Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

SYNGAP1 heterozygosity disrupts sensory processing by reducing touch-related activity within somatosensory cortex circuits

Abstract

In addition to cognitive impairments, neurodevelopmental disorders often result in sensory processing deficits. However, the biological mechanisms that underlie impaired sensory processing associated with neurodevelopmental disorders are generally understudied and poorly understood. We found that SYNGAP1 haploinsufficiency in humans, which causes a sporadic neurodevelopmental disorder defined by cognitive impairment, autistic features, and epilepsy, also leads to deficits in tactile-related sensory processing. In vivo neurophysiological analysis in Syngap1 mouse models revealed that upper-lamina neurons in somatosensory cortex weakly encode information related to touch. This was caused by reduced synaptic connectivity and impaired intrinsic excitability within upper-lamina somatosensory cortex neurons. These results were unexpected, given that Syngap1 heterozygosity is known to cause circuit hyperexcitability in brain areas more directly linked to cognitive functions. Thus, Syngap1 heterozygosity causes a range of circuit-specific pathologies, including reduced activity within cortical neurons required for touch processing, which may contribute to sensory phenotypes observed in patients.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Reduced sensory-evoked brain activity in Syngap1 SSC.
Fig. 2: Reduced ongoing and whisker-generated activity in SSC L2/3 neurons from awake Syngap1 mice.
Fig. 3: Reduced whisker responsiveness of SSC neurons in behaving Syngap1 mice.
Fig. 4: Reduced sensory responsiveness of L2/3 SSC neurons in Syngap1 mice is cortex-specific.
Fig. 5: Reduced sensory responsiveness in both excitatory and inhibitory neuronal populations in L2/3 SSC of Syngap1 mutants.
Fig. 6: In vivo patch clamp reveals that L2/3 SSC neurons in Syngap1 mutants have reduced sensory-evoked synaptic input.
Fig. 7: Syngap1 heterozygosity degrades synaptic connectivity and reduces intrinsic excitability of upper layer SSC neurons.
Fig. 8: Impaired texture discrimination and whisker-dependent go/no-go task performance in Syngap1 mice.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Sahin, M. & Sur, M. Genes, circuits, and precision therapies for autism and related neurodevelopmental disorders. Science 350, aab3897 (2015).

    Article  Google Scholar 

  2. Hull, J. V. et al. Resting-state functional connectivity in autism spectrum disorders: a review. Front. Psychiatry 7, 205 (2017).

    Article  Google Scholar 

  3. Just, M. A., Cherkassky, V. L., Keller, T. A. & Minshew, N. J. Cortical activation and synchronization during sentence comprehension in high-functioning autism: evidence of underconnectivity. Brain 127, 1811–1821 (2004).

    Article  Google Scholar 

  4. Zoghbi, H. Y. & Bear, M. F. Synaptic dysfunction in neurodevelopmental disorders associated with autism and intellectual disabilities. Cold Spring Harb. Perspect. Biol. 4, a009886 (2012).

    Article  Google Scholar 

  5. Robertson, C. E. & Baron-Cohen, S. Sensory perception in autism. Nat. Rev. Neurosci. 18, 671–684 (2017).

    Article  CAS  Google Scholar 

  6. Javitt, D. C. & Sweet, R. A. Auditory dysfunction in schizophrenia: integrating clinical and basic features. Nat. Rev. Neurosci. 16, 535–550 (2015).

    Article  CAS  Google Scholar 

  7. Peron, S. P., Freeman, J., Iyer, V., Guo, C. & Svoboda, K. A cellular resolution map of barrel cortex activity during tactile behavior. Neuron 86, 783–799 (2015).

    Article  CAS  Google Scholar 

  8. Feldmeyer, D. et al. Barrel cortex function. Prog. Neurobiol. 103, 3–27 (2013).

    Article  Google Scholar 

  9. Pluta, S. R., Lyall, E. H., Telian, G. I., Ryapolova-Webb, E. & Adesnik, H. Surround integration organizes a spatial map during active sensation. Neuron 94, 1220–1233.e5 (2017).

    Article  CAS  Google Scholar 

  10. Ogden, K. K., Ozkan, E. D. & Rumbaugh, G. Prioritizing the development of mouse models for childhood brain disorders. Neuropharmacology 100, 2–16 (2016).

    Article  CAS  Google Scholar 

  11. Hoischen, A., Krumm, N. & Eichler, E. E. Prioritization of neurodevelopmental disease genes by discovery of new mutations. Nat. Neurosci. 17, 764–772 (2014).

    Article  CAS  Google Scholar 

  12. Deciphering Developmental Disorders, S.; Deciphering Developmental Disorders Study. Large-scale discovery of novel genetic causes of developmental disorders. Nature 519, 223–228 (2015).

    Article  Google Scholar 

  13. Hamdan, F. F. et al. Mutations in SYNGAP1 in autosomal nonsyndromic mental retardation. N. Engl. J. Med. 360, 599–605 (2009).

    Article  CAS  Google Scholar 

  14. Rauch, A. et al. Range of genetic mutations associated with severe non-syndromic sporadic intellectual disability: an exome sequencing study. Lancet 380, 1674–1682 (2012).

    Article  CAS  Google Scholar 

  15. Deciphering Developmental Disorders, S.; Deciphering Developmental Disorders Study. Prevalence and architecture of de novo mutations in developmental disorders. Nature 542, 433–438 (2017).

    Article  Google Scholar 

  16. O’Roak, B. J. et al. Recurrent de novo mutations implicate novel genes underlying simplex autism risk. Nat. Commun. 5, 5595 (2014).

    Article  Google Scholar 

  17. Hamdan, F. F. et al. De novo SYNGAP1 mutations in nonsyndromic intellectual disability and autism. Biol. Psychiatry 69, 898–901 (2011).

    Article  CAS  Google Scholar 

  18. Berryer, M. H. et al. Mutations in SYNGAP1 cause intellectual disability, autism, and a specific form of epilepsy by inducing haploinsufficiency. Hum. Mutat. 34, 385–394 (2013).

    Article  CAS  Google Scholar 

  19. Parker, M. J. et al. De novo, heterozygous, loss-of-function mutations in SYNGAP1 cause a syndromic form of intellectual disability. Am. J. Med. Genet. A. 167A, 2231–2237 (2015).

    Article  Google Scholar 

  20. Mignot, C. et al. Genetic and neurodevelopmental spectrum of SYNGAP1-associated intellectual disability and epilepsy. J. Med. Genet. 53, 511–522 (2016).

    Article  CAS  Google Scholar 

  21. Carvill, G. L. et al. Targeted resequencing in epileptic encephalopathies identifies de novo mutations in CHD2 and SYNGAP1. Nat. Genet. 45, 825–830 (2013).

    Article  CAS  Google Scholar 

  22. von Stülpnagel, C. et al. SYNGAP1 mutation in focal and generalized epilepsy: a literature overview and a case report with special aspects of the EEG. Neuropediatrics 46, 287–291 (2015).

    Article  Google Scholar 

  23. Clement, J. P. et al. Pathogenic SYNGAP1 mutations impair cognitive development by disrupting maturation of dendritic spine synapses. Cell 151, 709–723 (2012).

    Article  CAS  Google Scholar 

  24. Ozkan, E. D. et al. Reduced cognition in Syngap1 mutants is caused by isolated damage within developing forebrain excitatory neurons. Neuron 82, 1317–1333 (2014).

    Article  CAS  Google Scholar 

  25. Agency for Healthcare Research and Quality. Registries for Evaluating Patient Outcomes: A User’s Guide [Internet]: Rare Disease Registries. (Gliklich, R. E., Dreyer, N. A., & Leavy, M. B.) Chapter 20 (US Department of Health & Human Services, Rockville, MD, USA, 2014).

  26. Berryer, M. H. et al. Decrease of SYNGAP1 in GABAergic cells impairs inhibitory synapse connectivity, synaptic inhibition and cognitive function. Nat. Commun. 7, 13340 (2016).

    Article  CAS  Google Scholar 

  27. Clement, J. P., Ozkan, E. D., Aceti, M., Miller, C. A. & Rumbaugh, G. SYNGAP1 links the maturation rate of excitatory synapses to the duration of critical-period synaptic plasticity. J. Neurosci. 33, 10447–10452 (2013).

    Article  CAS  Google Scholar 

  28. Dana, H. et al. Thy1-GCaMP6 transgenic mice for neuronal population imaging in vivo. PLoS One 9, e108697 (2014).

    Article  Google Scholar 

  29. Gonçalves, J. T., Anstey, J. E., Golshani, P. & Portera-Cailliau, C. Circuit level defects in the developing neocortex of Fragile X mice. Nat. Neurosci. 16, 903–909 (2013).

    Article  Google Scholar 

  30. Wolfe, J., Houweling, A. R. & Brecht, M. Sparse and powerful cortical spikes. Curr. Opin. Neurobiol. 20, 306–312 (2010).

    Article  CAS  Google Scholar 

  31. Margolis, D. J. et al. Reorganization of cortical population activity imaged throughout long-term sensory deprivation. Nat. Neurosci. 15, 1539–1546 (2012).

    Article  CAS  Google Scholar 

  32. Gorski, J. A. et al. Cortical excitatory neurons and glia, but not GABAergic neurons, are produced in the Emx1-expressing lineage. J. Neurosci. 22, 6309–6314 (2002).

    Article  CAS  Google Scholar 

  33. Gray, L. T. et al. Layer-specific chromatin accessibility landscapes reveal regulatory networks in adult mouse visual cortex. eLife 6, e21883 (2017).

    Article  Google Scholar 

  34. Diamond, M. E. & Arabzadeh, E. Whisker sensory system - from receptor to decision. Prog. Neurobiol. 103, 28–40 (2013).

    Article  Google Scholar 

  35. Nelson, S. B. & Valakh, V. Excitatory/inhibitory balance and circuit homeostasis in autism spectrum disorders. Neuron 87, 684–698 (2015).

    Article  CAS  Google Scholar 

  36. Muhia, M., Yee, B. K., Feldon, J., Markopoulos, F. & Knuesel, I. Disruption of hippocampus-regulated behavioural and cognitive processes by heterozygous constitutive deletion of SynGAP. Eur. J. Neurosci. 31, 529–543 (2010).

    Article  Google Scholar 

  37. Guo, X. et al. Reduced expression of the NMDA receptor-interacting protein SynGAP causes behavioral abnormalities that model symptoms of schizophrenia. Neuropsychopharmacology 34, 1659–1672 (2009).

    Article  CAS  Google Scholar 

  38. Sachidhanandam, S., Sermet, B. S. & Petersen, C. C. H. Parvalbumin-expressing GABAergic neurons in mouse barrel cortex contribute to gating a goal-directed sensorimotor transformation. Cell Rep. 15, 700–706 (2016).

    Article  CAS  Google Scholar 

  39. Yang, H., Kwon, S. E., Severson, K. S. & O’Connor, D. H. Origins of choice-related activity in mouse somatosensory cortex. Nat. Neurosci. 19, 127–134 (2016).

    Article  CAS  Google Scholar 

  40. Sachidhanandam, S., Sreenivasan, V., Kyriakatos, A., Kremer, Y. & Petersen, C. C. Membrane potential correlates of sensory perception in mouse barrel cortex. Nat. Neurosci. 16, 1671–1677 (2013).

    Article  CAS  Google Scholar 

  41. Ben-Sasson, A. et al. A meta-analysis of sensory modulation symptoms in individuals with autism spectrum disorders. J. Autism Dev. Disord. 39, 1–11 (2009).

    Article  Google Scholar 

  42. Marco, E. J., Hinkley, L. B., Hill, S. S. & Nagarajan, S. S. Sensory processing in autism: a review of neurophysiologic findings. Pediatr. Res. 69, 48R–54R (2011).

    Article  Google Scholar 

  43. Robertson, A. E. & Simmons, D. R. The relationship between sensory sensitivity and autistic traits in the general population. J. Autism Dev. Disord. 43, 775–784 (2013).

    Article  Google Scholar 

  44. Araki, Y., Zeng, M., Zhang, M. & Huganir, R. L. Rapid dispersion of SynGAP from synaptic spines triggers AMPA receptor insertion and spine enlargement during LTP. Neuron 85, 173–189 (2015).

    Article  CAS  Google Scholar 

  45. Kopanitsa, M. V. et al. Chronic treatment with a MEK inhibitor reverses enhanced excitatory field potentials in Syngap1 +/– mice. Pharmacol. Rep. 70, 777–783 (2018).

    Article  CAS  Google Scholar 

  46. Rumbaugh, G., Adams, J. P., Kim, J. H. & Huganir, R. L. SynGAP regulates synaptic strength and mitogen-activated protein kinases in cultured neurons. Proc. Natl. Acad. Sci. USA 103, 4344–4351 (2006).

    Article  CAS  Google Scholar 

  47. Vazquez, L. E., Chen, H. J., Sokolova, I., Knuesel, I. & Kennedy, M. B. SynGAP regulates spine formation. J. Neurosci. 24, 8862–8872 (2004).

    Article  CAS  Google Scholar 

  48. Walkup, W. G. et al. A model for regulation by SynGAP-alpha1 of binding of synaptic proteins to PDZ-domain ‘Slots’ in the postsynaptic density. eLife 5, e16813 (2016).

    Article  Google Scholar 

  49. Zeng, M., Bai, G. & Zhang, M. Anchoring high concentrations of SynGAP at postsynaptic densities via liquid-liquid phase separation. Small GTPases https://doi.org/10.1080/21541248.2017.1320350 (2017).

  50. Aceti, M. et al. Syngap1 haploinsufficiency damages a postnatal critical period of pyramidal cell structural maturation linked to cortical circuit assembly. Biol. Psychiatry 77, 805–815 (2015).

    Article  CAS  Google Scholar 

  51. Kim, J. H., Lee, H. K., Takamiya, K. & Huganir, R. L. The role of synaptic GTPase-activating protein in neuronal development and synaptic plasticity. J. Neurosci. 23, 1119–1124 (2003).

    Article  CAS  Google Scholar 

  52. Harrison, T. C., Sigler, A. & Murphy, T. H. Simple and cost-effective hardware and software for functional brain mapping using intrinsic optical signal imaging. J. Neurosci. Methods 182, 211–218 (2009).

    Article  Google Scholar 

  53. Chen-Bee, C. H., Kwon, M. C., Masino, S. A. & Frostig, R. D. Areal extent quantification of functional representations using intrinsic signal optical imaging. J. Neurosci. Methods 68, 27–37 (1996).

    Article  CAS  Google Scholar 

  54. Chen-Bee, C. H. et al. Visualizing and quantifying evoked cortical activity assessed with intrinsic signal imaging. J. Neurosci. Methods 97, 157–173 (2000).

    Article  CAS  Google Scholar 

  55. Chen-Bee, C. H. & Frostig, R. D. Variability and interhemispheric asymmetry of single-whisker functional representations in rat barrel cortex. J. Neurophysiol. 76, 884–894 (1996).

    Article  CAS  Google Scholar 

  56. Goldey, G. J. et al. Removable cranial windows for long-term imaging in awake mice. Nat. Protoc. 9, 2515–2538 (2014).

    Article  CAS  Google Scholar 

  57. Holtmaat, A. et al. Imaging neocortical neurons through a chronic cranial window. Cold Spring Harb. Protoc. 2012, 694–701 (2012).

    Article  Google Scholar 

  58. Dubbs, A., Guevara, J. & Yuste, R. moco: fast motion correction for calcium imaging. Front. Neuroinform. 10, 6 (2016).

    Article  Google Scholar 

  59. Patel, T. P., Man, K., Firestein, B. L. & Meaney, D. F. Automated quantification of neuronal networks and single-cell calcium dynamics using calcium imaging. J. Neurosci. Methods 243, 26–38 (2015).

    Article  CAS  Google Scholar 

  60. Desai, N. S., Siegel, J. J., Taylor, W., Chitwood, R. A. & Johnston, D. Matlab-based automated patch-clamp system for awake behaving mice. J. Neurophysiol. 114, 1331–1345 (2015).

    Article  CAS  Google Scholar 

  61. Desai, N. S. & Walcott, E. C. Synaptic bombardment modulates muscarinic effects in forelimb motor cortex. J. Neurosci. 26, 2215–2226 (2006).

    Article  CAS  Google Scholar 

  62. Ozkan, E. D. et al. Input-specific regulation of hippocampal circuit maturation by non-muscle myosin IIB. J. Neurochem. 134, 429–444 (2015).

    Article  CAS  Google Scholar 

  63. Agmon, A. & Connors, B. W. Thalamocortical responses of mouse somatosensory (barrel) cortex in vitro. Neuroscience 41, 365–379 (1991).

    Article  CAS  Google Scholar 

  64. Heyser, C. J. & Chemero, A. Novel object exploration in mice: not all objects are created equal. Behav. Processes 89, 232–238 (2012).

    Article  Google Scholar 

  65. Wu, H. P., Ioffe, J. C., Iverson, M. M., Boon, J. M. & Dyck, R. H. Novel, whisker-dependent texture discrimination task for mice. Behav. Brain Res. 237, 238–242 (2013).

    Article  Google Scholar 

  66. Gaffield, M. A., Amat, S. B., Bito, H. & Christie, J. M. Chronic imaging of movement-related Purkinje cell calcium activity in awake behaving mice. J. Neurophysiol. 115, 413–422 (2016).

    Article  CAS  Google Scholar 

  67. Slotnick, B. A simple 2-transistor touch or lick detector circuit. J. Exp. Anal. Behav. 91, 253–255 (2009).

    Article  Google Scholar 

  68. He, C. X. et al. Tactile defensiveness and impaired adaptation of neuronal activity in the Fmr1 knock-out mouse model of autism. J. Neurosci. 37, 6475–6487 (2017).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported in part by NIH grants from the National Institute of Mental Health (MH096847 and MH108408 to G.R. and MH105400 to C.A.M.), the National Institute for Neurological Disorders and Stroke (NS064079 to G.R. and NS083894 to J.M.C.), and the National Institute for Drug Abuse (DA034116 and DA036376 to C.A.M.). J.L.H. is supported by a National Institute for Neurological Disorders and Stroke Mentored Clinical Scientist Research Career Development Award (NS091381). The SYNGAP1 Natural History Study and Patient Registry is supported by a grant to M.W., J.L.H., and G.R. from The National Organization of Rare Disorders (NORD).

Author information

Authors and Affiliations

Authors

Contributions

S.D.M. performed experiments, designed experiments, analyzed data, co-wrote the manuscript, and edited the manuscript. E.D.O. performed experiments, designed experiments, analyzed data, and edited the manuscript. M.A. performed experiments, designed experiments and analyzed data. S.M. performed experiments, designed experiments, and analyzed data. E.M. performed experiments, designed experiments, and analyzed data. M.W. performed experiments, designed experiments, and analyzed data. N.L. performed experiments and analyzed data. T.V. performed experiments and analyzed data. M.A.G. designed experiments and interpreted data. J.M.C. designed experiments and interpreted data. J.L.H. performed experiments, designed experiments, analyzed data, interpreted data, and edited the manuscript. C.A.M. designed experiments, interpreted data, and edited the manuscript. G.R. conceived the study, designed experiments, interpreted data, co-wrote the manuscript, and edited the manuscript.

Corresponding author

Correspondence to Gavin Rumbaugh.

Ethics declarations

Competing interests

The authors declare no competing financial interests. M.W. is a paid employee of Bridge-the-GAP Educational Research Foundation. G.R. and J.L.H. are unpaid scientific advisors to Bridge-the-GAP Educational Research Foundation.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated supplementary information

Supplementary Figure 1 Impaired sensory-evoked GCaMP6s dynamics in SSC of Syngap1 mice.

(a) Example wide-field ΔF/F images superimposed over craniotomy pictures of WT and Het mice. (b-d) Traces of individual (b) and averaged (c) GCaMP6s responses, as well as a scatter plot showing the integrated ΔF/F signal amplitude (d) in response to 5 pulses at 5 Hz (Student’s t-test t(12)=4.09 p=0.0015). (e-g) Average time courses of GCaMP6s dynamics in response to a single whisker stimulation (e), 5 pulses at 5 Hz (f) and 5 pulses at 1 Hz (g) stimulation protocols. (h) Scatter plot showing ratio of fourth pulse to first pulse in 5 pulse at 1 Hz paradigm (Student’s t-test t(12)=4.16 p=0.0013). Peak (i, j) and integrated (k, l) ΔF/F responses for increasing number of pulses at 10 Hz (j, l) and increasing frequency of pulses at 5 pulses (i, k) (2-way RM-ANOVA : (i) F(1,12)=3.90 p=0.07 for genotype and F(2,24)=7.44 p=0.003 for genotype and frequency interaction; (j) F(1,12)=2.27 p=0.16 for genotype and F(4,48)=2.92 p=0.031 for genotype and frequency interaction; (k) F(1,12)=7.11 p=0.02 for genotype and F(2,24)=4.92 p=0.01 for genotype and frequency interaction; (l) F(1,12)=4.22 p=0.06 for genotype and F(4,48)=5.96 p=0.001 for genotype and frequency interaction). Open circles represent animal means, black lines or closed circles indicate population means and error bars or shaded areas indicate SEMs. Data were obtained n= 6 WT and n=8 Het mice from two cohorts of animals. All statistical tests were two-sided.

Supplementary Figure 2 Cluster analysis of ongoing and whisker-generated spike counts in SSC L2/3 neurons from awake Syngap1 mice.

(a) Relating to Fig. 2, graph showing the fraction of neurons clustered into low- (white), medium- (gray), and high-spiking (black) activity clusters. Mean spike count for each cluster is noted in the graph legend. Number of neurons in each cluster is denoted on the bar graph for each group. (b) Results of Chi-square analysis showing overall effect on how the neurons clustered across the four groups. (c) Results of Chi-square analyses demonstrating differences in each of the three activity clusters when comparing the four experimental groups. (d) Results of Chi-square analyses comparing the four groups to each other for all three activity clusters.

Supplementary Figure 3 Cell-type-specific recombination in Cux2-CreERT2 and Rbp4-Cre lines.

(a) Representative coronal brain sections of a Cux2-Cre+/-;Ai9+/- mouse, where Cre activity is reported by red fluorescence (black) from the TdTomato expression at PND60 (TMX injection at PND2). (b) Representative coronal brain section from a Cux2-Cre+/-;Ai9+/- mouse, where Cre activity is reported by TdTomato expression in the somatosensory cortex (red) at PND60 (TMX injection at PND2). (c) Representative coronal brain sections of a Rbp4Cre+/-;Ai6+/- mouse, where Cre activity is reported by green fluorescence (black) from the zsGreen expression at PND60. (d) Representative coronal brain section from a Rbp4Cre+/-;Ai9+/- mouse where Cre activity is reported by TdTomato expression in the somatosensory cortex (red) at PND60. Validation of the Cre driver lines were repeated in 2-3 mice per genotype.

Supplementary Figure 4 Additional performance metrics of Syngap1 mice in the go–no-go task.

Representative learning curves during Step 2 training for a WT (a, b) and a Het (c, d) mouse depicting probability of licking on Go (black) and NoGo (blue) trials (a, c) and discrimination index (b, d). Learning curves were similar for 6/7 WT and 7/7 Het mice. (e) Normalized bodyweights of mice from one day before water restriction to the completion of the experiment (2-way RM-ANOVA with Bonferroni’s multiple comparison, Genotype: F(1,12)=1.29, p=0.28; Day: F(45,540)=6.64, p= 1.19E-29; Genotype*Day interaction: F(45,540)=0.49, p=0.99). (f, g) Scatter plots of licking behavior during lick-port training depicting mean number of water licks (f, Unpaired t-test: t(12)=0.17, p=0.87) and total reward licks (g, Unpaired t-test t(12)=0.88, p=0.40). (h, i) Averaged response times across sessions during Go (h, 2-way RM-ANOVA with Bonferroni’s multiple comparison, Genotype: F(1,12)=0.33, p=0.58; Session: F(15,180)=16.84, p= 5.965E-27; Genotype*Session interaction: F(15,180)=0.89, p=0.57) and NoGo (i, 2-way RM-ANOVA with Bonferroni’s multiple comparison, Genotype: F(1,12)=0.0059, p=0.94; Session: F(15,180)=13.94, p= 4.88E-23; Genotype*Session interaction: F(15,180)=0.95, p=0.51) trials. (j-l) For Step 2 training, mean number of trials plotted against session (j, 2-way RM-ANOVA with Bonferroni’s multiple comparison, Genotype: F(1,12)=0.88 p=0.37; Session: F(15,180)=1.88 p=0.028; Genotype*Session interaction: F(15,180)=1.56 p=0.088), scatter plots of mean trials per session (k: Unpaired t-test t(12)=1.29, p=0.22) and total trials performed (l: Unpaired t-test t(12)=0.59, p=0.56). (m) Mean number of total licks for each session in Step 2 training (2-way RM-ANOVA with Bonferroni’s multiple comparison, Genotype: F(1,12)=2.08 p=0.18; Session: F(15,180)=3.51 p= 2.73E-5; Genotype*Session interaction: F(15,180)=2.58 p=0.0016). (n) Scatter plot of mean licks per trial in Step 2 training (Mann-Whitney test, U=11, p= 0.097). (o) Scatter plot of total number of licks in Step 2 training (Unpaired t-test t(12)=1.44, p=0.18). (e-o) Open circles represent individual animals, closed circles or solid black lines indicate population means and error bars or shaded area represent the SEMs. (a,c) Solid black and blue dashed lines indicate performance criteria for hits and FAs, respectively. (b,d) Solid black line indicates performance criteria for d’. All data obtained from n=7 WT and n=7 Het mice, from two cohorts of animals. All statistical tests were two-sided.

Supplementary information

Supplementary Figures 1–4

Reporting Summary

Supplementary Table 1

Registry entries with a completed sensory profile

Supplementary Table 2

Genotype of registry entries denoting touch-related sensory abnormalities

Supplementary Table 3

Up/down state properties of L2/3 barrel cortex neurons

Supplementary Video 1

Head-fixed wild-type mouse without a Botox injection

Supplementary Video 2

Head-fixed wild-type mouse after a Botox injection

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Michaelson, S.D., Ozkan, E.D., Aceti, M. et al. SYNGAP1 heterozygosity disrupts sensory processing by reducing touch-related activity within somatosensory cortex circuits. Nat Neurosci 21, 1–13 (2018). https://doi.org/10.1038/s41593-018-0268-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0268-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing