug-free mice received CNO (1mg/kg) or SAL 30 1218 minutes before the first trial in both days. The treatments were inverted meaning that those 1219 animals that received CNO during the preceding 7 days in this case were injected with SAL and 1220 the other way around. 2) CNO (1 mg/kg, ip) was injected 30 minutes after last trial at the day 1, 1221 2 and 3; subsequently mice received SAL 30 minutes after last trial at the day 4, 5 and 6 of the 1222 accelerated rotarod task. 1223 The DREADD ligand Clozapine-N-Oxide (CNO, TOCRIS, Bristol, UK) was dissolved in SAL (0.9% 1224 sodium chloride) and injected intraperitoneally at 1mg/kg. 1225 1226 5. Histology 1227 Mice were anesthetized with ketamine/xylazine (100 and 10 mg/kg, i.p., respectively) and 1228 rapidly perfused with ice-cold 4% paraformaldehyde in phosphate buffered SAL (PBS). The brains 1229 were carefully removed, postfixed in 4% paraformaldehyde for 24 h at 4 °C, cryoprotected in 1230 20% sucrose in PBS. The whole brain was cut into 40-μm-thick coronal sections on a cryostat 1231 (Thermo Scientific HM 560; Waltham, MA, USA). The sections were mounted on glass slides 1232 sealed with Mowiol mounting medium (Mowiol® 4-88; Sigma-Aldrich, France). Verification of 1233 virus injection site and DREADDs expression was assessed using a wide-field epifluorescent 1234 microscope (BX-43, Olympus, Waltham, MA, USA) using a mouse stereotaxic atlas (Paxinos and 1235 Franklin, 2004). We only kept mice showing a well targeted viral expression. Representative 1236 images of virus expression were acquired a Zeiss 800 Laser Scanning Confocal Microscope (×20 1237 objective, NA 0.8) (Carl Zeiss, Jena, Germany). Images were cropped and annotated using Zeiss 1238 Zen 2 Blue Edition software (for example images, see Fig. 5A, D). 1239 1240 1241 QUANTIFICATION AND STATISTICAL ANALYSIS 1242 Latency to fall in rotarod was analyzed by paired t-test for T1 vs T7 for each day (mean ± SEM). 1243 Distribution of numbers of trials in which cells were considered task related were compared 1244 using Mann-Whitney tests, Holm-Sidak corrected for multiple comparison between days. 1245 Evolution of the average firing rate (mean +/- SEM),) was normalized by subtracting the average 47

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1246 firing rate during the active part of the open-field session before the first rotarod trial for each 1247 day (one-way ANOVA repeated measured followed by Dunett Posthoc test). Linear regression 1248 between speed (open-field and rotarod) and neuronal activity were performed and we extracted 1249 the Pearson’s correlation coefficient, slope and its associated p-value. Comparisons between 1250 slops in open-field vs rotarod were performed using Wilcoxon test. One-way ANOVA repeated 1251 measure followed by Tukey Posthoc test was used to compare values of significant linear 1252 regression slopes for rotarod (mean +/- SEM) between phases. Monotonous relationship 1253 between significant linear regression slopes for rotarod and their associated Pearson’s 1254 correlation coefficient were assessed using Spearman Rank test for each brain structure and 1255 phases. Distributions of significant Pearson’s correlation coefficient were compared between 1256 brain structures using Mann Whitney test for each phase, followed by the computation of the 1257 overlapping index η (used as a unbiased non parametric estimate of the effect size). Mean 1258 change in quiet state firing rate between open-field after and before were compared using one- 1259 way ANOVA followed by Tukey Posthoc test for each phase, while the evolution of mean quiet 1260 state (open-field sessions and Inter-trial periods for each phase) was analyzed using one-way 1261 ANOVA followed by Dunett Posthoc test. Mean firing rate was normalized by subtracting the 1262 open-field before (OF1). Modulation of firing rate for chemogenetic experiments were analyzed 1263 using one-way ANOVA followed by Tukey Posthoc test for each phase. Latency to fall (mean ± 1264 S.E.M) in rotarod for chemogenetic experiments were analyzed using one-way ANOVA repeated 1265 measure followed by two types for Posthoc tests: paired t-test for repeated measure 1266 comparison and independent t-test for cross group comparisons. Locomotor activity (velocity) in 1267 open-field (mean ± S.E.M) was analyzed using two-way ANOVA repeated measure 1268 (treatment×moment) followed by t-test Posthoc (comparisons between treatments for each 1269 open-field session). Fixed speed rotarod (mean ± S.E.M) was analyzed using two-way ANOVA 1270 repeated measure (treatment×speed) followed by t-test Posthoc (comparisons between 1271 treatments for each speed steps). Footprint patterns parameters, horizontal bar, vertical pole 1272 and grid test were analyzed using one-way ANOVA. Data are represented as boxplots (median 1273 quartiles and interquartile range plus outliers). 1274 For the figure 6, to get a robust estimate of the initial and final performance of each day, we 1275 performed a linear regression on the latency to fall of each day and each animal; the within-day 48

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1276 and overnight loss was estimated from the start- and end-points of each regression segment. To 1277 estimate the interdependence of initial performance of the day, within day learning and inter- 1278 day learning, we used Deming linear regression, assuming equal variance of the noise of the 1279 measured quantity on the x- and y-axis. Confidence intervals were obtained by bootstrap. 1280

49

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1281 1282 REFERENCES 1283 Alexander, G.M., Rogan, S.C., Abbas, A.I., Armbruster, B.N., Pei, Y., Allen, J.A., Nonneman, R.J., 1284 Hartmann, J., Moy, S.S., Nicolelis, M.A., et al. (2009). Remote control of neuronal activity in 1285 transgenic mice expressing evolved G protein-coupled receptors. Neuron 63, 27-39. 1286 Anaclet, C., De Luca, R., Venner, A., Malyshevskaya, O., Lazarus, M., Arrigoni, E., and Fuller, P.M. 1287 (2018). Genetic Activation, Inactivation, and Deletion Reveal a Limited And Nuanced Role for 1288 Somatostatin-Containing Basal Forebrain Neurons in Behavioral State Control. The Journal of 1289 neuroscience : the official journal of the Society for Neuroscience 38, 5168-5181. 1290 Anaclet, C., Ferrari, L., Arrigoni, E., Bass, C.E., Saper, C.B., Lu, J., and Fuller, P.M. (2014). The 1291 GABAergic parafacial zone is a medullary slow wave sleep-promoting center. Nature 1292 neuroscience 17, 1217-1224. 1293 Anaclet, C., Pedersen, N.P., Ferrari, L.L., Venner, A., Bass, C.E., Arrigoni, E., and Fuller, P.M. (2015). 1294 Basal forebrain control of wakefulness and cortical rhythms. Nature communications 6, 8744. 1295 Aoki, S., Sato, Y., and Yanagihara, D. (2013). Lesion in the lateral cerebellum specifically produces 1296 overshooting of the toe trajectory in leading forelimb during obstacle avoidance in the rat. 1297 Journal of neurophysiology 110, 1511-1524. 1298 Armstrong, D.M., and Edgley, S.A. (1984). Discharges of nucleus interpositus neurones during 1299 locomotion in the cat. The Journal of physiology 351, 411-432. 1300 Aumann, T.D., and Horne, M.K. (1996). Ramification and termination of single axons in the 1301 cerebellothalamic pathway of the rat. The Journal of comparative neurology 376, 420-430. 1302 Baetens, K., Firouzi, M., Van Overwalle, F., and Deroost, N. (2020). Involvement of the 1303 cerebellum in the serial reaction time task (SRT) (Response to Janacsek et al.). NeuroImage 1304 220, 117114. 1305 Bearzatto, B., Servais, L., Cheron, G., and Schiffmann, S.N. (2005). Age dependence of strain 1306 determinant on mice motor coordination. Brain research 1039, 37-42. 1307 Bernard, J.A., and Seidler, R.D. (2013). Cerebellar contributions to visuomotor adaptation and 1308 motor sequence learning: an ALE meta-analysis. Frontiers in human neuroscience 7, 27. 1309 Bostan, A.C., Dum, R.P., and Strick, P.L. (2013). Cerebellar networks with the cerebral cortex and 1310 basal ganglia. Trends in cognitive sciences 17, 241-254.

50

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1311 Bostan, A.C., and Strick, P.L. (2010). The cerebellum and basal ganglia are interconnected. 1312 Neuropsychology review 20, 261-270. 1313 Brashers-Krug, T., Shadmehr, R., and Bizzi, E. (1996). Consolidation in human motor memory. 1314 Nature 382, 252-255. 1315 Buitrago, M.M., Schulz, J.B., Dichgans, J., and Luft, A.R. (2004). Short and long-term motor skill 1316 learning in an accelerated rotarod training paradigm. Neurobiology of learning and memory 81, 1317 211-216. 1318 Caggiano, V., Leiras, R., Goni-Erro, H., Masini, D., Bellardita, C., Bouvier, J., Caldeira, V., Fisone, G., 1319 and Kiehn, O. (2018). Midbrain circuits that set locomotor speed and gait selection. Nature 1320 553, 455-460. 1321 Caligiore, D., Pezzulo, G., Baldassarre, G., Bostan, A.C., Strick, P.L., Doya, K., Helmich, R.C., Dirkx, 1322 M., Houk, J., Jorntell, H., et al. (2017). Consensus Paper: Towards a Systems-Level View of 1323 Cerebellar Function: the Interplay Between Cerebellum, Basal Ganglia, and Cortex. Cerebellum 1324 16, 203-229. 1325 Cao, V.Y., Ye, Y., Mastwal, S., Ren, M., Coon, M., Liu, Q., Costa, R.M., and Wang, K.H. (2015). 1326 Motor Learning Consolidates Arc-Expressing Neuronal Ensembles in Secondary Motor Cortex. 1327 Neuron 86, 1385-1392. 1328 Carta, I., Chen, C.H., Schott, A.L., Dorizan, S., and Khodakhah, K. (2019). Cerebellar modulation of 1329 the reward circuitry and social behavior. Science 363. 1330 Chen, C.H., Fremont, R., Arteaga-Bracho, E.E., and Khodakhah, K. (2014). Short latency cerebellar 1331 modulation of the basal ganglia. Nature neuroscience 17, 1767-1775. 1332 Clopath, C., Badura, A., De Zeeuw, C.I., and Brunel, N. (2014). A cerebellar learning model of 1333 vestibulo-ocular reflex adaptation in wild-type and mutant mice. The Journal of neuroscience : 1334 the official journal of the Society for Neuroscience 34, 7203-7215. 1335 Cohen, D.A., Pascual-Leone, A., Press, D.Z., and Robertson, E.M. (2005). Off-line learning of 1336 motor skill memory: a double dissociation of goal and movement. Proceedings of the National 1337 Academy of Sciences of the United States of America 102, 18237-18241. 1338 Costa, R.M., Cohen, D., and Nicolelis, M.A. (2004). Differential corticostriatal plasticity during fast 1339 and slow motor skill learning in mice. Current biology : CB 14, 1124-1134.

51

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1340 Darmohray, D.M., Jacobs, J.R., Marques, H.G., and Carey, M.R. (2019). Spatial and Temporal 1341 Locomotor Learning in Mouse Cerebellum. Neuron 102, 217-231 e214. 1342 de Solages, C., Szapiro, G., Brunel, N., Hakim, V., Isope, P., Buisseret, P., Rousseau, C., Barbour, B., 1343 and Lena, C. (2008). High-frequency organization and synchrony of activity in the purkinje cell 1344 layer of the cerebellum. Neuron 58, 775-788. 1345 Doya, K. (1999). What are the computations of the cerebellum, the basal ganglia and the 1346 cerebral cortex? Neural networks : the official journal of the International Neural Network 1347 Society 12, 961-974. 1348 Doyon, J., Korman, M., Morin, A., Dostie, V., Hadj Tahar, A., Benali, H., Karni, A., Ungerleider, L.G., 1349 and Carrier, J. (2009). Contribution of night and day sleep vs. simple passage of time to the 1350 consolidation of motor sequence and visuomotor adaptation learning. Experimental brain 1351 research 195, 15-26. 1352 Durieux, P.F., Schiffmann, S.N., and de Kerchove d'Exaerde, A. (2012). Differential regulation of 1353 motor control and response to dopaminergic drugs by D1R and D2R neurons in distinct dorsal 1354 striatum subregions. The EMBO journal 31, 640-653. 1355 Galliano, E., Potters, J.W., Elgersma, Y., Wisden, W., Kushner, S.A., De Zeeuw, C.I., and Hoebeek, 1356 F.E. (2013). Synaptic transmission and plasticity at inputs to murine cerebellar Purkinje cells are 1357 largely dispensable for standard nonmotor tasks. The Journal of neuroscience : the official 1358 journal of the Society for Neuroscience 33, 12599-12618. 1359 Gomez, J.L., Bonaventura, J., Lesniak, W., Mathews, W.B., Sysa-Shah, P., Rodriguez, L.A., Ellis, R.J., 1360 Richie, C.T., Harvey, B.K., Dannals, R.F., et al. (2017). Chemogenetics revealed: DREADD 1361 occupancy and activation via converted clozapine. Science 357, 503-507. 1362 Gornati, S.V., Schäfer, C.B., Eelkman Rooda, O.H.J., Nigg, A.L., De Zeeuw, C.I., and Hoebeek, F.E. 1363 (2018). Differentiating Cerebellar Impact on Thalamic Nuclei. Cell reports 23, 2690-2704. 1364 Groszer, M., Keays, D.A., Deacon, R.M., de Bono, J.P., Prasad-Mulcare, S., Gaub, S., Baum, M.G., 1365 French, C.A., Nicod, J., Coventry, J.A., et al. (2008). Impaired synaptic plasticity and motor 1366 learning in mice with a point mutation implicated in human speech deficits. Current biology : 1367 CB 18, 354-362. 1368 Herzfeld, D.J., Kojima, Y., Soetedjo, R., and Shadmehr, R. (2018). Encoding of error and learning 1369 to correct that error by the Purkinje cells of the cerebellum. Nature neuroscience 21, 736-743. 52

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1370 Hewitt, A.L., Popa, L.S., and Ebner, T.J. (2015). Changes in Purkinje cell simple spike encoding of 1371 reach kinematics during adaption to a mechanical perturbation. The Journal of neuroscience : 1372 the official journal of the Society for Neuroscience 35, 1106-1124. 1373 Hintzen, A., Pelzer, E.A., and Tittgemeyer, M. (2018). Thalamic interactions of cerebellum and 1374 basal ganglia. Brain structure & function 223, 569-587. 1375 Hirata, H., Takahashi, A., Shimoda, Y., and Koide, T. (2016). Caspr3-Deficient Mice Exhibit Low 1376 Motor Learning during the Early Phase of the Accelerated Rotarod Task. PloS one 11, 1377 e0147887. 1378 Houck, B.D., and Person, A.L. (2015). Cerebellar Premotor Output Neurons Collateralize to 1379 Innervate the Cerebellar Cortex. The Journal of comparative neurology 523, 2254-2271. 1380 Huang, V.S., Haith, A., Mazzoni, P., and Krakauer, J.W. (2011). Rethinking motor learning and 1381 savings in adaptation paradigms: model-free memory for successful actions combines with 1382 internal models. Neuron 70, 787-801. 1383 Iscru, E., Serinagaoglu, Y., Schilling, K., Tian, J., Bowers-Kidder, S.L., Zhang, R., Morgan, J.I., 1384 DeVries, A.C., Nelson, R.J., Zhu, M.X., and Oberdick, J. (2009). Sensorimotor enhancement in 1385 mouse mutants lacking the Purkinje cell-specific Gi/o modulator, Pcp2(L7). Molecular and 1386 cellular neurosciences 40, 62-75. 1387 Ito, M. (2008). Control of mental activities by internal models in the cerebellum. Nature reviews 1388 Neuroscience 9, 304-313. 1389 Karni, A., Meyer, G., Rey-Hipolito, C., Jezzard, P., Adams, M.M., Turner, R., and Ungerleider, L.G. 1390 (1998). The acquisition of skilled motor performance: fast and slow experience-driven changes 1391 in . Proceedings of the National Academy of Sciences of the United States 1392 of America 95, 861-868. 1393 Khilkevich, A., Zambrano, J., Richards, M.M., and Mauk, M.D. (2018). Cerebellar implementation 1394 of movement sequences through feedback. eLife 7. 1395 Kida, H., Tsuda, Y., Ito, N., Yamamoto, Y., Owada, Y., Kamiya, Y., and Mitsushima, D. (2016). Motor 1396 Training Promotes Both Synaptic and Intrinsic Plasticity of Layer II/III Pyramidal Neurons in the 1397 Primary Motor Cortex. Cerebral cortex 26, 3494-3507.

53

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1398 Korman, M., Raz, N., Flash, T., and Karni, A. (2003). Multiple shifts in the representation of a 1399 motor sequence during the acquisition of skilled performance. Proceedings of the National 1400 Academy of Sciences of the United States of America 100, 12492-12497. 1401 Krakauer, J.W., Hadjiosif, A.M., Xu, J., Wong, A.L., and Haith, A.M. (2019). Motor Learning. 1402 Comprehensive Physiology 9, 613-663. 1403 Li, W., Ma, L., Yang, G., and Gan, W.B. (2017). REM sleep selectively prunes and maintains new 1404 synapses in development and learning. Nature neuroscience 20, 427-437. 1405 Longley, M., and Yeo, C.H. (2014). Distribution of neural plasticity in cerebellum-dependent 1406 motor learning. Progress in brain research 210, 79-101. 1407 Low, A.Y.T., Thanawalla, A.R., Yip, A.K.K., Kim, J., Wong, K.L.L., Tantra, M., Augustine, G.J., and 1408 Chen, A.I. (2018). Precision of Discrete and Rhythmic Forelimb Movements Requires a Distinct 1409 Neuronal Subpopulation in the Interposed Anterior Nucleus. Cell reports 22, 2322-2333. 1410 Machado, A.S., Darmohray, D.M., Fayad, J., Marques, H.G., and Carey, M.R. (2015). A 1411 quantitative framework for whole-body coordination reveals specific deficits in freely walking 1412 ataxic mice. eLife 4. 1413 Marple-Horvat, D.E., and Criado, J.M. (1999). Rhythmic neuronal activity in the lateral 1414 cerebellum of the cat during visually guided stepping. The Journal of physiology 518 ( Pt 2), 1415 595-603. 1416 Mauk, M.D., Li, W., Khilkevich, A., and Halverson, H. (2014). Cerebellar mechanisms of learning 1417 and plasticity revealed by delay eyelid conditioning. International review of neurobiology 117, 1418 21-37. 1419 Morehead, J.R., Qasim, S.E., Crossley, M.J., and Ivry, R. (2015). Savings upon Re-Aiming in 1420 Visuomotor Adaptation. The Journal of neuroscience : the official journal of the Society for 1421 Neuroscience 35, 14386-14396. 1422 Morton, S.M., and Bastian, A.J. (2006). Cerebellar contributions to locomotor adaptations during 1423 splitbelt treadmill walking. The Journal of neuroscience : the official journal of the Society for 1424 Neuroscience 26, 9107-9116. 1425 Muellbacher, W., Ziemann, U., Wissel, J., Dang, N., Kofler, M., Facchini, S., Boroojerdi, B., Poewe, 1426 W., and Hallett, M. (2002). Early consolidation in human primary motor cortex. Nature 415, 1427 640-644. 54

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1428 Nagai, H., de Vivo, L., Bellesi, M., Ghilardi, M.F., Tononi, G., and Cirelli, C. (2017). Sleep 1429 Consolidates Motor Learning of Complex Movement Sequences in Mice. Sleep 40. 1430 Nguyen-Vu, T.D., Kimpo, R.R., Rinaldi, J.M., Kohli, A., Zeng, H., Deisseroth, K., and Raymond, J.L. 1431 (2013). Cerebellar Purkinje cell activity drives motor learning. Nature neuroscience 16, 1734- 1432 1736. 1433 Ohmae, S., and Medina, J.F. (2015). Climbing fibers encode a temporal-difference prediction 1434 error during cerebellar learning in mice. Nature neuroscience 18, 1798-1803. 1435 Pastore, M., and Calcagnì, A. (2019). Measuring Distribution Similarities Between Samples: A 1436 Distribution-Free Overlapping Index. Frontiers in psychology 10, 1089. 1437 Paxinos, G.F., and Franklin, K.B.J. (2004). The Mouse Brain in Stereotaxic Coordinates 1438 (Amsterdam: Elsevier). 1439 Paz, R., Pelletier, J.G., Bauer, E.P., and Paré, D. (2006). Emotional enhancement of memory via 1440 amygdala-driven facilitation of rhinal interactions. Nature neuroscience 9, 1321-1329. 1441 Pedersen, N.P., Ferrari, L., Venner, A., Wang, J.L., Abbott, S.B.G., Vujovic, N., Arrigoni, E., Saper, 1442 C.B., and Fuller, P.M. (2017). Supramammillary glutamate neurons are a key node of the 1443 arousal system. Nature communications 8, 1405. 1444 Popa, D., Spolidoro, M., Proville, R.D., Guyon, N., Belliveau, L., and Léna, C. (2013). Functional 1445 role of the cerebellum in gamma-band synchronization of the sensory and motor cortices. The 1446 Journal of neuroscience : the official journal of the Society for Neuroscience 33, 6552-6556. 1447 Porras-Garcia, M.E., Ruiz, R., Perez-Villegas, E.M., and Armengol, J.A. (2013). Motor learning of 1448 mice lacking cerebellar Purkinje cells. Frontiers in neuroanatomy 7, 4. 1449 Proville, R.D., Spolidoro, M., Guyon, N., Dugué, G.P., Selimi, F., Isope, P., Popa, D., and Léna, C. 1450 (2014). Cerebellum involvement in cortical sensorimotor circuits for the control of voluntary 1451 movements. Nature neuroscience 17, 1233-1239. 1452 Raymond, J.L., and Medina, J.F. (2018). Computational Principles of Supervised Learning in the 1453 Cerebellum. Annual review of neuroscience 41, 233-253. 1454 Roth, B.L. (2016). DREADDs for Neuroscientists. Neuron 89, 683-694. 1455 Rothwell, P.E., Fuccillo, M.V., Maxeiner, S., Hayton, S.J., Gokce, O., Lim, B.K., Fowler, S.C., 1456 Malenka, R.C., and Sudhof, T.C. (2014). Autism-associated neuroligin-3 mutations commonly 1457 impair striatal circuits to boost repetitive behaviors. Cell 158, 198-212. 55

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1458 Rueda-Orozco, P.E., and Robbe, D. (2015). The striatum multiplexes contextual and kinematic 1459 information to constrain motor habits execution. Nature neuroscience 18, 453-460. 1460 Ruitenberg, M.F.L., Koppelmans, V., De Dios, Y.E., Gadd, N.E., Wood, S.J., Reuter-Lorenz, P.A., 1461 Kofman, I., Bloomberg, J.J., Mulavara, A.P., and Seidler, R.D. (2018). Neural correlates of multi- 1462 day learning and savings in sensorimotor adaptation. Scientific reports 8, 14286. 1463 Sakayori, N., Kato, S., Sugawara, M., Setogawa, S., Fukushima, H., Ishikawa, R., Kida, S., and 1464 Kobayashi, K. (2019). Motor skills mediated through cerebellothalamic tracts projecting to the 1465 central lateral nucleus. Molecular brain 12, 13. 1466 Samaei, A., Ehsani, F., Zoghi, M., Hafez Yosephi, M., and Jaberzadeh, S. (2017). Online and offline 1467 effects of cerebellar transcranial direct current stimulation on motor learning in healthy older 1468 adults: a randomized double-blind sham-controlled study. The European journal of 1469 neuroscience 45, 1177-1185. 1470 Sano, T., Kohyama-Koganeya, A., Kinoshita, M.O., Tatsukawa, T., Shimizu, C., Oshima, E., Yamada, 1471 K., Le, T.D., Akagi, T., Tohyama, K., et al. (2018). Loss of GPRC5B impairs synapse formation of 1472 Purkinje cells with cerebellar nuclear neurons and disrupts cerebellar synaptic plasticity and 1473 motor learning. Neuroscience research 136, 33-47. 1474 Sarnaik, R., and Raman, I.M. (2018). Control of voluntary and optogenetically perturbed 1475 locomotion by spike rate and timing of neurons of the mouse cerebellar nuclei. eLife 7. 1476 Sauerbrei, B.A., Lubenov, E.V., and Siapas, A.G. (2015). Structured Variability in Purkinje Cell 1477 Activity during Locomotion. Neuron 87, 840-852. 1478 Sausbier, M., Hu, H., Arntz, C., Feil, S., Kamm, S., Adelsberger, H., Sausbier, U., Sailer, C.A., Feil, R., 1479 Hofmann, F., et al. (2004). Cerebellar ataxia and Purkinje cell dysfunction caused by Ca2+- 1480 activated K+ channel deficiency. Proceedings of the National Academy of Sciences of the 1481 United States of America 101, 9474-9478. 1482 Schwartz, A.B., Ebner, T.J., and Bloedel, J.R. (1987). Responses of interposed and dentate 1483 neurons to perturbations of the locomotor cycle. Experimental brain research 67, 323-338. 1484 Seidler, R.D. (2010). Neural correlates of motor learning, transfer of learning, and learning to 1485 learn. Exercise and sport sciences reviews 38, 3-9.

56

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1486 Seidler, R.D., Purushotham, A., Kim, S.G., Ugurbil, K., Willingham, D., and Ashe, J. (2002). 1487 Cerebellum activation associated with performance change but not motor learning. Science 1488 296, 2043-2046. 1489 Shadmehr, R., and Holcomb, H.H. (1997). Neural correlates of motor memory consolidation. 1490 Science 277, 821-825. 1491 Spencer, R.M., and Ivry, R.B. (2009). Sequence learning is preserved in individuals with cerebellar 1492 degeneration when the movements are directly cued. Journal of cognitive neuroscience 21, 1493 1302-1310. 1494 Steriade, M. (1995). Two channels in the cerebellothalamocortical system. The Journal of 1495 comparative neurology 354, 57-70. 1496 Stroobants, S., Gantois, I., Pooters, T., and D'Hooge, R. (2013). Increased gait variability in mice 1497 with small cerebellar cortex lesions and normal rotarod performance. Behavioural brain 1498 research 241, 32-37. 1499 Venner, A., Anaclet, C., Broadhurst, R.Y., Saper, C.B., and Fuller, P.M. (2016). A Novel Population 1500 of Wake-Promoting GABAergic Neurons in the Ventral Lateral Hypothalamus. Current biology : 1501 CB 26, 2137-2143. 1502 Vinueza Veloz, M.F., Zhou, K., Bosman, L.W., Potters, J.W., Negrello, M., Seepers, R.M., Strydis, C., 1503 Koekkoek, S.K., and De Zeeuw, C.I. (2015). Cerebellar control of gait and interlimb coordination. 1504 Brain structure & function 220, 3513-3536. 1505 Watanave, M., Matsuzaki, Y., Nakajima, Y., Ozawa, A., Yamada, M., and Hirai, H. (2018). 1506 Contribution of Thyrotropin-Releasing Hormone to Cerebellar Long-Term Depression and 1507 Motor Learning. Frontiers in cellular neuroscience 12, 490. 1508 Yang, G., Pan, F., and Gan, W.B. (2009). Stably maintained dendritic spines are associated with 1509 lifelong memories. Nature 462, 920-924. 1510 Yang, Y., and Lisberger, S.G. (2014). Purkinje-cell plasticity and cerebellar motor learning are 1511 graded by complex-spike duration. Nature 510, 529-532. 1512 Yin, H.H., Mulcare, S.P., Hilario, M.R., Clouse, E., Holloway, T., Davis, M.I., Hansson, A.C., Lovinger, 1513 D.M., and Costa, R.M. (2009). Dynamic reorganization of striatal circuits during the acquisition 1514 and consolidation of a skill. Nature neuroscience 12, 333-341. 1515 57

bioRxiv preprint doi: https://doi.org/10.1101/2020.08.27.270330; this version posted August 27, 2020. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-ND 4.0 International license.

1516 1517

58