The R451C mutation in the neuroligin 3 gene induces an ASD-like behavioral phenotype in mice.
In addition to social impairments and stereotyped behaviors, the two major phenotypes supporting an ASD diagnosis, autistic children may also face difficulties in movement planning during locomotion, further pointing to the striatum as an anatomical substrate of ASD-associated phenotypes [61]. Previous work implicated striatal abnormalities as the basis of the social deficits and the restricted and repetitive behaviors characteristic of ASD [62–65]. Indeed, the striatum is anatomically and functionally organized in a ventral part, mediating limbic functions of motivation, reward, and emotion related behaviors, and a dorsal region, implicated in cognitive and motor functions, such as sensorimotor processing, goal-directed and automated behavior [66]. Such functional segregation relies on the topographical organization of the glutamatergic inputs to the striatum, which are distributed in a dorsomedial-to-ventrolateral arrangement from the sensorimotor, associative, and limbic cortices [67]. In particular, the dorsal striatum represents a key brain region for the initiation and control of movements, as well as for habitual and repetitive actions [68–76]. Despite this evidence, while there is a substantial amount of work on the role of the ventral striatum in ASD, the dorsal region has so far been poorly investigated.
We therefore analyzed the ASD-like phenotype of R451C-NL3 knock-in (KI) mice using a comprehensive battery of tests focusing on striatal-dependent motor, social and stereotyped behaviors. Accordingly, restricted interests and stereotypies arise from striatal dysfunction [62–65]. We assessed repetitive behaviors of R451C-NL3 mice by evaluating self-grooming during spontaneous activity in the open field test. Normal grooming behavior involves brief bouts lasting seconds to minutes of scratching and fur cleaning using the forelimbs [77]. We observed a significant increase in the number of complete self-grooming sequences in R451C-NL3 mice, with respect to WT littermates, indicating the expression of a stereotypic behavior in mutants (Fig. 1A1; WT = 5.250 ± 0.729, N = 12; R451C-NL3 = 7.30 ± 0.668, N = 10; Mann-Whitney test, *Pvalue= 0.022). We then utilized, for the first time in the R451C-NL3 mouse model, the marble-burying, a consolidated test in the study of rodent stereotypic behaviors [2, 45, 78]. Interestingly, this test highlighted a significant increase in the digging and burying behavior of mutants, compared to WT mice (Fig. 1A2,A3), further supporting an increase in repetitive behaviors. Specifically, the two-way ANOVA analysis showed a significant main effect of genotype (F(1,27) = 12.60, **Pvalue= 0.001), with a significant increase in the number of marbles buried within all the evaluated time-intervals (Fig. 1A2; Sidak's multiple comparisons post-hoc analysis: 0–5 min, **Pvalue= 0.009; 5–10 min, **Pvalue= 0.008; 10–15 min, *Pvalue= 0.043; 15–20 min, *Pvalue= 0.019; 20–25 min, Pvalue= 0.186; 25–30 min, *Pvalue= 0.033). Overall, throughout the test duration, R451C-NL3 mice buried significantly more marbles compared to their WT littermates (Fig. 1A3; WT = 12.0 ± 1.173, N = 16; R451C-NL3 = 16.23 ± 0.752, N = 13; Mann-Whitney test, **Pvalue= 0.004). We next investigated social behavior, the second core domain disrupted in ASD, which is based on the ventral striatum [79, 80]. To this aim, in order to evaluate social and novel preference of the mutant mice (Fig. 1B1 − 4) we employed the 3-chamber test, which is well-established to assess potential alterations in social behavior. R451C-NL3 mice showed no bias for either side of the chamber. When allowed to freely explore, KI mice spent significantly less time than WT littermates investigating the social stimulus and did not exhibit a significant preference for the social stimulus over the non-social stimulus (Fig. 1B1; Interaction time: WTobject= 60.08 ± 6.811s, WTsocial= 104.8 ± 10.81s, N = 10, Sidak’s multiple comparisons test *Pvalue= 0.018; R451C-NL3object = 75.20 ± 10.16s; R451C-NL3social = 80.11 ± 12.42s, N = 9, Sidak’s multiple comparisons test Pvalue= 0.9997; two-way ANOVA, F(1,34)interaction = 5.978, P = 0.448). R451C-NL3 mice accordingly displayed a significantly lower social preference index than WT mice (Fig. 1B2; WTsocial preference index= 0.633 ± 0.028, R451C-NL3social preference index= 0.509 ± 0.038; unpaired t-test *Pvalue= 0.018). In the second phase of the test, when allowed to freely interact with either a familiar or an unfamiliar mouse, WT mice spent significantly more time sniffing the unfamiliar mouse, whereas mutant mice showed no difference in the interaction time spent with either the familiar or the novel mouse (Fig. 1B3; Interaction time: WTfamiliar = 60.79 ± 7.317s, WTnovel= 104.5 ± 15.46s, N = 10, Sidak’s multiple comparisons test, *Pvalue= 0.037; R451C-NL3familiar = 76.54 ± 8.005s; R451C-NL3novel = 52.41 ± 10.12s, N = 9, Sidak’s multiple comparisons test, Pvalue= 0.586; two-way ANOVA, F(1,34)interaction = 9.670, **Pvalue= 0.0038). Hence, R451C-NL3 mice showed a reduced novel preference index compared to control animals (Fig. 1B2; WTnovel preference index= 0.615 ± 0.049, R451C-NL3novel preference index= 0.400 ± 0.056; Mann-Whitney test *Pvalue= 0.017). It is noteworthy that both clinical studies and preclinical research involving model organisms suggest that motivational impairments in ASD extend beyond social rewards and manifest in response to various nonsocial stimuli [81]. To analyze nonsocial motivation, we performed the sucrose splash test. Mice carrying the R451C-NL3 mutation showed a reduced dorsal self-grooming activity, compared to WT littermates (Fig. 1C1; WT = 10.42 ± 0.857, N = 12; R451C-NL3 = 7.8 ± 0.841, N = 10; unpaired t-test *Pvalue= 0.043), although no differences were observed in the grooming latency compared to their WT littermates (Fig. 1C2; WT = 38.75 ± 7.072 s, N = 12; R451C-NL3 = 41.70 ± 6.414 s, N = 10; unpaired t-test Pvalue= 0.765). Additionally, to exclude cognitive issues as a potential explanation of genotype differences, we assessed the recognition and spatial working memory of the mice using the NORT and Y-maze test, respectively. No significant differences were observed in either task: R451C-NL3 mice were able to discriminate between novel and familiar objects in the NORT test (suppl. Figure 1A1, A2), and exhibited normal spontaneous alternations in the Y-maze, indicating preserved spatial working memory (suppl. Figure 1B1). Analysis of the Y-maze test revealed a significant increase in the total arm entries of R451C-NL3 mice with respect to WT mice, suggesting a hyperactive phenotype of the mutant mice (suppl. Figure 1B2). Hyperactivity of R451C-NL3 mice was observed also in the open field test, where mutants showed a significant increase in the total horizontal distance traveled (Fig. 1D1; WT = 119.0 ± 8.376 m, N = 12; R451C-NL3 = 151.2 ± 6.640 m, N = 10; unpaired t-test **Pvalue= 0.009), and in the horizontal distance traveled in the last two time-intervals of the session (Figs. 1D2; Sidak’s multiple comparisons test 20–25 min, *Pvalue= 0.011; 25–30 min, *Pvalue= 0.043; Two-way ANOVA: main effect of time (F (2.116, 42.31) = 57.53, ****Pvalue< 0.0001, and genotype (F(1,20) = 8.528, **Pvalue= 0.0085). The average speed detected during open field test was not significantly different between the two genotypes (Fig. 1D3; WT = 0.097 ± 0.005 m/s, N = 12; R451C-NL3 = 0.108 ± 0.004 m/s, N = 10; unpaired t-test Pvalue= 0.112), suggesting that the increased locomotor activity of R451C-NL3 mice might be caused by striatal circuitry impairments affecting the ability to set goals, maintain a direction, and follow a smooth trajectory, rather than to enhanced motor skills [61]. To further explore motor function, we examined coordination and learning with the accelerated rotarod test. The performance improved similarly across the training sessions in mutant and control mice (Fig. 1E1; two-way ANOVA, F (4.707, 127.1)time= 37.08, ****P value < 0.0001). However, neither the genotype (Fig. 1E1; two-way ANOVA F(1, 27)genotype= 1.166, Pvalue= 0.289), nor the interaction between genotype and training (Fig. 1E1; two-way ANOVA F (11, 297)interaction= 1.190, Pvalue= 0.293) had a significant impact on the observed improvement. The linear regression analysis did not detect any significant differences between the genotypes in the initial coordination (Fig. 1B2; WT = 8.667 ± 0.724 r.p.m., N = 22; R451C-NL3 = 8.711 ± 0.784 r.p.m., N = 23; unpaired t-test, Pvalue= 0.967) or learning rate (Fig. 1B3; WT = 0.881 ± 0.099, N = 22; R451C-NL3 = 1.009 ± 0.099, N = 23; unpaired t-test, Pvalue= 0.365). Neverthless, R451C-NL3 mice showed a significantly better performance on the test day (day 4) compared to WT littermates (Fig. 1B4; WT = 129.0 ± 7.904 s, N = 22; R451C-NL3 = 155.4 ± 10.29 s, N = 19; unpaired t-test, *Pvalue= 0.045). These findings further point to an impairment of dorsal striatal circuits, which are essential for motor skill acquisition and motor learning [46].
Finally, anxiety-like behaviors were analyzed in both the open field and the EPM, to rule out a possible influence on the results of R451C-NL3 mice in the striatal-dependent behavioral tests (suppl. Figure 1A,B).
Altered protein expression profile in the dorsal striatum of R451C-NL3 mice.
To investigate possible alterations induced by the R451C-NL3 mutation in the dorsal striatum, we first analyzed the protein expression profile, by performing a proteomic analysis. A comparative analysis of differentially expressed proteins (DEPs) in dorsal striatum extracts of control and mutant mice was carried out using LC-MS-based shotgun proteomics. Overall, we quantified 911 proteins in the two mouse groups (N = 5 R451C-NL3; N = 5 WT) and selected 93 DEPs (suppl. Tables S1, S2) with a statistically significant maximum fold change (MFC) of protein expression level set at ≥ 1.3 (ANOVA p value ≤ 0.05). A hierarchical clustering analysis of DEPs was then performed, utilizing Euclidean correlation as the distance metric that is shown in a heat map dendrogram (Fig. 2A). This graphical representation (based on protein recurrence) illustrates the different protein expression in WT vs. KI striatum. Additionally, a volcano scatter plot was employed for rapid visual recognition of proteins exhibiting statistically significant fold change in each comparison (Fig. 2B).
Notably, our analysis showed a marked reduction of NL3 protein abundance (Fig. 2C1; WT = 14.59 ± 3.113, N = 5; R451C-NL3 = 5.482 ± 0.850, N = 5; unpaired t-test, *Pvalue= 0.022) within the dorsal striatum of KI, compared to WT, mice. Accordingly, the Western blot analysis showed that the R451C mutation causes NL3 protein downregulation also in the dorsal striatum (Fig. 2C2; WT = 0.955 ± 0.047, N = 3; R451C-NL3 = 0.217 ± 0.008, N = 3; unpaired t-test, ***Pvalue= 0.0001), similar to what previously reported in different brain areas of this mouse model [82, 83], as well as in heterologous expression cellular systems (De Jaco et al., 2010). Subsequently, the dataset of identified DEPs was utilized to conduct gene ontology enrichment analysis utilizing ShinyGO 0.80, a gene set enrichment tool (https://bioinformatics.sdstate.edu/go/) [84]. The enrichment analysis of DEPs within the dorsal striatum of R451C-NL3 mice identified several dysregulated pathways with respect to WT animals (suppl. Figure S2A,B). Notably, some of the pathways identified in our proteomic analysis are strongly linked to signal transduction pathways, including G protein-coupled receptor (GPCR)-mediated signaling, neurotransmitter receptors, and postsynaptic signal transduction. These observations are in line with experimental data supporting the pivotal role of NL3 in AMPA and NMDA receptor-mediated excitatory transmission in hippocampus and somatosensory cortex [85–87], in long-term synaptic plasticity [36, 86] and in dendritic spine turnover [88, 89].
The R451C mutation impairs the expression of corticostriatal synaptic plasticity.
Corticostriatal synapses express different forms of bidirectional plasticity, potentiation and depression, that can be induced experimentally by high-frequency stimulation (HFS) of afferent fibers [50], and by agonist-induced mGluI receptors activation [34]. As we previously showed [36], the high-frequency stimulation (HFS) of afferent fibers was not able to induce an activity-dependent LTD at dorsal striatum synapses of R451C-NL3 mice (Fig. 3A; EPSP amplitude post-HFS: WTPOST= 51.43 ± 4.78% of pre-HFS control, n = 18; paired t-test, **** Pvalue< 0.0001; R451C-NL3POST = 101.54 ± 3.74% of pre-HFS control, n = 17; paired t-test, Pvalue= 0.684; WTPOST vs. R451C-NL3POST, unpaired t-test, ****Pvalue< 0.0001). We therefore investigated the other forms of long-term synaptic plasticity that can be induced at corticostriatal synapses. We found that activity-dependent long-term potentiation (HFS-LTP), is also impaired in the R451C-NL3 dorsal striatum. Notably, delivery of the induction protocol in R451C-NL3 slices failed to induce an LTP, of amplitude similar to the one observed in WT (Fig. 3B1,2; EPSP amplitude post-HFS: WTPOST= 141.4 ± 2.427% of pre-HFS control, n = 6; paired t-test, ****Pvalue< 0.0001; R451C-NL3POST = 106.4 ± 0.634% of pre-HFS control, n = 6; paired t-test, ***Pvalue= 0.0002; WTPOST vs. R451C-NL3POST; unpaired t-test, ****Pvalue< 0.0001). Interestingly, several lines of evidence suggest a pathophysiological role of mGlu receptors in ASD [9]. We therefore examined the expression of mGluI-dependent pharmacological LTD in R451C-NL3 mice. Bath application of 50 µM (RS)-3,5-dihydroxyphenylglycine (3,5-DHPG; 10 minutes) resulted in a sustained LTD of the excitatory postsynaptic currents (EPSCs) in WT slices. Conversely, the same protocol in KI slices induced only a transient decrease of EPSC amplitude, which returned to levels not significantly different from baseline (Fig. 3C1,2; EPSC amplitude post-DHPG: WTPOST−DHPG = 65.35 ± 4.529% of pre-DHPG control, n = 5, paired t-test, **Pvalue= 0.0016; R451C-NL3POST − DHPG = 98.31 ± 0.530% of pre-DHPG control, n = 8, paired t-test, Pvalue= 0.843; WTPOST−DHPG vs. R451C-NL3POST − DHPG, unpaired t-test, *Pvalue= 0.011).
Exogenous activation of CB1 receptors rescues the expression of mGluI-induced LTD.
We previously showed that the activation of the endocannabinoid (eCB) signaling was able to induce a partial rescue of HFS-LTD in R451C-NL3 slices [36]. We therefore investigated whether the exogenous activation of the eCB system was able to rescue also the pharmacological LTD, induced by 3,5-DHPG application, in KI mice. Corticostriatal slices from mutant and WT mice were perfused with the highly selective CB1 receptor (CB1R) agonist ACEA (20 µM), 10 min before and during DHPG perfusion. The magnitude of DHPG-LTD in control slices was not modified by perfusion with ACEA, suggesting that activation of mGluI receptors induces a production of eCB sufficient for LTD expression in the WT striatum (Fig. 4A1,2; EPSC amplitude post-DHPG: WTPOST−DHPG = 66.68 ± 4.958% of pre-DHPG control, n = 7; paired t-test, ***Pvalue= 0.0005). In KI slices, exogenous CB1R activation was able to fully rescue the expression of DHPG-LTD, which showed a magnitude comparable to the pharmacological LTD of WT slices (Fig. 4A1,2; EPSC amplitude post-DHPG: R451C-NL3POST − DHPG = 72.69 ± 7.350% of pre-DHPG control, n = 4; paired t-test, *Pvalue= 0.033; WTPOST−DHPG vs. R451C-NL3POST − DHPG, unpaired t-test, Pvalue= 0.500). In line with these observations, DHPG-LTD expression was prevented by slice preincubation with the CB1R-specific antagonist AM-251, and only a mild depression of synaptic transmission was observed in both genotypes (Figs. 4B1,2; EPSC amplitude post-DHPG: WTPOST−DHPG = 88.77 ± 7.311% of pre-DHPG control, n = 6, paired t-test, Pvalue= 0.177; R451C-NL3POST − DHPG = 96.36 ± 2.863% of pre-DHPG control, n = 5; paired t-test, Pvalue= 0.273; WTPOST−DHPG vs. R451C-NL3POST − DHPG, unpaired t-test, Pvalue= 0.380).
Overall, these observations show a significant impairment of different forms of long-term synaptic plasticity, encompassing activity-dependent and pharmacological-induced, in the dorsal striatum of R451C-NL3 mice. In particular, the transient depression of EPSC amplitude induced by 3,5-DHPG and the rescue of pharmacological LTD obtained with the CB1R agonist ACEA suggest an impairment of mGluI-mediated signaling in R451C-NL3 mice. Indeed, mGluI receptors are implicated in eCB production [85], and play a pivotal role in modulating synaptic plasticity at excitatory glutamatergic synapses, by promptly regulating the necessary rise in intracellular calcium, needed for activity-dependent eCB retrograde signaling, and the redistribution of AMPA and NMDA receptors [90, 91].
Activation of mGluI receptors does not potentiate NMDA receptor-mediated currents in R451C-NL3 striatal slices.
To further investigate the function of mGluI receptors in the R451C-NL3 striatum, we performed additional electrophysiology experiments. As observed in different brain regions, in the dorsal striatum the NMDA glutamate receptor-mediated inward currents are potentiated by the simultaneous activation of mGlu5 and NMDA receptors. To verify the reproducibility of the NMDA-mediated currents, during whole-cell patch-clamp recordings from SPNs, NMDA (30 µM, 30 s) was applied twice, 5 min apart. NMDA induced transient inward currents of similar amplitude in SPNs from WT and R451C-NL3 mice (data not shown; WT = − 28.9 ± 2.802 pA, n = 17; R451C-NL3 = − 29.86 ± 3.55 pA, n = 24; Mann-Whitney test Pvalue = 0.923). After incubation with 3,5-DHPG (50 µM, 5 min), a significant enhancement of the NMDA-mediated inward current was observed in WT SPNs, but not in KI slices (Fig. 5A; WTDHPG= 160.1 ± 13.70% of pre-DHPG amplitude, n = 7, paired t-test **Pvalue = 0.004; R451C-NL3DHPG = 103.6 ± 2.868% of pre-DHPG amplitude, n = 9, paired t-test Pvalue = 0.250). In a further set of experiments, we additionally verified whether an mGlu5 receptor positive allosteric modulator (PAM), CDPPB, was able to potentiate the NMDA receptor-mediated responses in SPNs from mutant mice. As shown in Fig. 5B, similar to 3,5-DHPG, 50 µM CDPPB increased NMDA receptor-mediated inward currents in WT slices, whereas it failed to enhance the NMDA-induced current response in SPNs of R451C-NL3 mice (WTCDPPB= 162.3 ± 12.68% of pre-CDPPB amplitude, n = 3, paired t-test *Pvalue = 0.039; R451C-NL3CDPPB = 109.3 ± 6.26% of pre-CDPPB amplitude, n = 4, paired t-test Pvalue = 0.236). These data show that the positive modulation exerted by activation of mGlu5 receptors on NMDA-mediated responses is disrupted by the R451C-NL3 mutation. To verify if this alteration was dependent on a dysfunction of NMDA receptors, we further analyzed the ionotropic glutamate-mediated postsynaptic currents in SPNs. Our findings did not show significant differences in the AMPA/NMDA ratio between WT and R451C-NL3 slices (Fig. 5C; WT = 2.739 ± 0.287, n = 5; R451C-NL3 = 2.732 ± 0.273, n = 6; Mann-Whitney test, Pvalue= 0.931). AMPA- and NMDA-mediated current kinetics were also similar in the two genotypes (Fig. 5D1,2; VH= -70 mV, AMPA rise time: WT = 3.385 ± 0.195 ms, R451C-NL3 = 3.995 ± 0.497 ms, unpaired t-test Pvalue= 0.297; VH = + 40 mV, NMDA rise time: WT = 10.13 ± 1.387 ms, R451C-NL3 = 9.987 ± 1.005 ms, unpaired t-test Pvalue= 0.894; figure E1,2;F1,2; VH= -70 mV, AMPA decay time: WT = 27.32 ± 1.043 ms, R451C-NL3 = 38.0 ± 5.369 ms, unpaired t-test Pvalue= 0.098; AMPA weighted tau: WT = 9.748 ± 0.990 ms, R451C-NL3 = 13.70 ± 2.857 ms, unpaired t-test Pvalue= 0.329; VH = + 40 mV, NMDA decay time: WT = 396.3 ± 84.09 ms, R451C-NL3 = 466.5 ± 99.65 ms, unpaired t-test Pvalue= 0.610; NMDA weighted tau: WT = 166.9 ± 31.49 ms, R451C-NL3 = 233.6 ± 29.79 ms, unpaired t-test Pvalue= 0.175). Overall, these data indicate that the R451C mutation does not affect either the AMPA or the NMDA receptor function, hence suggesting an impairment of mGlu5 receptor function.
R451C-NL3 mice exhibit a partial postsynaptic remodeling in the dorsal striatum.
Overall, our findings suggest an impairment of mGluI receptor signaling in the dorsal striatum of R451C-NL3 mice, in line with several reports indicating alterations in the function of these receptors, particularly of mGlu-dependent LTD, in different experimental models of ASD [92, 93]. We first examined the protein levels of mGlu5 and mGlu1 receptors in total lysates of the dorsal striatum of WT and R451C-NL3 mice. We observed a slight, nonsignificant decrease in both mGlu1 and mGlu5 receptor subtypes (Fig. 6A,B; mGlu5: WT = 1.000 ± 0.090, N = 8; R451C-NL3 = 0.812 ± 0.069, N = 8; unpaired t-test P value = 0.121; mGlu1: WT = 1.000 ± 0.059, N = 6; R451C-NL3 = 0.853 ± 0.086, N = 6; unpaired t-test P value = 0.191). MGlu5 and mGlu1 receptors are mainly postsynaptic. PSD-95 and SHANK3 play a critical role in organizing and scaffolding signaling molecules at the PSD, influencing receptor trafficking, anchoring, and expression [94, 95]. Notably, we observed a significant increase in PSD-95 protein level in the dorsal striatum of R451C-NL3 mice (Fig. 6C; WT = 1.000 ± 0.045, N = 8; R451C-NL3 = 1.288 ± 0.0823, N = 10; unpaired t-test *Pvalue= 0.011), whereas the protein level of SHANK3 was significantly reduced (Fig. 6D; WT = 1.000 ± 0.137, N = 7; R451C-NL3 = 0.549 ± 0.073, N = 6; unpaired t-test *Pvalue= 0.019). Next, we examined the protein levels of subunits within the glutamate receptor family. Specifically, we focused on the NMDA (NR2A and NR2B) and AMPA (GluA1 and GluA2) protein expression levels in the dorsal striatum of WT and R451C-NL3 mice. Our analysis ruled out differences between the two genotypes in the expression levels of either NR2A (Fig. 6E; WT = 1 ± 0.053, N = 3; R451C-NL3 = 1.155 ± 0.225, N = 3; unpaired t-test Pvalue= 0.541) or NR2B (Fig. 6F; WT = 1 ± 0.084, N = 6; R451C-NL3 = 1.072 ± 0.152, N = 5; unpaired t-test Pvalue= 0.674). Likewise, our experiments reveal no differences in the expression levels of the AMPA receptor subunits GluA1 (Fig. 6G; WT = 1 ± 0.105, N = 5; R451C-NL3 = 0.853 ± 0.042, N = 6; unpaired t-test Pvalue= 0.195) and GluA2 (Fig. 6H; WT = 1 ± 0.156, N = 3; R451C-NL3 = 1.021 ± 0.062, N = 3; Mann-Whitney test Pvalue= 0.700). To determine the regional specificity of the observed protein expression changes, we performed a complementary Western blot analysis in the cortical region. This analysis examined the levels of mGluR5 and PSD-95. Our findings revealed no significant changes in the expression levels of either protein (suppl. figure S3A, 3B). Overall, our findings show that the R451C mutation in the NL3 gene impairs the function of the striatal glutamatergic synaptic machinery.
Altered expression of mGluI-associated synaptic proteins in the R451C-NL3 striatum.
Given the role of perisynaptic mGlu receptors in plasticity [26, 96], we analyzed their specific expression in synaptosomal preparations obtained from the dorsal striatum of both R451C-NL3 and WT mice. Western blot experiments showed a statistically significant decrease of mGlu5 receptor protein levels in R451C-NL3 with respect to WT striatal synaptosomes (Fig. 7A; WT = 100 ± 5.527, N = 6; R451C-NL3 = 66.85 ± 8.036, N = 6; unpaired t-test **Pvalue= 0.0068). Conversely, synaptic mGlu1 receptor levels were not different between R451C-NL3 and control littermates (Fig. 7B; WT = 100 ± 13,95 N = 4; R451C-NL3 = 80.29 ± 5.504, N = 6; unpaired t-test Pvalue= 0.1675). Since the synaptic scaffolding protein Homer is a known binding partner of group I mGlu receptors [97, 98] whose expression is required in PSD remodeling, modulation of glutamate receptor-mediated functions, regulation of calcium signaling, and plasticity of excitatory synapses [98, 99], we evaluated whether the reduction in mGlu5 receptor synaptic expression was associated with changes in Homer protein levels. We found no significant difference between R451C-NL3 and WT synaptosomal preparations (Fig. 7C; WT = 100 ± 7.464, N = 6; R451C-NL3 = 83.75 ± 6.764, N = 6; unpaired t-test Pvalue= 0.1376). In different ASD preclinical models, evidence of mGluI receptor dysfunction has been provided in the cortex [51]. We therefore analyzed the receptor protein levels in cortical synaptosomal preparations from R451C-NL3 and WT mice. Like in the striatum, also in the cortex we observed a significant reduction of mGlu5 receptor level in synaptosomes, whereas mGlu1 receptor and Homer levels were not altered (suppl. figure S3C,D,E; mGlu5: WT = 100 ± 8.254, N = 5; R451C-NL3 = 69.64 ± 5.02, N = 5; unpaired t-test *Pvalue= 0.0138; mGlu1: WT = 100 ± 6.389, N = 6; R451C-NL3 = 82.89 ± 6.496, N = 6; unpaired t-test **Pvalue= 0.0895; Homer: WT = 100 ± 6.43 N = 5; R451C-NL3 = 88.11 ± 8.51, N = 5; unpaired t-test Pvalue= 0.2480). Overall, our findings support a specific impairment of mGlu5 receptor in R451C-NL3 KI mice, suggesting that its dysfunction might underscore the described deficits of striatal synaptic plasticity [100–102].