Abstract

Full-length RIM1 and 2 are key components of the presynaptic active zone that ubiquitously control excitatory and inhibitory neurotransmitter release. Here, we report that the function of the small RIM isoform RIM4, consisting of a single C2 domain, is strikingly different from that of the long isoforms.

RIM4 is dispensable for neurotransmitter release but plays a postsynaptic, cell type-specific role in cerebellar Purkinje cells that is essential for normal motor function. In the absence of RIM4, Purkinje cell intrinsic firing is reduced and caffeine-sensitive, and dendritic integration of climbing fibre input is disturbed. Mice lacking RIM4, but not mice lacking RIM1/2, selectively in Purkinje cells exhibit a severe, hours-long paroxysmal dystonia. These episodes can also be induced by caffeine, ethanol or stress and closely resemble the deficits seen with mutations of the PNKD (paroxysmal non-kinesigenic dystonia) gene.

Our data reveal essential postsynaptic functions of RIM proteins and show non-overlapping specialized functions of a small isoform despite high homology to a single domain in the full-length proteins.

Introduction

The RIM family of proteins is known for its ubiquitous and important function in synaptic transmission and synaptic plasticity.1 Not only glutamatergic and GABAergic synapses depend on RIM proteins, but also varicosities that release neuromodulators, such as dopamine.2 Knockout (KO) mice of RIM1α, the full-length variant of the RIM1 gene, show deficits in spatial memory and social interaction as well as a decrease in prepulse inhibition.3,4 Mice deficient for both full-length variants, RIM1α and RIM2α, are not viable.5 Within the synapse, the long RIM family members RIM1 and RIM2 maintain a high density of presynaptic calcium channels and keep synaptic vesicles in a release-ready state.5-10 RIM1α and RIM2α comprise several well characterized protein domains through which they interact with most other active zone (AZ) proteins, the synaptic vesicle protein Rab3 and voltage-gated calcium channels (VGCCs).11 The C-terminal C2 domain of RIM1 and RIM2, the so-called C2B domain, allows membrane anchoring of RIM1/2 and is likely to be particularly important for the spatial coordination of AZ proteins, VGCCs and synaptic vesicles.12,13

The RIM protein family also includes two shorter members, RIM3 and RIM4,11 which are widely expressed throughout the CNS.14 These short RIM proteins comprise only a single C2-domain, which is highly homologous to the C2B domain of RIM1/2.11 As such, single domain proteins RIM3 and RIM4 are unlikely to have similar AZ-organizing functions as their long RIM family counterparts. We have previously shown, using a knockdown strategy, that RIM3 and RIM4 are important for dendritic outgrowth.14 It has also been described that RIM3 and RIM4,15 as well as the C2B domains of RIM1/2,16 can interact with β-subunits of VGCCs to suppress calcium-induced inactivation. However, it is not clear whether suppression of inactivation by RIM3/4 alters action potential-mediated vesicle release from classical synapses, where calcium-induced inactivation may only play a minor role. Furthermore, such suppression may be dominated by RIM1/2, which are abundant in presynapses and have a much higher affinity for β-subunits.15 Taken together, the exact role of RIM3 and RIM4 in CNS function remains largely unresolved.

The results presented here show that RIM4 plays a unique role in the CNS that is very different from the well studied function of RIM1/2. RIM4 acts in a strictly pathway-specific manner and is primarily required by Purkinje cells in the cerebellum. RIM4 does not play a role in transmitter release but acts postsynaptically to regulate somatodendritic excitability of Purkinje cells. During development, RIM4 is required for dendrite outgrowth of Purkinje cells and for the correct localization of synapses along their dendrites. Mice lacking RIM4 develop a severe motor phenotype in the form of a paroxysmal dystonia, not known to result from deletion of the long RIM isoforms. This dystonia shares many features with Cav2.1 channelopathies, such as tottering mice, but most closely resembles the paroxysmal non-kinesigenic dystonia (PNKD) caused by mutations in the PNKD protein that abolish binding to the C2B domain of RIM1/2.17

Materials and methods

A full description of the experimental procedures is provided in the Supplementary material.

Experimental model and subject details

Generation of mouse models and source of cell cultures used in this study are reported in the Supplementary material.

Immunoblotting

Immunoblotting of lysates from brain tissue and HEK293T cells was performed as previously described.14,18,19 Information about antibodies is reported in the Supplementary material.

Plasmids

The AAV-CMV-mRFP and AAV-CMV-mRFP-T2A-Flag-RIM4 plasmids were generated by restriction cloning. Details about restriction enzymes and primers are provided in the Supplementary material.

Recombinant AAV production

Recombinant AAV1/2 particles were generated by triple transfection of HEK293T cells and purified as described previously.19-21

Primary neuronal cultures

Primary cortical neurons were prepared from E16–E19 C57Bl/6N mouse brains, as previously described.19-21

Data analysis

Data are presented as means ± standard errors of the mean (SEM), unless otherwise noted. Statistical significance was determined as indicated in the figure legends. Values were considered significant at P < 0.05. Statistical analyses were performed using Igor Pro 8 or Prism GraphPad (version 6.02).

Results

RIM4 deficiency causes severe paroxysmal dyskinesia and dystonia

To investigate the functional role of RIM4 in the CNS, we generated constitutive RIM4 knockout (KOconst) mice by crossing conditional RIM4fl/fl mice with the ubiquitously expressed PGK Cre driver line (Fig. 1A). Ablation of RIM4 protein expression was verified in the hippocampus, cortex and cerebellum of RIM4 KOconst mice by quantitative immunoblotting using an isoform-specific antibody raised against the RIM4 N-terminus (Fig. 1B and C). RIM4 KOconst mice were born at the expected Mendelian ratio (data not shown) and showed no obvious abnormalities during the first two postnatal weeks. However, starting around postnatal Day 21 (P21), RIM4 KOconst mice gained less weight than wild-type (WT) littermates (Fig. 1D). Around this time, RIM4 KOconst mice began to develop hours-long spontaneous and stress-induced (e.g. handling) dyskinetic and dystonic episodes. RIM4 KOconst mice and their wild-type littermates were continuously EEG (cortex) and video monitored (24/7) in their home cage for five consecutive days. We roughly divided dyskinetic-dystonic episodes into three phases (I–III, Fig. 1E), which correspond well to the severity scales 2–4 (mild to severe) proposed for grading dystonia symptoms.22 Phase I (Fig. 1E and F) lasted up to 10 min and consisted of seconds-long repetitive jerking of the forelimbs and dystonic upward extensions of one of the hindlimbs. Ambulation was largely maintained and mice only occasionally fell to the side. Phase II (∼10–15 min) was characterized by additional repetitive and sustained neck flexion together with extension of the trunk. Abnormal postures and falls were more frequent and a typical movement was to push the head into the bedding while the hindlimbs performed a pedalling motion. Phase III lasted ∼3 h and was associated with severe immobility and persistent abnormal postures with almost no ambulation. RIM4 KOconst mice developed dystonic episodes mostly once per day (1 ± 0.98 per day; Fig. 1G). The most frequent onset of motor episodes was in the early morning hours, suggesting a circadian-like pattern (Fig. 1G). In the EEG recordings we did not detect spike-and-wave discharges associated with periods of immobility during the episodes.29-31 Both ethanol and caffeine also induced severe dyskinetic-dystonic episodes in RIM4 KOconst mice, phenocopying the spontaneously occurring episodes (90% and 100% of RIM4 KOconst mice, respectively; Fig. 1H).

Constitutive RIM4 KO mice exhibit ataxia and inducible severe and long-lasting dystonia. (A) Scheme showing (from top to bottom) the structure of the mouse Rims4 gene (exons shown as black boxes), an expanded map around exon 2, the targeting vector (red triangles = loxP sites, grey ovals = FRT recombination sites, light blue box = splice acceptor site, dark blue box = LacZ cassette, N = neomycin resistance cassette, DT = diphtheria toxin cassette), KO (knockout) first allele (LacZ cassette is spliced in after exon 1 resulting in LacZ expression and RIM4 KO), RIM4fl/fl allele (neomycin resistance cassette was removed by flp recombination), KO allele (Cre recombination deleted exon 2). (B) Representative immunoblot and quantitative analysis (C) of lysates from hippocampus (HC), cortex (CX) and cerebellum (CB) of RIM4 wild-type (WT) (+/+), heterozygous (+/−) and KOconst (−/−) mice. OE = HEK293 T cell lysate expressing RIM4. β-Actin was used as loading control. n = 10 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (D) Graph showing body weight development of RIM4 wild-type (+/+, black), heterozygous (+/−, grey) and KOconst (−/−, white) mice over 35 days. n = 5 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (E) Representative images of different phases of motor impairment in RIM4 KOconst mice: phases I, II and III. (F) Duration of the three phases of motor impairment in individual mice during the episode. The grey colour indicates the phase. Note the broken x-axis. (G) Raster plots showing the onset and duration of spontaneously occurring episodes of motor impairment (solid black line) during a 5-day observation period. WT n = 6, KOconstn = 10 mice (13–17 weeks). (H) Diagram showing the susceptibility of RIM4 WT and KOconst mice to the induction of an episode of motor impairment by the intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse) or caffeine (Caff, 25 mg/kg mouse). WT n = 8, KOconstn = 10 mice. (I) Example image of paw prints (left) and footfall/gait pattern (right) of adult RIM4 WT and KOconst mice on the CatWalk. (J) Quantification of the regularity index of the footfall/gait pattern (number of normal step sequences × 4 / number of paw placements), n = 3 per group, unpaired t-test. (K) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 WT and KOconst mice, n = 10 per group, unpaired t-test. (L) RIM4 WT and KOconst mice were placed in a cage with a running wheel and monitored for four consecutive days: regular wheel (RW) consisting of all 38 rungs for 2 days and complex wheel (CX) consisting of 22 rungs (16 rungs were selected and removed) for 2 days. Black boxes indicate time spent running in the wheel, black lines represent episodes of motor impairment. WT n = 7, KOconstn = 6 mice. (M) Graph of the average instantaneous running speed of the mice shown in L. One-way ANOVA tests with Sidak's post hoc correction.
Figure 1

Constitutive RIM4 KO mice exhibit ataxia and inducible severe and long-lasting dystonia. (A) Scheme showing (from top to bottom) the structure of the mouse Rims4 gene (exons shown as black boxes), an expanded map around exon 2, the targeting vector (red triangles = loxP sites, grey ovals = FRT recombination sites, light blue box = splice acceptor site, dark blue box = LacZ cassette, N = neomycin resistance cassette, DT = diphtheria toxin cassette), KO (knockout) first allele (LacZ cassette is spliced in after exon 1 resulting in LacZ expression and RIM4 KO), RIM4fl/fl allele (neomycin resistance cassette was removed by flp recombination), KO allele (Cre recombination deleted exon 2). (B) Representative immunoblot and quantitative analysis (C) of lysates from hippocampus (HC), cortex (CX) and cerebellum (CB) of RIM4 wild-type (WT) (+/+), heterozygous (+/−) and KOconst (−/−) mice. OE = HEK293 T cell lysate expressing RIM4. β-Actin was used as loading control. n = 10 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (D) Graph showing body weight development of RIM4 wild-type (+/+, black), heterozygous (+/−, grey) and KOconst (−/−, white) mice over 35 days. n = 5 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (E) Representative images of different phases of motor impairment in RIM4 KOconst mice: phases I, II and III. (F) Duration of the three phases of motor impairment in individual mice during the episode. The grey colour indicates the phase. Note the broken x-axis. (G) Raster plots showing the onset and duration of spontaneously occurring episodes of motor impairment (solid black line) during a 5-day observation period. WT n = 6, KOconstn = 10 mice (13–17 weeks). (H) Diagram showing the susceptibility of RIM4 WT and KOconst mice to the induction of an episode of motor impairment by the intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse) or caffeine (Caff, 25 mg/kg mouse). WT n = 8, KOconstn = 10 mice. (I) Example image of paw prints (left) and footfall/gait pattern (right) of adult RIM4 WT and KOconst mice on the CatWalk. (J) Quantification of the regularity index of the footfall/gait pattern (number of normal step sequences × 4 / number of paw placements), n = 3 per group, unpaired t-test. (K) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 WT and KOconst mice, n = 10 per group, unpaired t-test. (L) RIM4 WT and KOconst mice were placed in a cage with a running wheel and monitored for four consecutive days: regular wheel (RW) consisting of all 38 rungs for 2 days and complex wheel (CX) consisting of 22 rungs (16 rungs were selected and removed) for 2 days. Black boxes indicate time spent running in the wheel, black lines represent episodes of motor impairment. WT n = 7, KOconstn = 6 mice. (M) Graph of the average instantaneous running speed of the mice shown in L. One-way ANOVA tests with Sidak's post hoc correction.

Automated gait analysis using a video-based CatWalk system revealed that even outside of dyskinetic-dystonic episodes, RIM4 KOconst mice displayed deficiencies in motor coordination compared to the wild-type mice, such as gait abnormalities, as evidenced by an irregular footprint pattern (Fig. 1I and J). RIM4 KOconst mice also performed worse in the accelerating rotarod test and showed a significantly reduced latency to fall compared to wild-type littermates (Fig. 1K). RIM4 KOconst mice also performed worse in the ledge test (Supplementary Fig. 1A) and the hind limb clasping assay (Supplementary Fig. 1B). RIM4 wild-type and KOconst mice were given the opportunity to use a regular or complex running wheel in their home cage and were then observed for 24 h/day for four consecutive days (2 days regular wheel and 2 days complex wheel, Fig. 1L). Consistent with only minor deficits in regular walking, no difference in mean running speed was observed between RIM4 wild-type and KOconst mice when running on wheels equipped with rungs in a regular pattern. However, RIM4 KOconst mice performed significantly worse when the pattern of the rungs in the wheel was made more complex by removing individual rungs (16 of 38) (Fig. 1L and M). While RIM4 wild-type mice showed a small increase in their instantaneous average running speed even in the complex wheel, RIM4 KOconst mice became significantly slower. RIM4 KOconst mice had fewer running sessions per time, not taking the duration of episodes into account [wild-type, 36.1 ± 3.7 in running wheel (RW), 37.4 ± 2.8 in complex wheel (CW); KOconst 15.2 ± 3.2 in RW, 19.1 ± 2.2 in CW] than their wild-type littermates. Nevertheless, RIM KOconst mice occasionally attempted to use the wheel even when they were in a dystonic episode (black strokes on red bars in Fig. 1L). Taken together, the absence of RIM4 causes deficits in motor coordination and a circadian, long-lasting episodic motor impairment.

Synaptic transmission is not affected by RIM4 deficiency

To test whether ablation of RIM4 affects synaptic transmission, we performed whole-cell voltage clamp recordings from hippocampal CA1 pyramidal neurons in RIM4 wild-type and KOconst mice. Both the mean amplitude and frequency of miniature excitatory postsynaptic currents (mEPSCs) were not altered in RIM4 KOconst mice (Fig. 2A–C, 1 µM tetrodotoxin, 20 µM CNQX, 100 µM D-APV). Paired-pulse facilitation of field excitatory postsynaptic potentials (fEPSPs) recorded in the CA1 stratum radiatum and evoked by Schaffer collateral stimulation was unaffected in RIM4 KOconst mice (Fig. 2D). Furthermore, quantitative immunoblotting showed no significant changes in the expression of other components of the presynaptic release machinery, such as RIM1 and Liprin-α, in hippocampal lysates from RIM4 KOconst mice compared to wild-type (Fig. 2E). These findings were further supported by quantitative transcriptomic (total RNA sequencing, Supplementary Fig. 2A) and proteomic (Supplementary Fig. 2B) analyses, which did not reveal any significant changes in the levels of synaptic proteins in RIM4 KOconst mice. Levels of the other short RIM variant, RIM3, appeared increased in hippocampus (Fig. 2E), but not in cerebellum and cortex (Supplementary Fig. 2B, proteomics data). In conclusion, RIM4 appears to be dispensable for synaptic transmitter release in CA1 of the hippocampus.

In contrast to the ubiquitous role of long RIM family members, RIM4 does not act by modulating release. (A) Representative miniature excitatory postsynaptic current (mEPSC) recordings from RIM4 wild-type (WT, black) and KOconst (grey) mice. Scale bars = 2 s (horizontal) and 2 pA (vertical). (B and C) Cumulative distribution function (CDF) histograms of mean amplitude (B) and inter-event interval (C) of mEPSCs. n mice/n cells: WT 8/12, KOconst 4/5. (D) Representative traces of field excitatory postsynaptic potentials (fEPSPs) of RIM4 WT (black) and KOconst (grey) (left) and bar graph showing the mean paired-pulse ratio (PPR) (right). n mice/n cells: WT 8/32, KOconst 4/9, unpaired t-test. (E) Example images and quantification of immunoblots of hippocampal lysates from RIM4 WT and KOconst mice labelled with antibodies against RIM4, RIM1, RIM3, Liprin-α, PSD95 and β-actin. n = 5 per group. KO = knockout.
Figure 2

In contrast to the ubiquitous role of long RIM family members, RIM4 does not act by modulating release. (A) Representative miniature excitatory postsynaptic current (mEPSC) recordings from RIM4 wild-type (WT, black) and KOconst (grey) mice. Scale bars = 2 s (horizontal) and 2 pA (vertical). (B and C) Cumulative distribution function (CDF) histograms of mean amplitude (B) and inter-event interval (C) of mEPSCs. n mice/n cells: WT 8/12, KOconst 4/5. (D) Representative traces of field excitatory postsynaptic potentials (fEPSPs) of RIM4 WT (black) and KOconst (grey) (left) and bar graph showing the mean paired-pulse ratio (PPR) (right). n mice/n cells: WT 8/32, KOconst 4/9, unpaired t-test. (E) Example images and quantification of immunoblots of hippocampal lysates from RIM4 WT and KOconst mice labelled with antibodies against RIM4, RIM1, RIM3, Liprin-α, PSD95 and β-actin. n = 5 per group. KO = knockout.

RIM4 deletion specifically causes morphological changes in the cerebellum

The cerebellum constitutes a critical circuit for motor coordination and control and RIM4 is robustly expressed in the cerebellum (Fig. 1B and Supplementary Fig. 3). We analysed the size of the cerebellum in Nissl-stained sagittal sections of RIM4 wild-type and KOconst mice (Fig. 3A) and found a significant reduction in the cerebellar area in mice lacking RIM4 (Fig. 3B). In contrast, the size of the cerebral cortex and hippocampus was unchanged in RIM4 KOconst mice (Fig. 3B), despite robust RIM4 expression in those regions (Supplementary Fig. 3). Quantification of the size of the individual cerebellar layers revealed a significantly reduced area of the molecular layer (ML) in RIM4 KOconst mice, whereas the size of the granule cell layer and the white matter did not differ between wild-type and KOconst (Fig. 3C and D). In contrast to the reduced thickness of the molecular layer in RIM4 KOconst mice (Fig. 3E), the total length of the Purkinje cell layer (Fig. 3F) and the density of Purkinje cells along this length (Fig. 3G) were unchanged, as was the estimated number of Purkinje cells (Fig. 3H). We labelled the dendritic tree of individual Purkinje cells by single cell electroporation of TMR-Dextran and performed a 3D reconstruction of Purkinje cells in confocal image stacks (Fig. 3I). The area covered by the dendritic tree of Purkinje cells was significantly reduced in RIM4 KOconst mice compared to wild-type mice (Fig. 3J), and a 3D-Sholl analysis revealed a significantly reduced dendritic arborization starting at a distance of ∼50 µm from the soma (Fig. 3K). A detailed analysis showed that, in particular, fourth to sixth order dendritic segments were significantly shorter in Purkinje cells of RIM4 KOconst mice (Fig. 3L).

RIM4 plays a pathway-specific role in cerebellar Purkinje cells and appears dispensable in hippocampus and cerebral cortex. (A) Representative images of sagittal brain sections stained with fluorescent Nissl green from RIM4 wild-type (WT) (left) and KOconst (right) mice. White lines delineate the measured areas, cortex (CX), cerebellum (CB) and hippocampus (HC). Scale bars = 1 mm. (B) Bar graph showing the quantitative analysis of the area of the analysed brain regions, CX (WT 12.46 ± 0.12 mm2, KOconst 11.99 ± 0.39 mm2), CB (WT 7.78 ± 0.36 mm2, KOconst 6.47 ± 0.13 mm2) and HC (WT 1.43 ± 0.05 mm2, KOconst 1.43 ± 0.05 mm2), normalized to the WT value. n = 4 mice per group (12–15 weeks), unpaired t-test. (C) Example images of calbindin (magenta) and Hoechst (blue) stained cerebellar sections from RIM4 wild-type and KOconst mice. White lines indicate measured subcerebellar areas, molecular layer (ML), granule cell layer (GCL), white matter (WM). Scale bars = 1 mm. (D) Bar graph showing the quantitative analysis of the area of the analysed subcerebellar regions, molecular layer (ML) (WT 4.04 ± 0.25 mm2, KOconst 2.99 ± 0.15 mm2), granule cell layer (GCL) (WT 3.22 ± 0.22 mm2, KOconst 2.59 ± 0.1 mm2), white matter (WM) (WT 0.95 ± 0.13 mm2, KOconst 0.88 ± 0.07 mm2), normalized to the wild-type value. n = 4 mice per group (12–15 weeks), unpaired t-test. (E–H) ML thickness (WT 173.05 ± 4.36 μm, KOconst 151.08 ± 2.83 μm), total Purkinje cell layer (PCL) length, and Purkinje cell density (WT 4.18 ± 0.13 / 100 μm, KOconst 4.61 ± 0.14 / 100 μm) were measured in calbindin-stained cerebellar tissue from RIM4 wild-type and KOconst mice. The estimated number of Purkinje cells was calculated by multiplying the total PCL length by the Purkinje cell density. n = 4 mice per group (12–15 weeks), unpaired t-test. (I) Representative confocal image of Purkinje cells labelled with the red fluorescent dye TMR-Dextran (1 mM) in RIM4 wild-type and KOconst mice. Maximum intensity projections (top) and 3D reconstructed dendritic trees (right) with branch level annotations (fourth = blue; fifth = yellow; sixth = red). Scale bars = 50 µm. (J) Bar graph showing the quantification of the dendritic tree area in RIM4 wild-type and KOconst mice (WT 1.92 ± 0.12 × 10−2 mm2, KOconst 1.38 ± 0.08 × 10−2 mm2). n mice/n cells: WT 11/22, KOconst 10/30 (>100 days), unpaired t-test. (K) Graph showing Sholl intersections measured from 3D reconstructed Purkinje cells plotted against the radial distance from the cell body in RIM4 wild-type and KOconst mice. WT n = 11, KOconstn = 10 mice (>100 days), Kolmogorov-Smirnov test. (L) Bar graph showing quantification of dendritic length for each branching level (branch level 4: WT 1.6 ± 0.1 mm, KOconst 1.1 ± 0.04 mm; branch level 5: WT 1.4 ± 0.1 mm, KOconst 0.7 ± 0.04 mm; branch level 6: WT 0.9 ± 0.1 mm, KOconst 0.4 ± 0.04 mm). WT n = 11 KOconstn = 10 mice (>100 days), unpaired t-test and with Bonferroni's correction for multiple comparisons. (M) Cartoon of extracellular recordings of spontaneous spikes in Purkinje cells. The recording electrode was placed around the axon hillock of Purkinje cells. CNQX (20 µM) and 10 µM gabazine were bath-applied during recording. Right: Representative traces of spontaneous spikes in RIM4 wild-type and KOconst mice (left). Scale bars = 0.1 s (horizontal) and 0.1 mV (vertical). (N) Quantification of Purkinje cell mean firing in RIM4 wild-type and KOconst mice (WT 53.3 ± 0.04 Hz, KOconst 16.1 ± 0.85 Hz). n mice/n cells: WT 5/48; KOconst 6/56 (mice >100 days), unpaired t-test. (O) Representative traces of spikes before and after 1 mM caffeine application (left) in RIM4 wild-type and KOconst mice (left). Time course graph showing Purkinje cell firing rate before, during and after caffeine bath application. n mice/n cells: WT 4/20; KOconst 3/17 (mice >100 days), one-way ANOVA test. KO = knockout.
Figure 3

RIM4 plays a pathway-specific role in cerebellar Purkinje cells and appears dispensable in hippocampus and cerebral cortex. (A) Representative images of sagittal brain sections stained with fluorescent Nissl green from RIM4 wild-type (WT) (left) and KOconst (right) mice. White lines delineate the measured areas, cortex (CX), cerebellum (CB) and hippocampus (HC). Scale bars = 1 mm. (B) Bar graph showing the quantitative analysis of the area of the analysed brain regions, CX (WT 12.46 ± 0.12 mm2, KOconst 11.99 ± 0.39 mm2), CB (WT 7.78 ± 0.36 mm2, KOconst 6.47 ± 0.13 mm2) and HC (WT 1.43 ± 0.05 mm2, KOconst 1.43 ± 0.05 mm2), normalized to the WT value. n = 4 mice per group (12–15 weeks), unpaired t-test. (C) Example images of calbindin (magenta) and Hoechst (blue) stained cerebellar sections from RIM4 wild-type and KOconst mice. White lines indicate measured subcerebellar areas, molecular layer (ML), granule cell layer (GCL), white matter (WM). Scale bars = 1 mm. (D) Bar graph showing the quantitative analysis of the area of the analysed subcerebellar regions, molecular layer (ML) (WT 4.04 ± 0.25 mm2, KOconst 2.99 ± 0.15 mm2), granule cell layer (GCL) (WT 3.22 ± 0.22 mm2, KOconst 2.59 ± 0.1 mm2), white matter (WM) (WT 0.95 ± 0.13 mm2, KOconst 0.88 ± 0.07 mm2), normalized to the wild-type value. n = 4 mice per group (12–15 weeks), unpaired t-test. (EH) ML thickness (WT 173.05 ± 4.36 μm, KOconst 151.08 ± 2.83 μm), total Purkinje cell layer (PCL) length, and Purkinje cell density (WT 4.18 ± 0.13 / 100 μm, KOconst 4.61 ± 0.14 / 100 μm) were measured in calbindin-stained cerebellar tissue from RIM4 wild-type and KOconst mice. The estimated number of Purkinje cells was calculated by multiplying the total PCL length by the Purkinje cell density. n = 4 mice per group (12–15 weeks), unpaired t-test. (I) Representative confocal image of Purkinje cells labelled with the red fluorescent dye TMR-Dextran (1 mM) in RIM4 wild-type and KOconst mice. Maximum intensity projections (top) and 3D reconstructed dendritic trees (right) with branch level annotations (fourth = blue; fifth = yellow; sixth = red). Scale bars = 50 µm. (J) Bar graph showing the quantification of the dendritic tree area in RIM4 wild-type and KOconst mice (WT 1.92 ± 0.12 × 10−2 mm2, KOconst 1.38 ± 0.08 × 10−2 mm2). n mice/n cells: WT 11/22, KOconst 10/30 (>100 days), unpaired t-test. (K) Graph showing Sholl intersections measured from 3D reconstructed Purkinje cells plotted against the radial distance from the cell body in RIM4 wild-type and KOconst mice. WT n = 11, KOconstn = 10 mice (>100 days), Kolmogorov-Smirnov test. (L) Bar graph showing quantification of dendritic length for each branching level (branch level 4: WT 1.6 ± 0.1 mm, KOconst 1.1 ± 0.04 mm; branch level 5: WT 1.4 ± 0.1 mm, KOconst 0.7 ± 0.04 mm; branch level 6: WT 0.9 ± 0.1 mm, KOconst 0.4 ± 0.04 mm). WT n = 11 KOconstn = 10 mice (>100 days), unpaired t-test and with Bonferroni's correction for multiple comparisons. (M) Cartoon of extracellular recordings of spontaneous spikes in Purkinje cells. The recording electrode was placed around the axon hillock of Purkinje cells. CNQX (20 µM) and 10 µM gabazine were bath-applied during recording. Right: Representative traces of spontaneous spikes in RIM4 wild-type and KOconst mice (left). Scale bars = 0.1 s (horizontal) and 0.1 mV (vertical). (N) Quantification of Purkinje cell mean firing in RIM4 wild-type and KOconst mice (WT 53.3 ± 0.04 Hz, KOconst 16.1 ± 0.85 Hz). n mice/n cells: WT 5/48; KOconst 6/56 (mice >100 days), unpaired t-test. (O) Representative traces of spikes before and after 1 mM caffeine application (left) in RIM4 wild-type and KOconst mice (left). Time course graph showing Purkinje cell firing rate before, during and after caffeine bath application. n mice/n cells: WT 4/20; KOconst 3/17 (mice >100 days), one-way ANOVA test. KO = knockout.

Spontaneous firing of Purkinje cells is important for motor performance23,24 and is maintained in Purkinje cells in acute slices. We pharmacologically blocked both excitatory and inhibitory inputs and performed juxtacellular recordings of spontaneous activity of Purkinje cells (Fig. 3M). The mean firing frequency was strongly reduced in RIM4 KOconst compared to wild-type mice (Fig. 3N). Furthermore, application of caffeine, which induces episodes of dystonia in vivo (Fig. 1G), inhibited spontaneous firing of Purkinje cells in RIM4 KOconst mice but not in wild-type mice (Fig. 3O).

Altered synaptic integration by Purkinje cells

The firing activity of Purkinje cells in vivo is strongly modulated by synaptic input.25 Climbing fibre (CF) synaptic input from neurons in the inferior olive (IO) generates pronounced dendritic calcium signals,26 which have previously been studied in great detail in vivo.27-29 We performed GCaMP6f-based in vivo calcium imaging of Purkinje cells in the lobules V or VI of the cerebellar vermis in anaesthetized and head-fixed RIM4 KOconst and wild-type mice to examine synaptic activation of Purkinje cells (Fig. 4A). Wild-type mice displayed spontaneous calcium signals across the lateral extent of dendrites of Purkinje cells in vivo similar in amplitude, kinetics and frequency, as previously reported30,31 (Fig. 4B and Supplementary Fig. 4A and B). While calcium signals occurred with similar kinetics in RIM4 KOconst mice (rise time 20–80%, WT 55.9 ± 1.4 ms, KOconst 54.7 ± 3.5 ms; decay time, WT 288.5 ± 17.2 ms, KOconst 313 ± 23.1 ms, Supplementary Fig. 4A and B), their amplitudes were substantially reduced (Fig. 4C and D). Calcium signal amplitude distributions broadly followed Poisson distributions (dashed lines) with smaller values and reduced amplitude variability in RIM4 KOconst mice (Fig. 4D). Calcium signals occurred at a similar mean rate in wild-type and RIM4 KOconst mice (Fig. 4E), but their temporal pattern strikingly differed. The intervals between two consecutive calcium signals in Purkinje cells of wild-type mice showed three characteristic values (peaks in the histogram, bullet points in Fig. 4F) at ∼0.3, 2 and 5 s that clearly deviated from a random distribution. These characteristics were not observed in RIM4 KOconst mice, and the distribution of intervals resembled that of a random process (dashed line, Fig. 4F), suggesting that the rhythms of synaptic signals generated in the olivo-cerebellar loop and arriving at Purkinje cells are severely diminished. Next, we compared the temporal correlation of calcium signal activity between neighbouring Purkinje cells by calculating the Pearson's correlation coefficient of calcium signal event times. As shown in Fig. 4G, calcium activity was highly correlated for a group of Purkinje cells within a neighbourhood of approximately ±20 µm in wild-type mice. The degree of synchronicity of neighbouring Purkinje cells in RIM4 KOconst mice was greatly reduced, and the region of increased synchronicity was narrowed to approximately ±10 µm, indicating a pronounced alteration in the function of Purkinje cell ensembles (Fig. 4H).

Reduced amplitude and synchronicity of calcium spikes in Purkinje cells of RIM4 KOconst. (A) For in vivo two-photon calcium imaging, AAV particles encoding GCaMP6f were injected into the cerebellar vermis (lobules V and VI) and the cranial windows were positioned over the left lateral cerebellum (left). Fluorescence image showing selective expression of GCaMP6f (green) and Hoechst staining (blue) in Purkinje cells (right). Scale bar = 1 mm. (B) Representative two-photon non-descanned scans showing GCaMP6f expressing dendrites of Purkinje cells. Different coloured lines delineate the dendritic contours of individual Purkinje cells (PCs) and the corresponding example ΔF/F traces for RIM4 wild-type (WT, left) and KOconst (knockout, right). Line scan is indicated by dashed white line. Scale bars = 20 µm in scans, 10 s (horizontal) and 0.4 ΔF/F (vertical) in traces. (C) Quantitative analysis of the mean amplitude of spontaneous calcium signals in RIM4 wild-type and KOconst mice (WT 0.14 ± 0.02 ΔF/F, KOconst 0.06 ± 0.012 ΔF/F). n = 5 mice per group, unpaired t-test. (D) Histogram showing the data from (C) fitted with a Poisson distribution (dashed lines) to compare estimates of the numbers of inferior olive (IO) axon action potentials underlying the calcium signals on average. See the ‘Discussion’ section for details. Data were fitted with a modified Poisson function to match DF/F data: scale e−lambdalambdak / k! x­­–data were transformed onto natural numbers. We obtained lambda = 6.4 and scale = 61.4 for wild-type and lambda = 3.9 and scale = 87.4 for KO mice. (E) Bar graph showing the mean event rate of spontaneous calcium signals in RIM4 wild-type and KOconst mice (WT 0.19 ± 0.03 Hz, KOconst 0.17 ± 0.02 Hz). n = 5 mice per group, unpaired t-test. (F) Histogram showing the distribution of calcium event intervals in RIM4 wild-type and KOconst mice. Black dots indicate distinct peaks in the wild-type population. The KOconst data were fitted with an exponential distribution (grey dashed line, left). Representative ΔF/F traces of calcium spikes (marked in red). Scale bars = 5 s (horizontal), 0.1 ΔF/F (vertical). (G and H) Plots of the synchronicity of spontaneous calcium spikes in neighbouring Purkinje cells as a function of the mediolateral distance between cells for RIM4 wild-type and KOconst mice. The colour scale indicates the frequency of measured data-points (0 white to 1.0 dark blue). The mean synchronicity (black circles connected by a black line) was measured in each bin and the KOconst values were corrected according to the interdendritic interval (see the ‘Materials and methods’ section), two-way ANOVA test. Right: Raster plot of the temporal distribution of calcium events in RIM4 wild-type and KOconst. Black bars indicate individual calcium events in each dendrite and the temporal summation of events across neighbouring dendrites is colour-coded below (from 0 dark blue to 9 spikes bright yellow). Cross-correlation coefficient values are colour-coded and shown for each dendrite (0 white to 1 black). (I) For in vitro calcium imaging, Purkinje cells (PCs) were current-clamped with a recording electrode filled with the calcium indicator OGB-1 (200 μM) and imaged after climbing fibre (CF) stimulation. (J) Representative ΔF/F images showing CF-evoked postsynaptic calcium transients in Purkinje cell dendrites (J, ΔF at peak colour coded, from 0, black, to 1.5 × 104, light green) superimposed on averaged baseline scans (F, white). (K) Representative traces of calcium transients after CF-stimulation (left) and of currents from somatic current-clamp recordings (cells brought to −70 mV to suppress spiking and improve comparability) showing complex spikes (right) from RIM4 wild-type (black) and KOconst (grey) mice. Scale bars = calcium transients 1 s (horizontal), 0.1 ΔF/F (vertical), currents 5 ms (horizontal), 20 mV (vertical). (L) Bar graph showing the mean amplitude of CF stimulation-induced calcium transients in proximal (Prox.) and distal (Dist.) dendrites and in the whole dendritic tree (Tree) of RIM4 wild-type and KOconst mice (4–5 weeks) (WT: proximal 0.13 ± 0.01 ΔF/F, distal 0.34 ± 0.04 ΔF/F, whole tree 0.25 ± 0.03 ΔF/F; KOconst: proximal 0.16 ± 0.02 ΔF/F, distal 0.43 ± 0.06 ΔF/F, whole tree 0.34 ± 0.04 ΔF/F). n mice/n cells, WT 3/5; KOconst 5/8, unpaired t-test. (M) Individual (grey) and mean (black) traces of calcium transients after multiple CF stimuli (stimulus indicated by red box) and corresponding colour-coded profile of ΔF/F values (0 dark blue to 0.6 yellow) for RIM4 wild-type and KOconst mice. Scale bars = 0.5 s (horizontal), 0.1 ΔF/F (vertical). (N) Quantitative analysis of the relative amplitude in response to the first, second and third stimulus in RIM4 wild-type and KOconst mice (4–5 weeks) (WT, 2nd 0.69 ± 0.02, 3rd 0.75 ± 0.02; KOconst, 2nd 0.54 ± 0.02, 3rd 0.56 ± 0.02). n mice/n cells, WT 3/6; KOconstn = 5/10, unpaired t-test. (O) Evoked electrical response in Purkinje cells were recorded in voltage clamp mode (−70 mV holding potential) after climbing fibre stimulation (inter stimulation interval 50 ms). Middle: Mean excitatory postsynaptic potential (EPSC) traces from PCs in response to the first and second CF stimulus. Right: Quantification of the mean EPSC amplitude and paired-pulse ratio. n mice/n cells, WT 8/5; KOconst 5/9, unpaired t-test. Scale bars = 20 ms (horizontal), 0.2 nA (vertical).
Figure 4

Reduced amplitude and synchronicity of calcium spikes in Purkinje cells of RIM4 KOconst. (A) For in vivo two-photon calcium imaging, AAV particles encoding GCaMP6f were injected into the cerebellar vermis (lobules V and VI) and the cranial windows were positioned over the left lateral cerebellum (left). Fluorescence image showing selective expression of GCaMP6f (green) and Hoechst staining (blue) in Purkinje cells (right). Scale bar = 1 mm. (B) Representative two-photon non-descanned scans showing GCaMP6f expressing dendrites of Purkinje cells. Different coloured lines delineate the dendritic contours of individual Purkinje cells (PCs) and the corresponding example ΔF/F traces for RIM4 wild-type (WT, left) and KOconst (knockout, right). Line scan is indicated by dashed white line. Scale bars = 20 µm in scans, 10 s (horizontal) and 0.4 ΔF/F (vertical) in traces. (C) Quantitative analysis of the mean amplitude of spontaneous calcium signals in RIM4 wild-type and KOconst mice (WT 0.14 ± 0.02 ΔF/F, KOconst 0.06 ± 0.012 ΔF/F). n = 5 mice per group, unpaired t-test. (D) Histogram showing the data from (C) fitted with a Poisson distribution (dashed lines) to compare estimates of the numbers of inferior olive (IO) axon action potentials underlying the calcium signals on average. See the ‘Discussion’ section for details. Data were fitted with a modified Poisson function to match DF/F data: scale elambdalambdak / k! x­­–data were transformed onto natural numbers. We obtained lambda = 6.4 and scale = 61.4 for wild-type and lambda = 3.9 and scale = 87.4 for KO mice. (E) Bar graph showing the mean event rate of spontaneous calcium signals in RIM4 wild-type and KOconst mice (WT 0.19 ± 0.03 Hz, KOconst 0.17 ± 0.02 Hz). n = 5 mice per group, unpaired t-test. (F) Histogram showing the distribution of calcium event intervals in RIM4 wild-type and KOconst mice. Black dots indicate distinct peaks in the wild-type population. The KOconst data were fitted with an exponential distribution (grey dashed line, left). Representative ΔF/F traces of calcium spikes (marked in red). Scale bars = 5 s (horizontal), 0.1 ΔF/F (vertical). (G and H) Plots of the synchronicity of spontaneous calcium spikes in neighbouring Purkinje cells as a function of the mediolateral distance between cells for RIM4 wild-type and KOconst mice. The colour scale indicates the frequency of measured data-points (0 white to 1.0 dark blue). The mean synchronicity (black circles connected by a black line) was measured in each bin and the KOconst values were corrected according to the interdendritic interval (see the ‘Materials and methods’ section), two-way ANOVA test. Right: Raster plot of the temporal distribution of calcium events in RIM4 wild-type and KOconst. Black bars indicate individual calcium events in each dendrite and the temporal summation of events across neighbouring dendrites is colour-coded below (from 0 dark blue to 9 spikes bright yellow). Cross-correlation coefficient values are colour-coded and shown for each dendrite (0 white to 1 black). (I) For in vitro calcium imaging, Purkinje cells (PCs) were current-clamped with a recording electrode filled with the calcium indicator OGB-1 (200 μM) and imaged after climbing fibre (CF) stimulation. (J) Representative ΔF/F images showing CF-evoked postsynaptic calcium transients in Purkinje cell dendrites (J, ΔF at peak colour coded, from 0, black, to 1.5 × 104, light green) superimposed on averaged baseline scans (F, white). (K) Representative traces of calcium transients after CF-stimulation (left) and of currents from somatic current-clamp recordings (cells brought to −70 mV to suppress spiking and improve comparability) showing complex spikes (right) from RIM4 wild-type (black) and KOconst (grey) mice. Scale bars = calcium transients 1 s (horizontal), 0.1 ΔF/F (vertical), currents 5 ms (horizontal), 20 mV (vertical). (L) Bar graph showing the mean amplitude of CF stimulation-induced calcium transients in proximal (Prox.) and distal (Dist.) dendrites and in the whole dendritic tree (Tree) of RIM4 wild-type and KOconst mice (4–5 weeks) (WT: proximal 0.13 ± 0.01 ΔF/F, distal 0.34 ± 0.04 ΔF/F, whole tree 0.25 ± 0.03 ΔF/F; KOconst: proximal 0.16 ± 0.02 ΔF/F, distal 0.43 ± 0.06 ΔF/F, whole tree 0.34 ± 0.04 ΔF/F). n mice/n cells, WT 3/5; KOconst 5/8, unpaired t-test. (M) Individual (grey) and mean (black) traces of calcium transients after multiple CF stimuli (stimulus indicated by red box) and corresponding colour-coded profile of ΔF/F values (0 dark blue to 0.6 yellow) for RIM4 wild-type and KOconst mice. Scale bars = 0.5 s (horizontal), 0.1 ΔF/F (vertical). (N) Quantitative analysis of the relative amplitude in response to the first, second and third stimulus in RIM4 wild-type and KOconst mice (4–5 weeks) (WT, 2nd 0.69 ± 0.02, 3rd 0.75 ± 0.02; KOconst, 2nd 0.54 ± 0.02, 3rd 0.56 ± 0.02). n mice/n cells, WT 3/6; KOconstn = 5/10, unpaired t-test. (O) Evoked electrical response in Purkinje cells were recorded in voltage clamp mode (−70 mV holding potential) after climbing fibre stimulation (inter stimulation interval 50 ms). Middle: Mean excitatory postsynaptic potential (EPSC) traces from PCs in response to the first and second CF stimulus. Right: Quantification of the mean EPSC amplitude and paired-pulse ratio. n mice/n cells, WT 8/5; KOconst 5/9, unpaired t-test. Scale bars = 20 ms (horizontal), 0.2 nA (vertical).

The reduced amplitude of dendritic calcium signals in Purkinje cells in vivo prompted us to compare CF signalling in acute slices (confocal imaging). To this end, we examined calcium entry in patch-clamped Purkinje cells in current-clamp mode using the calcium indicator OGB-1 (200 µM) and confocal microscopy (Fig. 4I and J). CF stimulation induced clear complex spikes in Purkinje cells from wild-type and KO mice (Fig. 4K and Supplementary Fig. 4C and D). Unexpectedly, and in contrast to the in vivo recordings, CF stimulation elicited calcium transients in both groups that were similar in amplitude along the dendrites (Fig. 4K and L) as well as in kinetics (Supplementary Fig. 4E and F). However, we found that Purkinje cells in RIM4 KOconst integrated a series of CF stimuli at the most frequently observed interval of ∼0.3 s slightly less effectively than Purkinje cells in wild-type mice (Fig. 4M and N). We further compared the synaptic conductance of CF EPSCs by recording Purkinje cells in voltage-clamp mode during stimulation of a CF (Fig. 4O). CF EPSCs were significantly increased (∼1.6-fold) in Purkinje cells from RIM4 KOconst mice, whereas paired pulse depression (Fig. 4O) and the kinetics of synaptic currents (Supplementary Fig. 4G and H) were not changed.

Ablation of RIM4 in Purkinje cells recapitulates cerebellar pathology and motor dysfunction of the constitutive KO

To examine if ablation of RIM4 only in Purkinje cells is sufficient to recapitulate the spontaneous episodes of dystonia observed in RIM4 KOconst mice, we crossed conditional RIM4fl/fl mice with mice expressing Cre recombinase under the Purkinje cell-protein 2 (PCP2) promoter,32 resulting in a Purkinje cell-specific RIM4 knock-out (KOPCP2). We verified the specificity of RIM4 ablation in KOPCP2 mice by performing fluorescent in situ hybridization (RNAscope) experiments, which revealed the selective reduction of RIM4 mRNA in cerebellar Purkinje cells without changes in other principal neurons in the cortex and hippocampus, as well as other cell types such as granule cells (GCs) and molecular layer interneurons (MLIs) in the cerebellar cortex (Fig. 5A and B). As observed in the RIM4 KOconst line, RIM4 KOPCP2 mice exhibited a similarly reduced body weight compared to their wild-type littermates (Fig. 5C). Video monitoring revealed that RIM4 KOPCP2 mice also exhibited spontaneous dyskinetic-dystonic episodes (Fig. 5D), which were indistinguishable from that of RIM4 KOconst mice. Furthermore, as observed in RIM4 KOconst mice, caffeine administration induced dyskinetic-dystonic episodes in all RIM4 KOPCP2 mice tested, whereas ethanol was slightly less effective, inducing episodes in only 3/6 mice compared to 9/10 in RIM4 KOconst mice (Fig. 5E). Finally, RIM4 KOPCP2 mice also reproduced the poor baseline (non-episode time) performance of RIM4 KOconst mice on the accelerating rotarod (Fig. 5F) and the complex running wheel (Fig. 5G and H).

Specific deletion of RIM4 only in Purkinje cells reproduces motor dysfunction. (A) Representative fluorescent in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) and Nissl counterstain (blue) in cerebellar Purkinje cells, cortex and hippocampus in RIM4 wild-type (WT) and KOPCP2 mice. Scale bars in scans = 20 µm. The dashed white lines indicate the border between the granule cells (GC) and the Purkinje cells (PCs). (B) Quantification of RIM4 mRNA density in the three cerebellar subregions and pyramidal neurons in the cortex and hippocampus in RIM4 wild-type and KOPCP2 mice. Purkinje cells, granule cells, molecular layer interneurons (MLI), pyramidal neurons (PN), n = 4 mice per group, unpaired t-test. (C) Graph showing the development of body weight during the first 35 days after birth in RIM4 wild-type and KOPCP2 mice. n = 5 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (D) Onset and duration of episodes of motor impairment in RIM4 KOPCP2 during 5 days of continuous video monitoring. WT n = 5, KOPCP2n = 7 mice. (E) Diagram showing the susceptibility of RIM4 wild-type and KOPCP2 mice to the induction of an episode of motor impairment by intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse), or caffeine (Caff, 25 mg/kg mouse). WT n = 4–8, KOPCP2n = 6–10 mice. (F) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 wild-type and KOPCP2 mice (11–15 weeks), WT n = 10, KOPCP2n = 8 mice, unpaired t-test. (G) RIM4 wild-type and KOPCP2 mice (13–17 weeks) were placed in a cage with a running wheel and monitored for four consecutive days, regular wheel (RW) consisting of all 38 rungs for 2 days and complex wheel (CX) consisting of 22 rungs (16 rungs were randomly selected and removed) for 2 days. Black boxes indicate time spent running in the wheel, black lines represent episodes of motor impairment. WT n = 8 RIM4, KOPCP2n = 5 mice. (H) Graph of the average instantaneous running speed of the mice shown in G. One-way ANOVA tests with Sidak's post hoc correction. (I) Representative images of a sagittal brain slices stained with fluorescent Nissl green from RIM4 wild-type (left) and KOPCP2 (right) mice. White lines delineate the measured areas, cortex (CX), cerebellum (CB) and hippocampus (HC) (left, scale bars = 1 mm). Bar graph showing the quantitative analysis of the area of the analysed brain regions, CX (WT 1 ± 0.03, KOPCP2 0.93 ± 0.037), CB (WT 1 ± 0.02, KOPCP2 0.78 ± 0.037) and HC (WT 1 ± 0.078, KOPCP2 1.05 ± 0.044, areas normalized to mean of WT). n = 5 mice per group, unpaired t-test. (J) Representative images of cerebellar sections (vermis VI) stained with calbindin (magenta) and VGlut2 (green) in RIM4 wild-type and KOPCP2 mice. Scale bars = 50 µm. Line profiles of average VGlut2 intensity were measured from the Purkinje cell soma to the top of the molecular layer. (K) VGlut2 fluorescence intensity was plotted against the distance from the Purkinje cell soma for RIM4 WT (black) and KOPCP2 mice (grey). The average length of the molecular layer for both groups is shown below the data (percentiles, 25, 50 and 75, are indicated by ticks on the lines). WT n = 5, KOPCP2n = 6 mice (6–8 weeks), Kolmogorov-Smirnov test. (L) Juxtacellular recordings of Purkinje cells were performed in the presence of blockers for excitatory (20 μM CNQX) and inhibitory (10 μM gabazine) postsynaptic currents. Left: Quantitative analysis of the mean spontaneous firing rate during baseline recordings, unpaired t-test. Middle: Example traces of Purkinje cell firing at baseline and after 1 mM caffeine application in RIM4 wild-type (black) and KOPCP2 (grey) mice. Red boxes above the traces indicate the onset of individual spikes. Right: Time course of the firing rate before, during and after caffeine application. n mice/n cells, WT 20/297, KOPCP2 15/212 for spontaneous spike analysis and WT 5/20, KOPCP2 3/17 cells for caffeine application. (M) Average spontaneous action potentials recorded in whole cell current clamp mode at resting potential. (N) Comparison of intrinsic excitability properties obtained from whole cell current clamp recordings (n cells, WT 9, KOPCP2 10, except for input resistance: 6/6). (O) Deep cerebellar nuclei (DCN) neurons were filled with the red fluorescent dye TMR-dextran (200 µM) as shown in the fluorescence image. Purkinje cell axons were stimulated in the white matter. Example traces of inhibitory postsynaptic currents in Purkinje cells of RIM4 wild-type (black) and KOPCP2 (grey) mice, stimulus indicated by a red box. Quantitative analysis of the paired-pulse ratio (PPR) (WT 0.98 ± 0.03, KOPCP2 1.00 ± 0.02). Scale bars = 100 µm image, 20 ms (horizontal) and 0.1 nA (vertical) traces. n mice/n cells, WT 11/18, KOPCP2 6/13, unpaired t-test. (P) Parallel fibres (PFs) were electrically stimulated with inter-event intervals of 50 ms. Representative excitatory postsynaptic potential (EPSC) traces at PF-PC from wild-type (solid black line) and knockout (KO, dashed line) mice. Red boxes indicate stimulus. Bar graphs showing PPR for both synapse types in WT (black) and KO (white). n = mice/n = cells, WT 5/9, KOPCP2 3/6 at PF-PC synapses; WT 3/8, unpaired t-test.
Figure 5

Specific deletion of RIM4 only in Purkinje cells reproduces motor dysfunction. (A) Representative fluorescent in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) and Nissl counterstain (blue) in cerebellar Purkinje cells, cortex and hippocampus in RIM4 wild-type (WT) and KOPCP2 mice. Scale bars in scans = 20 µm. The dashed white lines indicate the border between the granule cells (GC) and the Purkinje cells (PCs). (B) Quantification of RIM4 mRNA density in the three cerebellar subregions and pyramidal neurons in the cortex and hippocampus in RIM4 wild-type and KOPCP2 mice. Purkinje cells, granule cells, molecular layer interneurons (MLI), pyramidal neurons (PN), n = 4 mice per group, unpaired t-test. (C) Graph showing the development of body weight during the first 35 days after birth in RIM4 wild-type and KOPCP2 mice. n = 5 mice per group, two-way ANOVA test with Bonferroni post hoc correction. (D) Onset and duration of episodes of motor impairment in RIM4 KOPCP2 during 5 days of continuous video monitoring. WT n = 5, KOPCP2n = 7 mice. (E) Diagram showing the susceptibility of RIM4 wild-type and KOPCP2 mice to the induction of an episode of motor impairment by intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse), or caffeine (Caff, 25 mg/kg mouse). WT n = 4–8, KOPCP2n = 6–10 mice. (F) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 wild-type and KOPCP2 mice (11–15 weeks), WT n = 10, KOPCP2n = 8 mice, unpaired t-test. (G) RIM4 wild-type and KOPCP2 mice (13–17 weeks) were placed in a cage with a running wheel and monitored for four consecutive days, regular wheel (RW) consisting of all 38 rungs for 2 days and complex wheel (CX) consisting of 22 rungs (16 rungs were randomly selected and removed) for 2 days. Black boxes indicate time spent running in the wheel, black lines represent episodes of motor impairment. WT n = 8 RIM4, KOPCP2n = 5 mice. (H) Graph of the average instantaneous running speed of the mice shown in G. One-way ANOVA tests with Sidak's post hoc correction. (I) Representative images of a sagittal brain slices stained with fluorescent Nissl green from RIM4 wild-type (left) and KOPCP2 (right) mice. White lines delineate the measured areas, cortex (CX), cerebellum (CB) and hippocampus (HC) (left, scale bars = 1 mm). Bar graph showing the quantitative analysis of the area of the analysed brain regions, CX (WT 1 ± 0.03, KOPCP2 0.93 ± 0.037), CB (WT 1 ± 0.02, KOPCP2 0.78 ± 0.037) and HC (WT 1 ± 0.078, KOPCP2 1.05 ± 0.044, areas normalized to mean of WT). n = 5 mice per group, unpaired t-test. (J) Representative images of cerebellar sections (vermis VI) stained with calbindin (magenta) and VGlut2 (green) in RIM4 wild-type and KOPCP2 mice. Scale bars = 50 µm. Line profiles of average VGlut2 intensity were measured from the Purkinje cell soma to the top of the molecular layer. (K) VGlut2 fluorescence intensity was plotted against the distance from the Purkinje cell soma for RIM4 WT (black) and KOPCP2 mice (grey). The average length of the molecular layer for both groups is shown below the data (percentiles, 25, 50 and 75, are indicated by ticks on the lines). WT n = 5, KOPCP2n = 6 mice (6–8 weeks), Kolmogorov-Smirnov test. (L) Juxtacellular recordings of Purkinje cells were performed in the presence of blockers for excitatory (20 μM CNQX) and inhibitory (10 μM gabazine) postsynaptic currents. Left: Quantitative analysis of the mean spontaneous firing rate during baseline recordings, unpaired t-test. Middle: Example traces of Purkinje cell firing at baseline and after 1 mM caffeine application in RIM4 wild-type (black) and KOPCP2 (grey) mice. Red boxes above the traces indicate the onset of individual spikes. Right: Time course of the firing rate before, during and after caffeine application. n mice/n cells, WT 20/297, KOPCP2 15/212 for spontaneous spike analysis and WT 5/20, KOPCP2 3/17 cells for caffeine application. (M) Average spontaneous action potentials recorded in whole cell current clamp mode at resting potential. (N) Comparison of intrinsic excitability properties obtained from whole cell current clamp recordings (n cells, WT 9, KOPCP2 10, except for input resistance: 6/6). (O) Deep cerebellar nuclei (DCN) neurons were filled with the red fluorescent dye TMR-dextran (200 µM) as shown in the fluorescence image. Purkinje cell axons were stimulated in the white matter. Example traces of inhibitory postsynaptic currents in Purkinje cells of RIM4 wild-type (black) and KOPCP2 (grey) mice, stimulus indicated by a red box. Quantitative analysis of the paired-pulse ratio (PPR) (WT 0.98 ± 0.03, KOPCP2 1.00 ± 0.02). Scale bars = 100 µm image, 20 ms (horizontal) and 0.1 nA (vertical) traces. n mice/n cells, WT 11/18, KOPCP2 6/13, unpaired t-test. (P) Parallel fibres (PFs) were electrically stimulated with inter-event intervals of 50 ms. Representative excitatory postsynaptic potential (EPSC) traces at PF-PC from wild-type (solid black line) and knockout (KO, dashed line) mice. Red boxes indicate stimulus. Bar graphs showing PPR for both synapse types in WT (black) and KO (white). n = mice/n = cells, WT 5/9, KOPCP2 3/6 at PF-PC synapses; WT 3/8, unpaired t-test.

Selective absence of RIM4 from Purkinje cells was sufficient to reproduce the morphological changes: we observed a similar reduction in cerebellar size in RIM4 KOPCP2 mice (WT 1 ± 0.02, KOPCP2 0.78 ± 0.037, area normalized to WT mean) (Fig. 5I andSupplementary Fig. 5A–F), while the cortex and hippocampus were unaffected (Fig. 5I). No obvious loss of Purkinje cells was observed (Supplementary Fig. 5G). Selective RIM4 ablation in Purkinje cells altered the distribution of climbing fibre synapses along their dendrites: immunolabelling against the vesicular glutamate transporter 2 (VGlut2) and calbindin revealed a pronouncedly more proximal location of climbing fibre synapses compare to wild-type mice (Fig. 5J and K).

The electrophysiological properties of Purkinje cells in RIM4 KOPCP2 mice also phenocopied those identified in RIM4 KOconst Purkinje cells. Spontaneous activity of Purkinje cells in slices in the presence of both excitatory and inhibitory blockers was significantly reduced in juxtacellular recordings and showed a significantly increased sensitivity to bath application of 1 mM caffeine (Fig. 5L). Current clamp recordings from patch-clamped Purkinje cells in RIM4 KOPCP2 mice revealed broadened action potentials and an increased and delayed fast afterhyperpolarization when compared to RIM4 wild-type mice (Fig. 5M and N), whereas action potential amplitude, resting membrane potential and input resistance were unchanged.

To address if neurotransmitter release from synapses of Purkinje cells is affected by RIM4 ablation, we evaluated GABA release at the Purkinje cell-deep cerebellar nuclei (DCN) synapse (Fig. 5O). We performed voltage clamp recordings from neurons in the DCN. Inhibitory postsynaptic currents were isolated by 20 µM CNQX and 100 µM D-APV and we electrically stimulated Purkinje cell axons in nearby white matter (200–300 µm distance from the recorded soma). Paired-pulse ratios (PPRs) were very similar between wild-type and RIM4 KOPCP2 mice (Fig. 5O, right), suggesting that deletion of RIM4 from Purkinje cells does not affect transmitter release at their presynapses. PPR of parallel fibre (PF)-EPSCs in Purkinje cells of RIM4 KOPCP2 mice, in which RIM4 was absent postsynaptically, was also not altered (Fig. 5P).

We repeated in vivo calcium imaging experiments in wild-type and RIM4 KOPCP2 mice under the same conditions but injected conditional GCaMP6f virus (AAV2/1-pAAV.CAG.Flex.GCaMP6f.WPRE.SV40) to label Purkinje cells specifically (Supplementary Fig. 5H). Spontaneous calcium signals in dendrites of Purkinje cells occurred at a similar frequency as in the constitutive RIM4 WT/KO groups. As observed in the constitutive RIM4 KO line, the amplitudes of calcium signals in Purkinje cells of RIM4 KOPCP2 mice were substantially smaller (WT 0.42 ± 0.06 ΔF/F, KOPCP2 0.12 ± 0.02 ΔF/F) and the spatial synchrony of calcium signals across neighbouring Purkinje cells in a micro-zone was reduced when compared to wild-type mice (Supplementary Fig. 5I–K). These experiments show that the loss of RIM4 from Purkinje cells alone is sufficient to trigger alterations of synaptic integration in cerebellar circuits.

At the ultrastructural level of Purkinje cells in RIM4 KOPCP2 mice, their somata, dendrites and synapses appeared overall normal with no signs of pathology or degeneration (Fig. 6). However, we found PF synapses in RIM4 KOPCP2 mice to occasionally merge to large multi-synapse varicosities contacting multiple spines of Purkinje cells (Fig. 6A and B). Large multi-synapse varicosities with three and more release sites were almost exclusively observed in RIM4 KOPCP2 mice (Fig. 6C). While the size (area) of single synapse boutons in both wild-type and RIM4 KOPCP2 mice was smaller than 0.5 µm2, almost all large multi-synapse boutons showed cross-sections of >0.5 µm2 (Supplementary Fig. 6). Similar structures have previously been observed in the dystonic tottering mouse line,33 suggesting commonalities in the mechanisms of cerebellar pathology. However, large multi-synapse varicosities (at least three synapses) were observed in only 31 of 922 counted PF synapses (Supplementary Fig. 6) of RIM4 KOPCP2 mice, making it difficult to judge their pathophysiological relevance.

Increased frequency of multi-synapse boutons in RIM4 KOPCP2 mice. (A and B) Examples of scanning transmission electron micrographs of the molecular layer of the cerebellum (n = 4 RIM4 WT and n = 4 KOPCP2 mice were analysed for this figure). Asterisks mark Purkinje cell dendrites in the overview images. Arrowheads (orange) mark the release sites in the presynaptic varicosities (orange). Regions for analysis were chosen ∼50 µm distant to the Purkinje cell layer. Varicosities formed on dendrites (and not on spines) were not included in this analysis. (C) Quantitative analysis of abundance of presynaptic varicosities of single and multi-varicosities. n = 4 mice per group, 2500 µm2 region of interest analysed per mouse. Data are presented as sum ± 99% confidence interval of the proportions. Data were statistically compared using the Poisson means test with alpha = 1%, resulting in a significance level of ∼5% when taking into account a Bonferroni-type of correction for three simultaneous comparisons. For RIM4 wild-type (WT) and KOPCP2 mice 243, 234, 215, 220 and 261, 260, 215, 186 synapses were analysed per region of interest, respectively. Proportions were calculated after summing the counts across all mice within each group. (D) Example climbing fibre (CF) synapses identified in scanning transmission electron micrographs. Dendrites are shaded green and were identified by long longitudinal structures typically filled with either filaments (smaller branches, wild-type example) or pronounced smooth endoplasmic reticulum and mitochondria (larger branches, KOPCP2 example). Climbing fibre synapses are shaded blue and in some cases the axon (the CF) connecting individual release sites was visible (right). (E and F) Total counts of parallel fibre (PF) and CF synapse release sites, respectively. As CF synapses occur less frequently, a larger area was counted for this type of synapse. Counting regions of interest (ROIs, each 30 × 30 µm2) were equally distributed across mice and randomly placed in the proximal, medial and distal parts of the molecular layer. Release site counts were summed across mice within the wild-type and the KOPCP2 groups and plotted as sum ± 99% confidence interval per region. The Poisson means test with alpha = 1%, resulting in a significance level of ∼5% when taking into account a Bonferroni-type of correction for three simultaneous comparisons, was used to assess statistical significance. KO = knockout.
Figure 6

Increased frequency of multi-synapse boutons in RIM4 KOPCP2 mice. (A and B) Examples of scanning transmission electron micrographs of the molecular layer of the cerebellum (n = 4 RIM4 WT and n = 4 KOPCP2 mice were analysed for this figure). Asterisks mark Purkinje cell dendrites in the overview images. Arrowheads (orange) mark the release sites in the presynaptic varicosities (orange). Regions for analysis were chosen ∼50 µm distant to the Purkinje cell layer. Varicosities formed on dendrites (and not on spines) were not included in this analysis. (C) Quantitative analysis of abundance of presynaptic varicosities of single and multi-varicosities. n = 4 mice per group, 2500 µm2 region of interest analysed per mouse. Data are presented as sum ± 99% confidence interval of the proportions. Data were statistically compared using the Poisson means test with alpha = 1%, resulting in a significance level of ∼5% when taking into account a Bonferroni-type of correction for three simultaneous comparisons. For RIM4 wild-type (WT) and KOPCP2 mice 243, 234, 215, 220 and 261, 260, 215, 186 synapses were analysed per region of interest, respectively. Proportions were calculated after summing the counts across all mice within each group. (D) Example climbing fibre (CF) synapses identified in scanning transmission electron micrographs. Dendrites are shaded green and were identified by long longitudinal structures typically filled with either filaments (smaller branches, wild-type example) or pronounced smooth endoplasmic reticulum and mitochondria (larger branches, KOPCP2 example). Climbing fibre synapses are shaded blue and in some cases the axon (the CF) connecting individual release sites was visible (right). (E and F) Total counts of parallel fibre (PF) and CF synapse release sites, respectively. As CF synapses occur less frequently, a larger area was counted for this type of synapse. Counting regions of interest (ROIs, each 30 × 30 µm2) were equally distributed across mice and randomly placed in the proximal, medial and distal parts of the molecular layer. Release site counts were summed across mice within the wild-type and the KOPCP2 groups and plotted as sum ± 99% confidence interval per region. The Poisson means test with alpha = 1%, resulting in a significance level of ∼5% when taking into account a Bonferroni-type of correction for three simultaneous comparisons, was used to assess statistical significance. KO = knockout.

In the same samples, we also assessed the density of PF and CF synapses in the proximal (0–30 µm), middle (30–60 µm) and distal (60–90 µm) parts of the molecular layer. In each part of the molecular layer, we counted individual release sites [only asymmetric, glutamatergic, defined by vesicles, narrow synaptic cleft and postsynaptic density (PSD)] and classified them as either PF or CF synapses if they terminated on spines (Fig. 6A and B) or large dendritic shafts (Fig. 6D), respectively. We observed opposite trends in the density of PF and CF synapses (release sites). While the density of PF synapses in the proximal and medial part of the molecular layer was reduced by ∼20% in RIM4 KOPCP2 mice (Fig. 6E), the density of CF synapses in these proximal parts did not differ. Conversely, in the distal part of the molecular layer (60–90 µm) PF synapse density was comparable between groups but there were significantly fewer CF synapses in RIM4 KOPCP2 mice. This is consistent with the idea that the developmental process of territory assignment between CF and PF synapses is mutually linked34 and further supports our confocal data, which also indicate that CF synapses in RIM4 KOPCP2 mice extend less into the distal region of Purkinje cell dendrites (Fig. 5J and K).

RIM4 in excitatory cortical, hippocampal or inferior olive neurons is dispensable for normal motor function

By performing histochemical stainings for the activity of the enzyme β-galactosidase (lacZ) in mice expressing lacZ under the endogenous RIM4 promoter, we observed very strong labelling in IO neurons (Fig. 7A). IO neurons play an important role in motor control by the olivo-cerebellar loop. RIM4 was also found to be expressed, although at substantially lower levels, in hippocampus and cortex. To test for the potential contribution of the IO and the hippocampus/cortex to dyskinesia and dystonia, we also aimed to delete RIM4 specifically from these regions only. We crossed conditional RIM4fl/fl mice with two lines expressing Cre recombinase under the control of either the corticotropin releasing factor (CRF; Fig. 7B)35 or the NeuroD subfamily 6 (NEX; Fig. 7C)36 promoter to generate RIM4 KOCRF and RIM4 KONEX mice, respectively. We verified cell type-specific knockout by quantifying RIM4 mRNA levels in the respective brain regions and cerebellum using fluorescent in situ hybridization (RNAScope) (Fig. 7D and E). In contrast to RIM4 KOconst and KOPCP2 mice, the overall brain size and the size of various subregions were unaltered in both RIM4 KOCRF and KONEX mice (Fig. 7F and G). Five consecutive days of video monitoring revealed that neither of these conditional RIM4 KO lines exhibited spontaneous (Fig. 7H) nor caffeine-induced episodes of dyskinesia/dystonia or any other motor impairments (Fig. 7I). Furthermore, the motor performance of RIM4 KOCRF and RIM4 KONEX mice on an accelerating rotarod (Fig. 7J) and on a simple or complex running wheel (data not shown) was comparable to that of their wild-type littermates. These results strongly suggest a highly Purkinje cell-specific role of RIM4.

RIM4 plays a pathway specific role in cerebellar Purkinje cells that is distinct from the function of the long RIM isoforms. (A) β-Galactosidase staining of brain sections from RIM4 knockout (KO) first mice, in which LacZ is expressed under the control of the endogenous RIM4 promoter. Cerebellum (CB), inferior olive (IO), cortex (CX) and hippocampus (HC). Scale bar, sagittal section = 1 mm, CB and IO = 50 mm. n = 5 mice per group. (B and C) Cartoons showing the major regions in which RIM4 was ablated in RIM4 KOCRF (IO neurons) and RIM4 KONEX (excitatory neurons in the cortex and hippocampus). (D) Representative in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) and Nissl counterstain (red) in the CB and IO of RIM4 wild-type (WT) and KOCRF mice. Scale bar = 20 µm. Right: Quantification of RIM4 mRNA density in the CB and IO of wild-type and RIM4 KOCRF mice. n = 4 mice per group, unpaired t-test. (E) Representative in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) in the and Nissl counterstain (red) CB and CX of wild-type and RIM4 KONEX mice. Scale bar = 20 µm. Right: Quantification of RIM4 mRNA density in the CB and IO of RIM4 wild-type and KOCRF mice. n = 4 mice per group, unpaired t-test. (F) Representative images of brain sections stained with fluorescent Nissl from RIM4 wild-type and KOCRF mice (left). Bar graph showing the quantitative analysis of the size of the analysed brain regions, CX, CB and HC, normalized to the wild-type value. n = 3 mice per group, unpaired t-test. (G) Representative images of brain sections stained with fluorescent Nissl from RIM4 wild-type and KONEX mice (left). Bar graph showing the quantitative analysis of the size of the analysed brain regions, CX, CB and HC, normalized to the wild-type value. n = 3 mice per group, unpaired t-test. (H) Raster plots showing the onset and duration of spontaneously occurring episodes of motor impairment (solid black line) during a continuous 5-day observation period. RIM4 KOconstn = 10, KOPCP2n = 6, KOCRF, KONEX and RIM1/2 KOPCP2n = 3. (I) Diagram showing the susceptibility of RIM4 wild-type and KOconst, KOPCP2, KOCRF, KONEX and RIM1/2 KOPCP2 to the induction of an episode of motor impairment by intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse), or caffeine (Caff, 25 mg/kg mouse). WT n = 3–10, KO n = 5–10 mice. (J) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 wild-type and KOconst, KOPCP2, KOCRF, KONEX and RIM1/2 KOPCP2 mice. RIM4 WT n = 6, KOCRFn = 9, RIM4 WT n = 4, KONEXn = 3 and RIM1/2 WT n = 10 and KOPCP2n = 10, unpaired t-test. (K) RIM1/2 conditional mice were crossed with the Purkinje cell-specific PCP2 Cre driver line. Juxtacellular recordings of Purkinje cells were performed in the presence of blockers for excitatory (20 μM CNQX) and inhibitory (10 μM gabazine) postsynaptic currents. Example traces of Purkinje cell firing in RIM1/2 WT (black) and KOPCP2 (grey) mice. Right: Quantitative analysis of the mean spontaneous firing rate during baseline recordings. WT n = 3 mice, 56 cells, RIM1/2 KOPCP2n = 3 mice, 55 cells. Bar graph, unpaired t-test. GCL = granule cell layer; ML = molecular layer; PCL = Purkinje cell layer.
Figure 7

RIM4 plays a pathway specific role in cerebellar Purkinje cells that is distinct from the function of the long RIM isoforms. (A) β-Galactosidase staining of brain sections from RIM4 knockout (KO) first mice, in which LacZ is expressed under the control of the endogenous RIM4 promoter. Cerebellum (CB), inferior olive (IO), cortex (CX) and hippocampus (HC). Scale bar, sagittal section = 1 mm, CB and IO = 50 mm. n = 5 mice per group. (B and C) Cartoons showing the major regions in which RIM4 was ablated in RIM4 KOCRF (IO neurons) and RIM4 KONEX (excitatory neurons in the cortex and hippocampus). (D) Representative in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) and Nissl counterstain (red) in the CB and IO of RIM4 wild-type (WT) and KOCRF mice. Scale bar = 20 µm. Right: Quantification of RIM4 mRNA density in the CB and IO of wild-type and RIM4 KOCRF mice. n = 4 mice per group, unpaired t-test. (E) Representative in situ hybridization (RNAScope) images showing the density of RIM4 mRNA (green) in the and Nissl counterstain (red) CB and CX of wild-type and RIM4 KONEX mice. Scale bar = 20 µm. Right: Quantification of RIM4 mRNA density in the CB and IO of RIM4 wild-type and KOCRF mice. n = 4 mice per group, unpaired t-test. (F) Representative images of brain sections stained with fluorescent Nissl from RIM4 wild-type and KOCRF mice (left). Bar graph showing the quantitative analysis of the size of the analysed brain regions, CX, CB and HC, normalized to the wild-type value. n = 3 mice per group, unpaired t-test. (G) Representative images of brain sections stained with fluorescent Nissl from RIM4 wild-type and KONEX mice (left). Bar graph showing the quantitative analysis of the size of the analysed brain regions, CX, CB and HC, normalized to the wild-type value. n = 3 mice per group, unpaired t-test. (H) Raster plots showing the onset and duration of spontaneously occurring episodes of motor impairment (solid black line) during a continuous 5-day observation period. RIM4 KOconstn = 10, KOPCP2n = 6, KOCRF, KONEX and RIM1/2 KOPCP2n = 3. (I) Diagram showing the susceptibility of RIM4 wild-type and KOconst, KOPCP2, KOCRF, KONEX and RIM1/2 KOPCP2 to the induction of an episode of motor impairment by intraperitoneal administration of saline (NaCl), ethanol (EtOH, 2 g/kg mouse), or caffeine (Caff, 25 mg/kg mouse). WT n = 3–10, KO n = 5–10 mice. (J) Bar graph showing the latency to fall off an accelerating rotarod (4 to 40 rpm in 300 s) for RIM4 wild-type and KOconst, KOPCP2, KOCRF, KONEX and RIM1/2 KOPCP2 mice. RIM4 WT n = 6, KOCRFn = 9, RIM4 WT n = 4, KONEXn = 3 and RIM1/2 WT n = 10 and KOPCP2n = 10, unpaired t-test. (K) RIM1/2 conditional mice were crossed with the Purkinje cell-specific PCP2 Cre driver line. Juxtacellular recordings of Purkinje cells were performed in the presence of blockers for excitatory (20 μM CNQX) and inhibitory (10 μM gabazine) postsynaptic currents. Example traces of Purkinje cell firing in RIM1/2 WT (black) and KOPCP2 (grey) mice. Right: Quantitative analysis of the mean spontaneous firing rate during baseline recordings. WT n = 3 mice, 56 cells, RIM1/2 KOPCP2n = 3 mice, 55 cells. Bar graph, unpaired t-test. GCL = granule cell layer; ML = molecular layer; PCL = Purkinje cell layer.

To explore if RIM1/2 share functions in cerebellar motor circuits with RIM4, we used the same PCP2-Cre driver line that fully reproduced the symptoms of the RIM4 KOconst and deleted RIM1 and RIM2 from Purkinje cells. However, RIM1/2 KOPCP2 mice did not show episodes of motor impairments during video monitoring in the home cage (Fig. 7H), nor could they be induced by caffeine injection (Fig. 7I). In addition, their performance on the accelerating rotarod (Fig. 7J) and their spontaneous firing activity were unaltered (Fig. 7K). Thus, RIM4 plays a unique role in Purkinje cells, which is not replicated by the long RIM family members.

Differential developmental requirement of RIM4

Our RIM4 ablation strategies resulted in a loss of the protein from early development on (constitutive KO) and around postnatal day (P) 6 for PCP2-Cre.32 Thus, it is not clear whether the structural and functional changes observed in adult mice are due to an acute deficiency of RIM4 in adulthood or whether there is an earlier and transient requirement for RIM4 that results in irreversible changes that persist into adulthood. To address this, we delivered RIM4 to Purkinje cells in RIM4-deficient mice by injecting recombinant adeno-associated virus (rAAV) particles expressing a mRFP1-T2A-Flag-RIM4 cassette (Fig. 8A) at either P0 (RIM4P0 based on RIM4 KOconst) or 3 months of age (RIM4Adult based on RIM4 KOPCP2) and examined whether key structural and functional alterations could be reversed. RIM4 wild-type and KO transduced with rAAV-mRFP served as controls (Fig. 8A). Viral delivery of RIM4 at P0 reversed the dendritic tree abnormalities, reduced the mean firing frequency and reverted the increased caffeine sensitivity of Purkinje cell firing observed in RIM4 KOconst mice (Fig. 8B, C, F and G). In contrast, when we virally delivered RIM4 to adult RIM4 KOPCP2 mice, we observed a divergent effect: while dendritic arborization was not restored (Fig. 8D and E), both the mean frequency and the caffeine sensitivity of Purkinje cell firing were restored to levels seen in wild-type mice (see values earlier) (Fig. 8F and G). Delivery of mRFP-RIM4 to wild-type (OEP0, OEAdult) mice did not affect dendrite branching or firing behaviour of Purkinje cells (Supplementary Fig. 7). Taken together, these results suggest that there is a critical window of time during early development, which requires RIM4 for the correct establishment of the dendritic tree of Purkinje cells. The absence of RIM4 during this critical period results in an irreversible alteration of the dendritic branching pattern. In contrast, the functional alterations observed in adult mice result from the acute absence of RIM4 and can be restored by providing RIM4 protein.

Electrophysiological alterations induced by RIM4 deficiency are reversible, but the reduced complexity of the dendritic tree is not. (A) Cartoon of the viral vectors that were injected into the vermis (lobules IV–VI), RIM4 expression (rAAV2/1-CMV-mRFP-T2A-Flag-RIM4) and control (rAAV2/1-CMV-mRFP) (left). Representative fluorescence images of acute sections showing strong expression of mRFP in Purkinje cells (right). Scale bars = 500 μm and 50 μm. (B and D) AAV particles were injected into the cerebellar vermis at either P0 (B) or 12 weeks (Adult, D), scale bar = 30 μm. Experiments were performed 8 weeks after injection for the P0 group (B) and 4–6 weeks after injection for the adult treatment group (D). Control animals wild-type (WT) and RIM4 knockout (KO) were injected with virus expressing mRFP and treatment animals, P0 and Adult, were injected with virus expressing RIM4 and mRFP (treatment). Right: Representative maximum intensity projections of Purkinje cells labelled with Alexa 488 (green) for all experimental groups. (C and E) Bar graphs showing the quantitative analysis of the dendritic tree area for P0 (WT 2.2 ± 0.1 × 104 μm2, KOconst 1.2 ± 0.1 × 104 μm2, RIM4P0 1.8 ± 0.2 × 104 μm2) and adult (WT 2.0 ± 0.1 × 104 μm2, KOconst 1.2 ± 0.1 × 104 μm2, RIM4P0 1.3 ± 0.1 × 104 μm2) treatment groups. P0 treatment, n = 3 mice per group, n = cells, WT n = 10, KOconstn = 14, RIM4P0n = 11; Adult treatment, n mice/n cells, WT 8/23 cells, KOPCP2 10 mice/42, RIM4Adult 6/37, unpaired t-test. (F) Representative traces from recordings of spontaneously firing Purkinje cells from wild-type, RIM4 KOPCP2, P0 treatment and adult treatment. Right: Bar graph showing the mean firing rate for the four groups (WT 33.7 ± 2.4 Hz, KO 25.0 ± 2.4 Hz, RIM4P0 35.0 ± 3.5 Hz, RIM4Adult 37.9 ± 5.3 Hz). n mice/n cells, WT n = 3/10, KOPCP2n= 4/13, RIM4Adultn = 4/5, RIM4Adultn = 6/28, unpaired t-test. Scale bars = 0.1 s (horizontal ) and 0.5 mV (vertical). (G) Time course of PC firing rate of Purkinje cells from the four experimental groups before and during caffeine application (1 mM, 10 min) (left). Bar graph showing the mean firing rate during caffeine application relative to baseline (right) (WT 91.5 ± 1.5%, KO 59.4 ± 12.0%, RIM4P0 102.7 ± 11.5%, RIM4Adult 89.2 ± 2.9%). n mice/n cells, WT 3/9, KOPCP2 4/13, RIM4Adult 3/7, RIM4Adult 6/16, unpaired t-test. Legend of F also applies to G.
Figure 8

Electrophysiological alterations induced by RIM4 deficiency are reversible, but the reduced complexity of the dendritic tree is not. (A) Cartoon of the viral vectors that were injected into the vermis (lobules IV–VI), RIM4 expression (rAAV2/1-CMV-mRFP-T2A-Flag-RIM4) and control (rAAV2/1-CMV-mRFP) (left). Representative fluorescence images of acute sections showing strong expression of mRFP in Purkinje cells (right). Scale bars = 500 μm and 50 μm. (B and D) AAV particles were injected into the cerebellar vermis at either P0 (B) or 12 weeks (Adult, D), scale bar = 30 μm. Experiments were performed 8 weeks after injection for the P0 group (B) and 4–6 weeks after injection for the adult treatment group (D). Control animals wild-type (WT) and RIM4 knockout (KO) were injected with virus expressing mRFP and treatment animals, P0 and Adult, were injected with virus expressing RIM4 and mRFP (treatment). Right: Representative maximum intensity projections of Purkinje cells labelled with Alexa 488 (green) for all experimental groups. (C and E) Bar graphs showing the quantitative analysis of the dendritic tree area for P0 (WT 2.2 ± 0.1 × 104 μm2, KOconst 1.2 ± 0.1 × 104 μm2, RIM4P0 1.8 ± 0.2 × 104 μm2) and adult (WT 2.0 ± 0.1 × 104 μm2, KOconst 1.2 ± 0.1 × 104 μm2, RIM4P0 1.3 ± 0.1 × 104 μm2) treatment groups. P0 treatment, n = 3 mice per group, n = cells, WT n = 10, KOconstn = 14, RIM4P0n = 11; Adult treatment, n mice/n cells, WT 8/23 cells, KOPCP2 10 mice/42, RIM4Adult 6/37, unpaired t-test. (F) Representative traces from recordings of spontaneously firing Purkinje cells from wild-type, RIM4 KOPCP2, P0 treatment and adult treatment. Right: Bar graph showing the mean firing rate for the four groups (WT 33.7 ± 2.4 Hz, KO 25.0 ± 2.4 Hz, RIM4P0 35.0 ± 3.5 Hz, RIM4Adult 37.9 ± 5.3 Hz). n mice/n cells, WT n = 3/10, KOPCP2n= 4/13, RIM4Adultn = 4/5, RIM4Adultn = 6/28, unpaired t-test. Scale bars = 0.1 s (horizontal ) and 0.5 mV (vertical). (G) Time course of PC firing rate of Purkinje cells from the four experimental groups before and during caffeine application (1 mM, 10 min) (left). Bar graph showing the mean firing rate during caffeine application relative to baseline (right) (WT 91.5 ± 1.5%, KO 59.4 ± 12.0%, RIM4P0 102.7 ± 11.5%, RIM4Adult 89.2 ± 2.9%). n mice/n cells, WT 3/9, KOPCP2 4/13, RIM4Adult 3/7, RIM4Adult 6/16, unpaired t-test. Legend of F also applies to G.

Discussion

We have identified an unexpected function for the short RIM family member, RIM4, which differs from the presynaptic organizing role of the long RIMs, RIM1/2, in the CNS. RIM4 does not play a significant presynaptic role, but is crucial for somatodendritic signal integration, and exerts this function specifically in Purkinje cells. Furthermore, the deletion of RIM4 causes a severe motor disorder that is not present in knockout mice of any other RIM family member, highlighting the systemic importance of the distinct cellular function of RIM4.

RIM4 does not play a role in the regulation of synaptic vesicle release

Paired-pulse plasticity of synaptic responses is a robust indicator of transmitter release probability. Mice deficient for the long RIMs (RIM1/2) exhibit changes in paired-pulse plasticity of synaptic responses and a decreased release probability.37-40 We found that paired-pulse plasticity was not altered in the absence of RIM4 at the PC-DCN synapse (RIM4 KOPCP2), the PF-PC, the CF-PC synapses and the Schaffer collateral synapses (both RIM4 KOconst). Furthermore, calcium sensitivity at the PF-PC synapse and the frequency of mEPSCs in the hippocampus were not affected by RIM4 deletion. Taken together, these results strongly suggest that RIM4 has no apparent presynaptic function. This is consistent with the view that the presynaptic role of the long RIMs in organizing voltage-gated calcium channels and synaptic vesicles in the AZ requires their multi-domain structure, whereas RIM4 comprises only a single C2 domain and a short isoform-specific N-terminal sequence.14

Somatodendritic role of RIM4

While we found little evidence for a presynaptic role of RIM4, we observed a number of structural and functional changes suggesting that RIM4 exerts important somatodendritic functions in Purkinje cell excitability. Unlike our previous study, which did not allow us to conclude whether the observed reduced dendritic tree14 might be secondary to a presynaptic effect, our current dataset now provides strong evidence that RIM4 in Purkinje cells is required for proper dendritic development: deletion of RIM4 specifically in Purkinje cells resulted in reduced dendritic branching, and this branching could be restored by viral delivery of RIM4 only to Purkinje cells. Our functional measurements clearly demonstrate a role for RIM4 in the organization of somatic excitability. In the absence of synaptic input (pharmacologically blocked), Purkinje cells displayed decreased firing rates, broadened action potentials and heightened sensitivity to inhibition of firing by caffeine due to the absence of RIM4. Firing deficits were replicated through selective RIM4 ablation in Purkinje cells. Furthermore, by reintroducing RIM4 via viral expression in adult RIM4 KOPCP2 mice, firing rates and caffeine sensitivity were restored to wild-type levels.

RIM4 plays a unique key role in Purkinje cells

Constitutive deletion of RIM4 from all cells, starting early during development, and selective deletion of RIM4 only from Purkinje cells during the second postnatal week resulted in qualitatively and quantitatively the same phenotype. This strongly suggests that RIM4 is of paramount importance for Purkinje cells only from adolescence onward. Our data suggest that the absence of RIM4 throughout development and in other neuron types leads to either no changes or subtle effects. However, it could be argued that constitutive deletion of RIM4 may cause additional deficits that would be masked if RIM4 played similar roles in functionally opposing neurons, such as excitatory and inhibitory neurons. We tested this idea by generating RIM4 KONEX mice in which RIM4 is primarily absent from cortical and hippocampal excitatory neurons, but which show no signs of motor deficits. The finding that deletion of RIM4 from neurons of the IO, which express RIM4 at exceptionally high levels compared to all other neurons in the brain, did not result in any obvious changes in the RIM4 KOCRF mice is unexpected and surprising. It should be noted that many of our assays were primarily focused on analysing motor performance, and it may well be that less severe disturbances caused by RIM4 deficiency in non-Purkinje cells remained undetected.

Postnatal deletion of RIM4 only from Purkinje cells was sufficient to induce (i) baseline motor dysfunction (wheel, rotarod); (ii) spontaneous and caffeine/ethanol induction of dyskinetic/dystonic episodes; (iii) reduced spontaneous firing in Purkinje cells; (iv) enhanced suppression of firing by caffeine; and (v) alterations of in vivo calcium signals in Purkinje cells. These findings strongly suggest that cerebellar Purkinje cells are responsible not only for reduced baseline motor coordination, but also for the induction of dyskinetic/dystonic episodes. In the absence of RIM4 in Purkinje cells, caffeine induces a pronounced weakening of tonic firing of Purkinje cells, leading to unbalanced activity in the cerebellar nucleo-olivary loop via disinhibition of the deep cerebellar nuclei, the target of Purkinje cells. The important and primary role of RIM4 in Purkinje cells in this process is also exemplified by the fact that the firing deficit and increased caffeine responsiveness can be restored by viral delivery of RIM4 to Purkinje cells. A causal role of Purkinje cells in ataxia and the initiation of paroxysmal dystonia is consistent with the finding that deletion of P/Q-type VGCCs only from Purkinje cells also mimics key aspects of tottering mice.41,42

Aberrant synaptic integration in Purkinje cells

The three most prominent features of abnormal signal integration were (i) loss of characteristic frequencies of calcium signals; (ii) substantially reduced amplitude of calcium signals; and (iii) reduced spatial extent of correlation of calcium activity of neighbouring Purkinje cells. In our view, it is unlikely that these changes are mechanistically attributable to alterations intrinsic to Purkinje cells. Rather, as discussed later, a loss of firing rhythms and conversion to random firing of IO neurons could explain the three hallmark changes in calcium signalling in Purkinje cells.

The appearance of spontaneous dendritic calcium signals in Purkinje cells reflects CF activity and thus action potential firing of IO neurons. The loss of the three characteristic frequencies of calcium signals in Purkinje cells of KO mice in vivo (2, 0.5, 0.2 Hz, Fig. 4F) clearly indicates altered firing patterns of IO neurons. In KO mice, the occurrence of calcium signals in Purkinje cells appeared almost random, suggesting that the action potential firing pattern of the presynaptic IO neuron is also random in KO mice.

A disturbed firing pattern of IO neurons may also explain the greatly reduced in vivo calcium (GCaMP) signal amplitude in Purkinje cells of RIM4 KO mice. Such dendritic calcium signals are typically not caused by single action potentials, but rather by bursts of action potentials from IO neurons.43 The more action potentials IO neurons fire per burst, the larger the dendritic calcium transients in Purkinje cells. Because the intervals between action potentials within an IO burst are short (∼5 ms),43 action potentials do not appear as individual calcium transients but accumulate into a relatively slow composite GCaMP signal (rise time ∼50 ms, decay time constant ∼300 ms) (Supplementary Fig. 4A and B). If the accumulation was perfectly linear, the amplitude of the GCaMP signal would directly reflect how many action potentials the IO neuron fired per burst. Random fluctuations in the number of action potentials per burst in IO neurons should follow a Poisson distribution, and so should then, if accumulation is more or less linear, the calcium signal amplitudes in Purkinje cells. Indeed, the distribution of calcium signal amplitudes in both wild-type and RIM4 KO mice could be fit reasonably well by a Poisson distribution. The fit parameters suggested that, on average, IO neurons in RIM4 KO mice fire only ∼60% of action potentials per burst compared to wild-type littermates (lambda was reduced to 60%). Thus, action potentials in IO neurons of RIM4 KO mice do not coincide as much as in wild-type mice and do not show the clustered occurrence of high frequency firing (∼200 Hz)43 such that there are fewer action potentials per bursts. This means that not only the three slow characteristic frequencies that can be measured directly with GCaMP imaging are lost, but also a high frequency component that underlies the formation of longer and more pronounced bursts, supporting the view of an overall loss of coordination of action potential firing towards a random firing pattern of IO neurons.

Neighbouring Purkinje cells form so-called micro-zones, which are defined by commonalities in afferents and in the target areas of their projections. CF activity in Purkinje cells within such micro-zones has been reported to be temporally correlated due to the synchronized action potential activity of IO neurons projecting to that micro-zone.44 In our dataset, we also found this spatiotemporal correlation of dendritic calcium activity amongst a dozen Purkinje cells, most likely reflecting CF activity, within ∼20 µm in wild-tye mice (Fig. 4G and H). We observed this correlation to a significantly lesser degree in RIM4 KO and RIM4 KOPCP2 mice. These findings suggest that not only have individual IO neurons lost their typical firing pattern, but also that IO neurons that belong to a micro-zone of Purkinje cells demonstrate weakened temporal correlation with each other. Taken together, the modifications of in vivo calcium signals seen in Purkinje cells of RIM4 KO and RIM4 KOPCP2 mice can be attributed to altered firing patterns of IO neurons. This could explain why calcium transients recorded in response to experimental electrical stimulation of Purkinje cells in acute brain slices did not appear to differ between the two groups (Fig. 4K and L). Of course, it should be noted that our conclusions are indirect. We inferred alterations in IO neuron firing from changes in calcium signals that we recorded in Purkinje cells without direct recordings from IO neurons.

Deleting RIM4 specifically in IO neurons does not cause spontaneous or induced motor episodes, coordination problems or structural abnormalities of the cerebellum (Fig. 7). On the other hand, deleting RIM4 only in Purkinje cells reproduced the in vivo calcium signalling deficits of Purkinje cells in the RIM4 KOconst mice showing that the presumed alterations of the firing patterns of IO neurons are likely not due to the loss of RIM4 from IO neurons. Firing patterns of IO neurons could be altered because the RIM4 deficiency of Purkinje cells leads to an alteration of some cell-to-cell signalling between Purkinje cells and IO neurons. This would necessitate an unusual, long-range signalling from climbing fibre synapses back to the somata of IO neurons in the brain stem to regulate their excitability. We believe it is more likely that the altered firing pattern of IO neurons rather is a consequence of chronically altered network activity in olivo-cerebellar circuits. Almost every night, mice experience episodes of severe motor impairments and this will be accompanied by aberrant network activity, not only in Purkinje cells. Purkinje cells are connected to IO neurons via the cerebellar nucleo-olivary projection and an adaptation of this tri-synaptic circuit to altered somatodendritic physiology and aberrant activity of Purkinje cells in RIM4 KO mice would not be unexpected.

If the calcium signal abnormalities in Purkinje cells may be secondary to aberrant network activity, the question arises as to which of the other alterations observed in the Purkinje cells of RIM4 KOconst or KOPCP2 mice are directly caused by the absence of RIM4. As explained earlier, the reduced firing frequency and the enhanced caffeine effect are most likely directly due to the loss of RIM4 in Purkinje cells. Furthermore, loss of RIM4 in Purkinje cells also caused an aberrant distribution and density of CF synapses (Figs 5K and 6F). Normally, there is a developmental shift of CF synapses to more distal locations,45 resulting in more widespread activation of Purkinje cell dendrites. As we found that CF synapses preferentially accumulate on proximal dendrites in RIM4 KO mice, it appears that this developmental repositioning as well as the usual segregation of PF synapses (Fig. 6E) do not fully occur in the absence of RIM4. We did not show a reversal of those phenomena by viral delivery of RIM4 and therefore it could be argued that these changes are secondary to the periodic episodes of motor problems. However, it is known that the redistribution of CFs along the dendritic tree is completed around the end of the third postnatal week,46,47 whereas the episodes of dyskinesia and dystonia begin thereafter, with the beginning of the third postnatal week. For this reason, we propose that the CF redistribution is not secondary to the motor episodes but a consequence of the loss of RIM4 from Purkinje cells. A similar proximal distribution of CF synapses was observed in KO mice lacking the auxiliary VGCC subunit α2δ2.48 Purkinje cells in these mice also showed increased CF EPSC amplitudes and computational modelling suggested that the more proximal location of the synapses can largely explain the increased amplitude. This effect may be even stronger in our case because not only are the synapses more proximal, but the dendrites are also shorter and less branched, making the dendritic tree more electrically compact and facilitating the spread of dendritic charge into the somatic region sampled by the recording pipette.

Taken together, we propose that RIM4 deletion from Purkinje cells leads to disturbed cellular signalling and causes aberrant dendritic arborization, climbing fibre innervation and firing, and that the resulting abnormal inhibitory output of Purkinje cells is responsible for the deteriorated motor coordination, the caffeine-induced and the nocturnal episodes of motor impairment.

What might be the molecular mechanism of RIM4's performance?

Our experiments do not allow us to identify a direct interaction partner of RIM4 or to pinpoint the exact molecular mechanism. However, comparison with other mouse lines with similar phenotypes and with functions of the RIM family of proteins may allow us to narrow down the list of possible candidate mechanisms.

The paroxysmal occurrence of dyskinetic and dystonic attacks lasting hours and their inducibility by caffeine, ethanol and stress are hallmarks of a group of mouse lines with mutations in or loss of the CACNA1A gene, which encodes the alpha subunit of the P/Q-type VGCC. This group includes the Tottering, Rocker and Rolling Nagoya mice, among others, and represents a number of genetic models of the human disease episodic ataxia 2 (EA2).49,50 Many of these mutants are associated with decreased calcium entry through P/Q-type VGCCs, which are the most prominent calcium channel type in Purkinje cells. We did not observe a reduction in calcium entry in situ, which could be taken as evidence for altered function of P/Q-type calcium channels. However, considering that CF-induced calcium signals in Purkinje cells in acute slices (OGB-1) were of similar magnitude in RIM4 KO mice (Fig. 4K and L), although CF EPSCs were greatly increased (Fig. 4O), it appears that recruitment of calcium entry by CF synapses may be less effective in RIM4 KO mice and it is possible that a loss-of-function of P/Q VGCCs is functionally compensated by a regulation of the AMPA-R component of CF synaptic responses in Purkinje cells to restore the levels of calcium entry seen in wild-type mice. RIM family proteins, particularly RIM1/2, are required for maintaining normal levels of presynaptic calcium channels at the AZ.9,10 This function of RIM1/2 is thought to involve a direct interaction of the RIM1/2 PDZ domain with the C-terminus of P/Q, N and R-type VGCCs.9 A PDZ domain is not present in RIM4. In addition to the PDZ domain, the C-terminal C2B domain of RIM1/2 and RIM3/4 has been reported to interact with the β-subunits of VGCCs.15,16 This interaction weakens the calcium-dependent inactivation of VGCCs, thereby increasing calcium entry. This weakening of calcium-dependent inactivation may not significantly affect the brief period of calcium entry mediated by single action potentials, as they occur in presynapses, and therefore not be important for transmitter release (cf. Figs 2 and 5O). Nevertheless, it could substantially elevate calcium entry during longer lasting depolarizations such as complex spike signals (∼30 ms). Thus, the absence of RIM4 may lead to enhanced calcium-dependent inactivation and reduced current through VGCCs in Purkinje cells, causing dyskinesia/dystonia/ataxia as observed in mice with mutated CACNA1A genes. Notably, mutations in the β4 subunit (CACNB4, lethargic mice) also result in paroxysmal dyskinesias, similar to what we observed in RIM4 KO mice.51 However, despite many phenotypic similarities between strains with mutated CACNA1A or CACNB4 genes and RIM4 KO mice, there are some important differences. First, these mouse lines exhibit epileptic seizures (absence) in addition to dyskinesia, even when VGCCs are deleted only in Purkinje cells.41 In RIM4 mice, we never observed signs of such seizures during continuous video and EEG monitoring. Second, the paroxysmal motor episodes of tottering mutant mice respond to treatment with acetazolamide,52,53 whereas this drug had no effect in our RIM4 KOconst mouse line (Supplementary Fig. 9). Third, in RIM4 KO mice, Purkinje cell action potential firing is characterized by heightened regularity (and reduced frequency) and rarely exhibits bursting behaviour (Supplementary Fig. 10). Conversely, in tottering mice, Purkinje cells fire with greater irregularity,54,55 indicating increased bursting. These differences between dystonic/ataxic syndromes associated with CACNA1 and our RIM4 KO mouse lines may raise scepticism regarding the notion that deletion of RIM4 in Purkinje cells acts through altered VGCC signalling.

A second form of paroxysmal dyskinesia and dystonia, characterized by episodes of motor dysfunction lasting for hours, but without seizures, is not associated with VGCCs but with mutations in the MR-1/PNKD gene.56,57 The gene and protein were named after the familial human disease ‘paroxysmal non-kinesigenic dyskinesia’, which occurs in patients carrying mutations (missense p.A7V, p.A9V) in the PNKD gene.58 Mice transgenically expressing the mutant form of the CNS-specific L (long) isoform of the PNKD protein exhibit symptoms resembling the human disease and suffer from paroxysmal dyskinetic and dystonic attacks very similar to what we observed in RIM4 KO mice.59 Dystonic motor episodes can also be triggered in these mice through caffeine and ethanol injections or stress (such as handling them).59 Notably, knockout mice lacking PNKD do not show dystonia episodes.59 Thus, motor episodes do not occur in the presence or absence of PNKD, but only when the mutant form of PNKD is expressed. This highlights the importance of the N-terminal mutations in causing the disease and has led to the conclusion that these mutations confer a gain-of-function, albeit a systemically deleterious function, and has sparked interest in the relevance of this mutation to the pathology. Thus, the absence of the PNKD protein, and therefore any function it might normally have, does not lead to the phenotype. The phenotype only occurs if the mutant protein is expressed, making the consequences of these mutations of particular interest. Using an interaction assay, it was found that wild-type PNKD interacts with the C2B domains of RIM1 and RIM2, but the mutant PNKD does not.17 However, neurons expressing the mutant form of PNKD that does not bind RIM1/2 showed normal synaptic release.17 This suggests that RIM1/2 binding to PNKD does not play a significant role in the occurrence of the motor phenotype, which is consistent with our observation when RIM1/2 is knocked out in Purkinje cells. Notably, the RIM4 C2 domain shares high homology with the C2B domains of RIM1/2 and strongly binds to PNKD (Supplementary Fig. 8). It could be expected that the A7V and A9V mutations in PNKD also block this interaction. PNKD is highly expressed in cerebellar Purkinje cells,59 so it is conceivable that PNKD and RIM4 interact in Purkinje cells in vivo. The similar motor phenotypes resulting from PNDK mutation to a non-RIM binding form and RIM4 deletion, and the lack of such phenotypes following PNKD or RIM1/2 deletion (specifically in Purkinje cells, our study) or reduction of RIM1/2 levels due to PNKD deletion,17 suggest that the PNKD-RIM4 interaction may be required for normal somatodendritic function of Purkinje cells. If this interaction does not occur in Purkinje cells either due to loss of RIM4 or if it is blocked by the mutations in the N-terminus of L-PNKD, PNKD could negatively impact the normal function of Purkinje cells, leading to paroxysmal dyskinesia and dystonia. The cellular function of the PNKD protein is currently unknown, but wild-type PNKD may be involved in the regulation of cellular oxidative stress.56 Mice expressing the mutant form of PNKD show altered dopaminergic signalling in striatal pathways.59,60 However, research about mice with mutations in other brain regions, particularly the cerebellum, is currently unavailable. If murine PNKD mechanistically intersects with RIM4 deficiency as we propose, then PNKD mutant mice should replicate our key findings of altered Purkinje cell physiology. However, even if our hypothesis is correct, it should be noted that RIM4 may fulfil multiple functions in Purkinje cells. Interaction with β-subunits of VGCCs and possibly other effectors may play additional roles in causing the phenotype. Future research is needed to identify the precise molecular interactions and partners responsible for RIM4's previously unidentified role.

Data availability

Further information and requests for resources should be directed to and will be fulfilled upon reasonable request by Susanne Schoch McGovern ([email protected]) and Dirk Dietrich ([email protected]).

Acknowledgements

We would like to acknowledge the assistance of the Virus (supported in part by SFB1089), Proteomics (M. Sylvester), Genomics, Electron Microscopy (H. Beckert, L. Maus) and Bioinformatics Core Facilities of the University of Bonn Medical Faculty and the Genomics and Bioinformatics platform at LIMES. We thank Sabine Opitz, Pia Trebing, Shayne Gilgenbach and Jessica Klemmer for excellent technical assistance, Polina Gulakova for cloning RIM4 viral vectors, Elizabeth Matthews and Sabrina Delattre for performing initial electrophysiological experiments, Tony Cijsouw for initial immunocytochemical trials and discussions, Julia Wolf for support in preparing tissue for electron microscopy, Khondker Ushna Sameen Islam and Erick Martinez-Chavez for help with RNAScope, Ileana Christea for binding assays, Martin Schwarz for help with clearing of PCP2-Cre::B6.Cg-Gt(ROSA)26Sortm14(CAG−tdTomato)Hze/J mouse brains, Lena Gschossmann for support in setting up the in vivo imaging analysis and S. Goebbels and K. Nave for kindly providing Neurod6tm1(cre)Kan mice. Graphical abstract was created with BioRender.com.

Funding

Our work was supported by the German Research Foundation (DFG) (S.S.: SFB1089, SCHO 820/4-1, SCHO 820/6-1, SCHO 820/7-1, SCHO 820/5-2, SPP1757; D.D.: SFB1089, SPP1757, INST1172 15, DI853/3-5&7, INST 217/785-1), the Federal State of Nordrhein-Westfalen (iBehave, S.S.), the BONFOR program of the Medical Faculty of the University of Bonn (S.S., D.D.).

Competing interests

The authors report no competing interests.

Supplementary material

Supplementary material is available at Brain online.

References

1

Lonart
G
.
RIM1: An edge for presynaptic plasticity
.
Trends Neurosci
.
2002
;
25
:
329
332
.

2

Liu
C
,
Kershberg
L
,
Wang
J
,
Schneeberger
S
,
Kaeser
PS
.
Dopamine secretion is mediated by sparse active zone-like release sites
.
Cell
.
2018
;
172
:
706
718.e15
.

3

Powell
CM
,
Schoch
S
,
Monteggia
L
, et al.
The presynaptic active zone protein RIM1α is critical for normal learning and memory
.
Neuron
.
2004
;
42
:
143
153
.

4

Blundell
J
,
Kaeser
PS
,
Südhof
TC
,
Powell
CM
.
RIM1alpha and interacting proteins involved in presynaptic plasticity mediate prepulse inhibition and additional behaviors linked to schizophrenia
.
J Neurosci
.
2010
;
30
:
5326
5333
.

5

Schoch
S
,
Mittelstaedt
T
,
Kaeser
PS
, et al.
Redundant functions of RIM1alpha and RIM2alpha in Ca(2+)-triggered neurotransmitter release
.
EMBO J.
2006
;
25
:
5852
5863
.

6

Castillo
PE
,
Schoch
S
,
Schmitz
F
,
Südhof
TC
,
Malenka
RC
.
RIM1alpha is required for presynaptic long-term potentiation
.
Nature
.
2002
;
415
:
327
330
.

7

Calakos
N
,
Schoch
S
,
Südhof
TC
,
Malenka
RC
.
Multiple roles for the active zone protein RIM1alpha in late stages of neurotransmitter release
.
Neuron
.
2004
;
42
:
889
896
.

8

Schoch
S
,
Castillo
PE
,
Jo
T
, et al.
RIM1alpha forms a protein scaffold for regulating neurotransmitter release at the active zone
.
Nature
.
2002
;
415
:
321
326
.

9

Kaeser
PS
,
Deng
L
,
Wang
Y
, et al.
RIM proteins tether Ca2+ channels to presynaptic active zones via a direct PDZ-domain interaction
.
Cell
.
2011
;
144
:
282
295
.

10

Han
Y
,
Kaeser
PS
,
Südhof
TC
,
Schneggenburger
R
.
RIM determines Ca2+ channel density and vesicle docking at the presynaptic active zone
.
Neuron
.
2011
;
69
:
304
316
.

11

Wang
Y
,
Südhof
TC
.
Genomic definition of RIM proteins: Evolutionary amplification of a family of synaptic regulatory proteins
.
Genomics
.
2003
;
81
:
126
137
.

12

Koushika
SP
,
Richmond
JE
,
Hadwiger
G
,
Weimer
RM
,
Jorgensen
EM
,
Nonet
ML
.
A post-docking role for active zone protein Rim
.
Nat Neurosci
.
2001
;
4
:
997
1005
.

13

de Jong
APH
,
Roggero
CM
,
Ho
MR
, et al.
RIM c2b domains target presynaptic active zone functions to PIP2-containing membranes
.
Neuron
.
2018
;
98
:
1
23
.

14

Alvarez-Baron
E
,
Michel
K
,
Mittelstaedt
T
, et al.
RIM3γ and RIM4γ are key regulators of neuronal arborization
.
J Neurosci
.
2013
;
33
:
824
839
.

15

Uriu
Y
,
Kiyonaka
S
,
Miki
T
, et al.
Rab3-interacting molecule γ isoforms lacking the rab3-binding domain induce long lasting currents but block neurotransmitter vesicle anchoring in voltage-dependent P/Q-type Ca2+ channels
.
J Biol Chem
.
2010
;
285
:
21750
21767
.

16

Kiyonaka
S
,
Wakamori
M
,
Miki
T
, et al.
RIM1 confers sustained activity and neurotransmitter vesicle anchoring to presynaptic Ca2+ channels
.
Nat Neurosci
.
2007
;
10
:
691
701
.

17

Shen
Y
,
Ge
WP
,
Li
Y
, et al.
Protein mutated in paroxysmal dyskinesia interacts with the active zone protein RIM and suppresses synaptic vesicle exocytosis
.
Proc Natl Acad Sci
.
2015
;
112
:
2935
2941
.

18

Zürner
M
,
Mittelstaedt
T
,
tom Dieck
S
,
Becker
A
,
Schoch
S
.
Analyses of the spatiotemporal expression and subcellular localization of liprin-α proteins
.
J Comp Neurol.
2011
;
519
:
3019
3039
.

19

Müller
JA
,
Betzin
J
,
Santos-Tejedor
J
, et al.
A presynaptic phosphosignaling hub for lasting homeostatic plasticity
.
Cell Rep.
2022
;
39
:
110696
.

20

Woitecki
AMH
,
Müller
JA
,
van Loo
KMJ
,
Sowade
RF
,
Becker
AJ
,
Schoch
S
.
Identification of synaptotagmin 10 as effector of NPAS4-mediated protection from excitotoxic neurodegeneration
.
J Neurosci
.
2016
;
36
:
2561
2570
.

21

van Loo
KMJ
,
Rummel
CK
,
Pitsch
J
, et al.
Calcium channel subunit α2δ4 is regulated by early growth response 1 and facilitates epileptogenesis
.
J Neurosci
.
2019
;
39
:
3175
3187
.

22

Jinnah
HA
,
Hess
EJ
.
Animal models of movement disorders
.
Sect A: Sci Found
.
2005
;
15
:
3083
3094
.

23

Walter
JT
,
Alviña
K
,
Womack
MD
,
Chevez
C
,
Khodakhah
K
.
Decreases in the precision of Purkinje cell pacemaking cause cerebellar dysfunction and ataxia
.
Nat Neurosci
.
2006
;
9
:
389
397
.

24

Kalume
F
,
Yu
FH
,
Westenbroek
RE
,
Scheuer
T
,
Catterall
WA
.
Reduced sodium current in Purkinje neurons from Nav1.1 mutant mice: Implications for ataxia in severe myoclonic epilepsy in infancy
.
J Neurosci
.
2007
;
27
:
11065
11074
.

25

D’Angelo
E
.
Physiology of the cerebellum
.
Handb Clin Neurol
.
2018
;
154
:
85
108
.

26

Ross
WN
,
Werman
R
.
Mapping calcium transients in the dendrites of Purkinje cells from the Guinea-pig cerebellum in vitro
.
J Physiol
.
1987
;
389
:
319
336
.

27

Schultz
SR
,
Kitamura
K
,
Post-Uiterweer
A
,
Krupic
J
,
Häusser
M
.
Spatial pattern coding of sensory information by climbing fiber-evoked calcium signals in networks of neighboring cerebellar Purkinje cells
.
J Neurosci : Off J Soc Neurosci
.
2009
;
29
:
8005
8015
.

28

Kitamura
K
,
Hausser
M
.
Dendritic calcium signaling triggered by spontaneous and sensory-evoked climbing fiber input to cerebellar Purkinje cells in vivo
.
J Neurosci
.
2011
;
31
:
10847
10858
.

29

Roh
SE
,
Kim
SH
,
Ryu
C
, et al.
Direct translation of climbing fiber burst-mediated sensory coding into post-synaptic Purkinje cell dendritic calcium
.
eLife
.
2020
;
9
:
e61593
.

30

Najafi
F
,
Giovannucci
A
,
Wang
SSH
,
Medina
JF
.
Sensory-Driven enhancement of calcium signals in individual Purkinje cell dendrites of awake mice
.
Cell Rep
.
2014
;
6
:
792
798
.

31

Kitamura
K
,
Kano
M
.
Dendritic calcium signaling in cerebellar Purkinje cell
.
Neural Netw
.
2013
;
47
:
11
17
.

32

Barski
JJ
,
Dethleffsen
K
,
Meyer
M
.
Cre recombinase expression in cerebellar Purkinje cells
.
Genesis
.
2000
;
28
(
3–4
):
93
98
.

33

Rhyu
IJ
,
Abbott
LC
,
Walker
DB
,
Sotelo
C
.
An ultrastructural study of granule cell/Purkinje cell synapses in tottering (tg/tg), leaner (tg(la)/tg(la)) and compound heterozygous tottering/leaner (tg/tg(la)) mice
.
Neuroscience
.
1999
;
90
:
717
728
.

34

Ichikawa
R
,
Hashimoto
K
,
Miyazaki
T
, et al.
Territories of heterologous inputs onto Purkinje cell dendrites are segregated by mGluR1-dependent parallel fiber synapse elimination
.
Proc Natl Acad Sci
.
2016
;
113
:
2282
2287
.

35

Taniguchi
H
,
He
M
,
Wu
P
, et al.
A resource of Cre driver lines for genetic targeting of GABAergic neurons in cerebral cortex
.
Neuron
.
2011
;
71
:
995
1013
.

36

Goebbels
S
,
Bormuth
I
,
Bode
U
,
Hermanson
O
,
Schwab
MH
,
Nave
K
.
Genetic targeting of principal neurons in neocortex and hippocampus of NEX-Cre mice
.
Genesis
.
2006
;
44
:
611
621
.

37

Kaeser
PS
,
Kwon
HB
,
Chiu
CQ
,
Deng
L
,
Castillo
PE
,
Südhof
TC
.
RIM1alpha and RIM1beta are synthesized from distinct promoters of the RIM1 gene to mediate differential but overlapping synaptic functions
.
J Neurosci
.
2008
;
28
:
13435
13447
.

38

Kupferschmidt
DA
,
Augustin
SM
,
Johnson
KA
,
Lovinger
DM
.
Active zone proteins RIM1αβ are required for normal corticostriatal transmission and action control
.
J Neurosci
.
2019
;
39
:
1457
1470
.

39

Narboux-Nême
N
,
Evrard
A
,
Ferezou
I
, et al.
Neurotransmitter release at the thalamocortical synapse instructs barrel formation but not axon patterning in the somatosensory cortex
.
J Neurosci
.
2012
;
32
:
6183
6196
.

40

Kaeser
PS
,
Deng
L
,
Fan
M
,
Südhof
TC
.
RIM genes differentially contribute to organizing presynaptic release sites
.
Proc Natl Acad Sci U S A.
2012
;
109
:
11830
11835
.

41

Mark
MD
,
Maejima
T
,
Kuckelsberg
D
, et al.
Delayed postnatal loss of P/Q-type calcium channels recapitulates the absence epilepsy, dyskinesia, and ataxia phenotypes of genomic Cacna1A mutations
.
J Neurosci.
2011
;
31
:
4311
4326
.

42

Todorov
B
,
Kros
L
,
Shyti
R
, et al.
Purkinje cell-specific ablation of Cav2.1 channels is sufficient to cause cerebellar ataxia in mice
.
Cerebellum Lond Engl
.
2012
;
11
:
246
258
.

43

Mathy
A
,
Ho
SSN
,
Davie
JT
,
Duguid
IC
,
Clark
BA
,
Häusser
M
.
Encoding of oscillations by axonal bursts in Inferior olive neurons
.
Neuron
.
2009
;
62
:
388
399
.

44

Gruijl
JRD
,
Bosman
LWJ
,
Zeeuw
CID
,
Jeu
MTGD
. Inferior olive: All ins and outs. In:
Manto
M
,
Schmahmann
JD
,
Rossi
F
,
Gruol
DL
,
Koibuchi
N
, eds.
Handbook of the cerebellum and cerebellar disorders
.
Springer
;
2013
:
1013
1058
.

45

Wilson
AM
,
Schalek
R
,
Suissa-Peleg
A
, et al.
Developmental rewiring between cerebellar climbing fibers and Purkinje cells begins with positive feedback synapse addition
.
Cell Rep
.
2019
;
29
:
2849
2861.e6
.

46

Hashimoto
K
,
Kano
M
.
Synapse elimination in the developing cerebellum
.
Cell Mol Life Sci
.
2013
;
70
:
4667
4680
.

47

Kano
M
,
Watanabe
T
,
Uesaka
N
,
Watanabe
M
.
Multiple phases of climbing fiber synapse elimination in the developing cerebellum
.
Cerebellum
.
2018
;
17
:
722
734
.

48

Beeson
KA
,
Beeson
R
,
Westbrook
GL
,
Schnell
E
.
α2δ-2 protein controls structure and function at the cerebellar climbing fiber synapse
.
J Neurosci Official J Soc Neurosci
.
2019
;
40
:
2403
2415
.

49

Hess
EJ
,
Jinnah
HA
. Mouse models of dystonia. In:
Movement disorders
.
Academic Press
;
2015
:
465
481
.

50

Rose
SJ
,
Hess
EJ
. Mouse models of episodic ataxia type 2. In:
LeDoux
MS
, ed.
Movement disorders
. 2nd edn.
Academic Press
;
2015
:
809
814
.

51

Khan
Z
,
Jinnah
HA
.
Paroxysmal dyskinesias in the lethargic mouse mutant
.
J Neurosci
.
2002
;
22
:
8193
8200
.

52

Cramer
SW
,
Popa
LS
,
Carter
RE
,
Chen
G
,
Ebner
TJ
.
Abnormal excitability and episodic low-frequency oscillations in the cerebral cortex of the tottering mouse
.
J Neurosci
.
2015
;
35
:
5664
5679
.

53

Gray
MM
,
Naik
A
,
Ebner
TJ
,
Carter
RE
.
Altered brain state during episodic dystonia in tottering mice decouples primary motor cortex from limb kinematics
.
Dystonia
.
2023
;
2
:
10974
.

54

Alviña
K
,
Khodakhah
K
.
The therapeutic mode of action of 4-aminopyridine in cerebellar ataxia
.
J Neurosci
.
2010
;
30
:
7258
7268
.

55

Alviña
K
,
Khodakhah
K
.
K Ca channels as therapeutic targets in episodic ataxia type-2
.
J Neurosci
.
2010
;
30
:
7249
7257
.

56

Lee
HY
,
Xu
Y
,
Huang
Y
, et al.
The gene for paroxysmal non-kinesigenic dyskinesia encodes an enzyme in a stress response pathway
.
Hum Mol Genet
.
2004
;
13
:
3161
3170
.

57

Rainier
S
,
Thomas
D
,
Tokarz
D
, et al.
Myofibrillogenesis regulator 1 gene mutations cause paroxysmal dystonic choreoathetosis
.
Arch Neurol
.
2004
;
61
:
1025
1029
.

58

Kuruvilla
KP
,
LeDoux
MS
. Mouse models of PNKD. In:
LeDoux
MS
, ed.
Movement disorders
. 2nd edn.
Academic Press
;
2015
:
779
783
.

59

Yang Lee
H
,
Nakayama
J
,
Xu
Y
, et al.
Dopamine dysregulation in a mouse model of paroxysmal nonkinesigenic dyskinesia
.
J Clin Investig
.
2012
;
122
:
507
518
.

60

Nelson
AB
,
Girasole
AE
,
Lee
HY
,
Ptáček
LJ
,
Kreitzer
AC
.
Striatal indirect pathway dysfunction underlies motor deficits in a mouse model of paroxysmal dyskinesia
.
J Neurosci
.
2022
;
42
:
2835
2848
.

Author notes

Dirk Dietrich and Susanne Schoch contributed equally to this work.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://dbpia.nl.go.kr/pages/standard-publication-reuse-rights)

Supplementary data