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:

Dendritic channelopathies contribute to neocortical and sensory hyperexcitability in Fmr1−/y mice

Abstract

Hypersensitivity in response to sensory stimuli and neocortical hyperexcitability are prominent features of Fragile X Syndrome (FXS) and autism spectrum disorders, but little is known about the dendritic mechanisms underlying these phenomena. We found that the primary somatosensory neocortex (S1) was hyperexcited in response to tactile sensory stimulation in Fmr1/y mice. This correlated with neuronal and dendritic hyperexcitability of S1 pyramidal neurons, which affect all major aspects of neuronal computation, from the integration of synaptic input to the generation of action potential output. Using dendritic electrophysiological recordings, calcium imaging, pharmacology, biochemistry and a computer model, we found that this defect was, at least in part, attributable to the reduction and dysfunction of dendritic h- and BKCa channels. We pharmacologically rescued several core hyperexcitability phenomena by targeting BKCa channels. Our results provide strong evidence pointing to the utility of BKCa channel openers for the treatment of the sensory hypersensitivity aspects of FXS.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Accelerated spread of evoked neocortical activity in Fmr1−/y mice.
Figure 2: Increase in sensory stimulus evoked responses of neocortical pyramidal neurons in Fmr1−/y mice.
Figure 3: Increase in AP duration, ADP and firing output of neocortical pyramidal neurons of Fmr1−/y mice in vitro.
Figure 4: Altered dendritic properties of neocortical neurons of Fmr1−/y mice in vitro.
Figure 5: Enhanced synaptic summation and AP backpropagation in distal dendrites of Fmr1−/y neurons in vitro.
Figure 6: Decrease in h-channel expression in distal dendrites of Fmr1−/y neurons in vitro.
Figure 7: Computer simulation of the effect of BKCa channel reduction on dendritic excitability.
Figure 8: Rescue of dendritic hyperexcitability and sensory hypersensitivity by boosting BKCa channel activity.

Similar content being viewed by others

References

  1. Gallagher, A. & Hallahan, B. Fragile X-associated disorders: a clinical overview. J. Neurol. 259, 401–413 (2012).

    Article  CAS  PubMed  Google Scholar 

  2. Krueger, D.D. & Bear, M.F. Toward fulfilling the promise of molecular medicine in Fragile X Syndrome. Annu. Rev. Med. 62, 411–429 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Santoro, M.R., Bray, S.M. & Warren, S.T. Molecular mechanisms of Fragile X Syndrome: a twenty-year perspective. Annu. Rev. Pathol. 7, 219–245 (2012).

    CAS  PubMed  Google Scholar 

  4. Baron-Cohen, S., Ashwin, E., Ashwin, C., Tavassoli, T. & Chakrabarti, B. Talent in autism: hyper-systemizing, hyper-attention to detail and sensory hypersensitivity. Phil. Trans. R. Soc. Lond. B 364, 1377–1383 (2009).

    Article  Google Scholar 

  5. Merenstein, S.A. et al. Molecular-clinical correlations in males with an expanded FMR1 mutation. Am. J. Med. Genet. 64, 388–394 (1996).

    Article  CAS  PubMed  Google Scholar 

  6. Miller, L.J. et al. Electrodermal responses to sensory stimuli in individuals with fragile X syndrome: a preliminary report. Am. J. Med. Genet. 83, 268–279 (1999).

    Article  CAS  PubMed  Google Scholar 

  7. Baranek, G.T. et al. Developmental trajectories and correlates of sensory processing in young boys with fragile X syndrome. Phys. Occup. Ther. Pediatr. 28, 79–98 (2008).

    Article  PubMed  Google Scholar 

  8. Rotschafer, S.E. & Razak, K.A. Auditory processing in fragile x syndrome. Front. Cell. Neurosci. 8, 19 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Chen, L. & Toth, M. Fragile X mice develop sensory hyperreactivity to auditory stimuli. Neuroscience 103, 1043–1050 (2001).

    Article  CAS  PubMed  Google Scholar 

  10. Nielsen, D.M., Derber, W.J., McClellan, D.A. & Crnic, L.S. Alterations in the auditory startle response in Fmr1 targeted mutant mouse models of fragile X syndrome. Brain Res. 927, 8–17 (2002).

    Article  CAS  PubMed  Google Scholar 

  11. Rotschafer, S. & Razak, K. Altered auditory processing in a mouse model of fragile X syndrome. Brain Res. 1506, 12–24 (2013).

    Article  CAS  PubMed  Google Scholar 

  12. Michalon, A. et al. Chronic pharmacological mGlu5 inhibition corrects Fragile X in adult mice. Neuron 74, 49–56 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Gibson, J.R., Bartley, A.F., Hays, S.A. & Huber, K.M. Imbalance of neocortical excitation and inhibition and altered UP states reflect network hyperexcitability in the mouse model of Fragile X Syndrome. J. Neurophysiol. 100, 2615–2626 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  14. 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  CAS  PubMed  PubMed Central  Google Scholar 

  15. Hays, S.A., Huber, K.M. & Gibson, J.R. Altered neocortical rhythmic activity states in Fmr1 KO Mice are due to enhanced mGluR5 signaling and involve changes in excitatory circuitry. J. Neurosci. 31, 14223–14234 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Harlow, E.G. et al. Critical period plasticity is disrupted in the barrel cortex of Fmr1 knockout mice. Neuron 65, 385–398 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Paluszkiewicz, S.M., Olmos-Serrano, J.L., Corbin, J.G. & Huntsman, M.M. Impaired inhibitory control of cortical synchronization in fragile X syndrome. J. Neurophysiol. 106, 2264–2272 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Deng, P.Y. et al. FMRP regulates neurotransmitter release and synaptic information transmission by modulating action potential duration via BK channels. Neuron 77, 696–711 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Brown, M.R. et al. Fragile X mental retardation protein controls gating of the sodium-activated potassium channel Slack. Nat. Neurosci. 13, 819–821 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Lee, H.Y. & Jan, L.Y. Fragile X syndrome: mechanistic insights and therapeutic avenues regarding the role of potassium channels. Curr. Opin. Neurobiol. 22, 887–894 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Szlapczynska, M., Bonnan, A., Ginger, M. & Frick, A. Plasticity and pathology of dendritic intrinsic excitability. in Horizons in Neuroscience Research, vol. 14 (eds. Costa, A. & Villalba, E.) 41–88 (Nova Science Publishers, New York, 2014).

  22. Brager, D.H., Akhavan, A.R. & Johnston, D. Impaired dendritic expression and plasticity of h-channels in the Fmr1−/y mouse model of Fragile X Syndrome. Cell Reports 1, 225–233 (2012).

    Article  CAS  PubMed  Google Scholar 

  23. Routh, B.N., Johnston, D. & Brager, D.H. Loss of functional A-type potassium channels in the dendrites of CA1 pyramidal neurons from a mouse model of Fragile X Syndrome. J. Neurosci. 33, 19442–19450 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Spruston, N. Pyramidal neurons: dendritic structure and synaptic integration. Nat. Rev. Neurosci. 9, 206–221 (2008).

    Article  CAS  PubMed  Google Scholar 

  25. London, M. & Häusser, M. Dendritic computation. Annu. Rev. Neurosci. 28, 503–532 (2005).

    Article  CAS  PubMed  Google Scholar 

  26. Frick, A. & Johnston, D. Plasticity of dendritic excitability. J. Neurobiol. 64, 100–115 (2005).

    Article  CAS  PubMed  Google Scholar 

  27. Bureau, I., Shepherd, G.M.G. & Svoboda, K. Circuit and plasticity defects in the developing somatosensory cortex of Fmr1 knock-out mice. J. Neurosci. 28, 5178–5188 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Nimchinsky, E.A., Oberlander, A.M. & Svoboda, K. Abnormal development of dendritic spines in FMR1 knock-out mice. J. Neurosci. 21, 5139–5146 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ferezou, I., Bolea, S. & Petersen, C. Visualizing the cortical representation of whisker touch: voltage-sensitive dye imaging in freely moving mice. Neuron 50, 617–629 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Ferezou, I. et al. Spatiotemporal dynamics of cortical sensorimotor integration in behaving mice. Neuron 56, 907–923 (2007).

    Article  CAS  PubMed  Google Scholar 

  31. Potez, S. & Larkum, M.E. Effect of common anesthetics on dendritic properties in layer 5 neocortical pyramidal neurons. J. Neurophysiol. 99, 1394–1407 (2008).

    Article  PubMed  Google Scholar 

  32. Takahashi, H. & Magee, J.C. Pathway interactions and synaptic plasticity in the dendritic tuft regions of CA1 pyramidal neurons. Neuron 62, 102–111 (2009).

    Article  CAS  PubMed  Google Scholar 

  33. Larkum, M.E., Kaiser, K. & Sakmann, B. Calcium electrogenesis in distal apical dendrites of layer 5 pyramidal cells at a critical frequency of back-propagating action potentials. Proc. Natl. Acad. Sci. USA 96, 14600–14604 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kole, M.H.P., Hallermann, S. & Stuart, G.J. Single Ih channels in pyramidal neuron dendrites: properties, distribution, and impact on action potential output. J. Neurosci. 26, 1677–1687 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Robinson, R.B. Hyperpolarization-activated cation currents: from molecules to physiological function. Annu. Rev. Physiol. 65, 453–480 (2003).

    Article  CAS  PubMed  Google Scholar 

  36. Ulrich, D. Dendritic resonance in rat neocortical pyramidal cells. J. Neurophysiol. 87, 2753–2759 (2002).

    Article  PubMed  Google Scholar 

  37. Pérez-Garci, E., Gassmann, M., Bettler, B. & Larkum, M.E. The GABAB1b isoform mediates long-lasting inhibition of dendritic Ca2+ spikes in layer 5 somatosensory pyramidal neurons. Neuron 50, 603–616 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Benhassine, N. & Berger, T. Large-conductance calcium-dependent potassium channels prevent dendritic excitability in neocortical pyramidal neurons. Pflugers Arch. 457, 1133–1145 (2009).

    Article  CAS  PubMed  Google Scholar 

  39. Notomi, T. & Shigemoto, R. Immunohistochemical localization of I-h channel subunits, HCN1–4, in the rat brain. J. Comp. Neurol. 471, 241–276 (2004).

    Article  CAS  PubMed  Google Scholar 

  40. Salkoff, L., Butler, A., Ferreira, G., Santi, C. & Wei, A. High-conductance potassium channels of the SLO family. Nat. Rev. Neurosci. 7, 921–931 (2006).

    Article  CAS  PubMed  Google Scholar 

  41. Faber, E.S. & Sah, P. Ca2+-activated K+ (BK) channel inactivation contributes to spike broadening during repetitive firing in the rat lateral amygdala. J. Physiol. (Lond.) 552, 483–497 (2003).

    Article  CAS  Google Scholar 

  42. Shruti, S., Clem, R.L. & Barth, A.L. A seizure-induced gain-of-function in BK channels is associated with elevated firing activity in neocortical pyramidal neurons. Neurobiol. Dis. 30, 323–330 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hay, E., Hill, S., Schuermann, F., Markram, H. & Segev, I. Models of neocortical layer 5b pyramidal cells capturing a wide range of dendritic and perisomatic active properties. PLoS Comput. Biol. 7, e1002107 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Nardi, A. & Olesen, S.-P. BK channel modulators: a comprehensive overview. Curr. Med. Chem. 15, 1126–1146 (2008).

    Article  CAS  PubMed  Google Scholar 

  45. Hébert, B. et al. Rescue of fragile X syndrome phenotypes in Fmr1 KO mice by a BKCa channel opener molecule. Orphanet J. Rare Dis. 9, 124 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Golding, N.L., Jung, H.Y., Mickus, T. & Spruston, N. Dendritic calcium spike initiation and repolarization are controlled by distinct potassium channel subtypes in CA1 pyramidal neurons. J. Neurosci. 19, 8789–8798 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Feldmeyer, D. Excitatory neuronal connectivity in the barrel cortex. Front. Neuroanat. 6, 24 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Cruz-Martín, A., Crespo, M. & Portera-Cailliau, C. Delayed stabilization of dendritic spines in Fragile X Mice. J. Neurosci. 30, 7793–7803 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Cruz-Martín, A., Crespo, M. & Portera-Cailliau, C. Glutamate induces the elongation of early dendritic protrusions via mGluRs in wild type mice, but not in fragile X mice. PLoS ONE 7, e32446 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. He, C.X. & Portera-Cailliau, C. The trouble with spines in fragile X syndrome: density, maturity and plasticity. Neuroscience 251, 120–128 (2013).

    Article  CAS  PubMed  Google Scholar 

  51. Mientjes, E.J. et al. The generation of a conditional Fmr1 knock out mouse model to study Fmrp function in vivo. Neurobiol. Dis. 21, 549–555 (2006).

    Article  CAS  PubMed  Google Scholar 

  52. Bakker, C.E. et al. Fmr1 knockout mice: a model to study Fragile X mental retardation. Cell 78, 23–33 (1994).

    Google Scholar 

  53. Palmer, L.M. et al. The cellular basis of GABA(B)-mediated interhemispheric inhibition. Science 335, 989–993 (2012).

    Article  CAS  PubMed  Google Scholar 

  54. Margrie, T.W., Brecht, M. & Sakmann, B. In vivo, low-resistance, whole-cell recordings from neurons in the anesthetized and awake mammalian brain. Pflugers Arch. 444, 491–498 (2002).

    Article  CAS  PubMed  Google Scholar 

  55. Beltramo, R. et al. Layer-specific excitatory circuits differentially control recurrent network dynamics in the neocortex. Nat. Neurosci. 16, 227–234 (2013).

    Article  CAS  PubMed  Google Scholar 

  56. Cheetham, C.E. & Fox, K. Presynaptic development at L4 to l2/3 excitatory synapses follows different time courses in visual and somatosensory cortex. J. Neurosci. 30, 12566–12571 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. de Kock, C.P.J., Bruno, R.M., Spors, H. & Sakmann, B. Layer- and cell type–specific suprathreshold stimulus representation in rat primary somatosensory cortex. J. Physiol. (Lond.) 581, 139–154 (2007).

    Article  CAS  Google Scholar 

  58. Grewe, B.F., Bonnan, A. & Frick, A. Back-propagation of physiological action potential output in dendrites of slender-tufted L5A pyramidal neurons. Front. Cell. Neurosci. 4, 13 (2010).

    PubMed  PubMed Central  Google Scholar 

  59. Mohajerani, M.H. et al. Mirrored bilateral slow-wave cortical activity within local circuits revealed by fast bihemispheric voltage-sensitive dye imaging in anesthetized and awake mice. J. Neurosci. 30, 3745–3751 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Piguel, N.H. et al. Scribble1/AP2 complex coordinates NMDA receptor endocytic recycling. Cell Rep. published online 10.1016/j.celrep.2014.09.017 (7 October 2014).

  61. Hines, M.L. & Carnevale, N.T. The NEURON simulation environment. Neural Comput. 9, 1179–1209 (1997).

    Article  CAS  PubMed  Google Scholar 

  62. Shao, L.R., Halvorsrud, R., Borg-Graham, L. & Storm, J.F. The role of BK-type Ca2+-dependent K+ channels in spike broadening during repetitive firing in rat hippocampal pyramidal cells. J. Physiol. (Lond.) 521, 135–146 (1999).

    Article  CAS  Google Scholar 

  63. Bailey, K.R., Rustay, N.R. & Crawley, J.N. Behavioral phenotyping of transgenic and knockout mice: practical concerns and potential pitfalls. ILAR J. 47, 124–131 (2006).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank B. Sakmann, P. Seeburg, the Max Planck Society and the Deutsche Forschungsgemeinschaft for initial support of this study. We thank J.-C. Delpech, S. Dos Santos Carvalho and W. Crusio for technical and scientific advice. Some data was acquired using equipment of the Bordeaux Imaging Center, the ESPCI ParisTech and Biochemistry Facility of the Bordeaux Neurocampus. This research was funded by FRAXA Research Foundation, INSERM, CNRS, Conseil de la Region d'Aquitaine, Fondation Jérôme Lejeune, Fédération pour la Recherche sur le Cerveau, Fondation pour la Recherche Médicale (SPF20130526794) and Labex-BRAIN (ANR-10-LABEX-43).

Author information

Authors and Affiliations

Authors

Contributions

A.F. conceived the project. Y.Z. and A.B. collected the in vitro electrophysiological and calcium imaging data. G.B. carried out the in vivo electrophysiological experiments. A.B. and N.S. performed the western blot experiments. I.F. performed the VSD imaging experiments. S.P. and M.G. carried out the behavioral experiments. G.L. collected the NEURON modeling data. B.O. provided the second-generation Fmr1/y mouse line. J.R. assisted in data interpretation. A.F., M.G., A.B., Y.Z., G.B. and I.F. wrote the paper and all of the other authors provided feedback.

Corresponding author

Correspondence to Andreas Frick.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Input-output function of the apical dendrite in Fmr1–/y and WT neurons in vitro.

(a) Experimental design: Dendritic whole-cell recordings were performed near the major branch-point. (b) Dendritic voltage responses (top and middle traces; WT: Vm = – 62 mV; Fmr1–/y: Vm = – 61 mV) to suprathreshold current injections at the dendrite (bottom) for a WT and an Fmr1–/y neuron. (c) Average number of spikes as a function of the current injected (Fmr1–/y, n = 59; WT, n = 54; p < 0.05). Data are shown as means ± s.e.m. *p < 0.05 (Fmr1–/y compared to WT). Statistical significance was calculated by repeated measure two-way ANOVA (c).

Supplementary Figure 2 Enhanced summation of simulated EPSPs (sEPSPs) in distal dendrites of Fmr1–/y neurons in vitro.

(a) Experimental design: Dendritic whole-cell recordings were performed near the major branch-point. (b) A train of five sEPSPs at 50 Hz was evoked by current wave injections in a WT dendrite and an Fmr1–/y dendrite (WT: Vm = – 57 mV; Fmr1–/y: Vm = – 58 mV). (c) Average summation ratio of the 5th to 1st sEPSP amplitudes (Fmr1–/y: 1.33 ± 0.03, n = 9; WT: 1.17 ± 0.05, n = 6; p < 0.05). Data are shown as means ± s.e.m. *p < 0.05 (Fmr1–/y compared to WT). Statistical significance was calculated by unpaired Student's t-test (c).

Supplementary Figure 3 Dendritic calcium spikes in Fmr1–/y and WT neurons in vitro.

(a) Experimental design: Dendritic whole-cell recordings were performed near the major branch-point. (b) Example traces of dendritic calcium spikes evoked by current wave injections in control and their abolition following bath-application of the voltage-gated Ca2+ channel blockers NiCl2 (100 µM) and CdCl2 (500 µM) in a WT dendrite (Vm = – 62 mV). (c) Example traces of dendritic calcium spikes in a WT dendrite and an Fmr1–/y dendrite (WT: Vm = – 62 mV; Fmr1–/y: Vm = – 59 mV). (d) Average voltage-time integral of dendritic calcium spikes (Fmr1–/y, 2.58 ± 0.14 mV*s, n = 4; WT, 1.91 ± 0.05 mV*s, n = 4; p < 0.01). Data are shown as means ± s.e.m. **p < 0.01 (Fmr1–/y compared to WT). Statistical significance was calculated by unpaired Student's t-test (d).

Supplementary Figure 4 Hyperexcitability of L5 pyramidal neurons in vivo.

(a) Experimental design: Whole-cell recordings were made from L5 pyramidal neurons in the primary somatosensory cortex of anaesthetized Fmr1–/y and WT mice. Example of a recorded L5 pyramidal neuron (biocytin, red fluorescence) and neuronal cell bodies (DAPI, blue). Arrowheads mark the cell body in L5 and the trajectory of the apical dendritic arbor throughout all the layers towards the pia. Scale bar: 100 µm. (b) Sample traces of trains of three APs at 120 Hz were superimposed and aligned to the last AP of each train (APs are truncated). The difference in ADP amplitude is indicated by an arrow, and the inset illustrates the difference in the AP half-width (solid line). (c) The amplitude of the ADP at 40, 120, and 170 Hz was significantly greater for Fmr1–/y neurons compared to WT neurons (Fmr1–/y: n = 9; WT: n = 9; p < 0.001). (d) The ratio of the half-width of the 3rd and 1st APs at 40, 120, and 170 Hz was significantly greater for Fmr1–/y neurons (Fmr1–/y: n = 9; WT: n = 9; p < 0.001). Data are shown as means ± s.e.m. Statistical significance was calculated by repeated measure two-way ANOVA (c-d).

Supplementary Figure 5 Dendritic cell-attached voltage-clamp recordings of Ih.

(a) Experimental design. Dendritic cell-attached voltage-clamp recordings of Ih were performed near the major branch-point. (b) Examples of Ih measurements and blockade of currents with Cs+ (3 mM, green trace), ZD7288 (100 µM, blue trace), or both drugs (orange trace). Traces are individual measurements.

Supplementary Figure 6 Full-length pictures of the blots presented in Figure 6.

Representative Western blot from Fmr1–/y and WT somatosensory cortex extracts for the membrane-bound fraction of HCN1 and HCN2, and for GAPDH.

Supplementary Figure 7 Effect of h-channel blocker ZD7288 on EPSP summation and resonance frequency in Fmr1–/y and WT neurons.

(a) Experimental design: Synaptic input was evoked by extracellular L1 stimulation while dendritic whole-cell recordings were performed in the distal dendrites of L5B pyramidal neurons. (b) Dendritic voltage traces for EPSPs summation before and after 10 min application of ZD7288 (100 µM) in an Fmr1–/y and a WT neuron. (c) Average increase in the EPSP summation ratio following ZD7288 application (Fmr1–/y: 31 ± 9 %, n = 9; WT: 132 ± 14 %, n = 10). (d) Dendritic voltage responses (top and middle) to the ZAP20 stimulus before (left column) and after (right column) ZD7288 application in an Fmr1–/y and a WT neuron. (e) Impedance amplitude as a function of frequency calculated from corresponding color-coded traces in (d). (f) Average resonance frequency at a holding potential of – 65 mV following ZD7288 application (Fmr1–/y: 0.45 ± 0.14 Hz, n = 10; WT: 0.76 ± 0.15 Hz, n = 10). Data are shown as mean ± s.e.m. ***p < 0.001, *p < 0.05 (ZD7288 compared to baseline). Statistical significance was calculated by repeated measure two-way ANOVA with Bonferroni’s multiple comparisons test (c, f).

Supplementary Figure 8 Effect of h-channel blocker ZD7288 on subthreshold membrane properties of Fmr1–/y and WT neurons.

(a) Experimental design: Dendritic whole-cell recordings were performed in the distal dendrites of L5B pyramidal neurons. (b) Dendritic voltage responses (top; WT: Vm = – 62 mV; Fmr1–/y: Vm = – 61 mV) to subthreshold current injections (bottom) before and after 10 min application of the h-channel blocker ZD7288 (100 µM) in a WT (left) and an Fmr1–/y (right) neuron. (c) Average values of input resistance in control and following ZD7288 application (Fmr1–/y-control 74.5 ± 11.1 MΩ, Fmr1–/y-ZD7288 120.1 ± 10.1 MΩ, n = 13; WT-control 51.6 ± 3.6 MΩ, WT-ZD7288 122.7 ± 8.0 MΩ, n = 21). (d) Average sag ratio values in control and following ZD7288 application (Fmr1–/y-control 24.4 ± 3.0 %, Fmr1–/y-ZD7288 1.4 ± 0.2 %, n = 13; WT-control 34.6 ± 2.4 %, WT-ZD7288 1.2 ± 0.2 %, n = 22). The sag response was abolished after blocking Ih (e) Average membrane time constant values before and after Ih blockade (Fmr1–/y-control 13.2 ± 2.6 ms, Fmr1–/y-ZD7288 30.7 ± 4.0 ms, n = 11; WT-control 8.7 ± 1.3 ms, WT-ZD7288 26.9 ± 2.6 ms, n = 11). Data are shown as mean ± s.e.m. ***p < 0.001 (ZD7288 compared to baseline). Statistical significance was calculated by repeated measure two-way ANOVA with Bonferroni’s multiple comparisons test (c–e).

Supplementary Figure 9 Effect of h-channel blocker ZD7288 on firing properties of Fmr1–/y and WT neurons.

(a) Experimental design: Dendritic whole-cell recordings were performed in the distal dendrites of L5B pyramidal neurons. (b) Dendritic voltage responses to suprathreshold current injections into the dendrite (bottom) in control (top traces) and after ZD7288 application (middle traces) in a WT (left) and an Fmr1–/y (right) neuron. (c) Average number of dendritic spikes as a function of the current injected before (solid line) and after (dashed line) blocking h-channels (ZD7288: Fmr1–/y, n = 12; WT, n =16; p = 0.005). Data are shown as mean ± s.e.m. Statistical significance was calculated by repeated measure two-way ANOVA.

Supplementary Figure 10 Reduced effect of Ih blocker ZD7288 on dendritic calcium signaling in Fmr1–/y neurons as compared to WT neurons.

(a) Experimental design: Ca2+ transients evoked by back-propagating APs were measured using 2-photon laser scanning microscopy in line-scan mode at the major apical branch-point as relative changes in Fluo-5F (green) fluorescence (normalized to Alexa-594 (red) fluorescence). (b) Dendritic calcium traces in response to a somatic burst of three APs at 100 Hz before and after ZD7288 (20 µM). (c) Average effect of ZD7288 on the amplitude of Ca2+ transients evoked by trains of three bAPs at 100 Hz (Fmr1–/y: 7.5 ± 21.3 %, n = 12; WT: 207.2 ± 61.7 %, n = 10). Data are shown as mean ± s.e.m. ***p < 0.001 (ZD7288 compared to baseline). Statistical significance was calculated by repeated measure two-way ANOVA with Bonferroni’s multiple comparisons test.

Supplementary Figure 11 Rescue of action potential duration by boosting BKCa channel activity.

(a) Experimental design: Ca2+ transients evoked by back-propagating APs were measured using 2-photon laser scanning microscopy in line-scan mode at the major apical branch-point as relative changes in Fluo-5F (green) fluorescence (normalized to Alexa-594 (red) fluorescence). (b) Dendritic calcium traces in response to a somatic burst of three APs at 100 Hz before and after ZD7288 (20 µM). (c) Average effect of ZD7288 on the amplitude of Ca2+ transients evoked by trains of three bAPs at 100 Hz (Fmr1–/y: 7.5 ± 21.3 %, n = 12; WT: 207.2 ± 61.7 %, n = 10). Data are shown as mean ± s.e.m. ***p < 0.001 (ZD7288 compared to baseline). Statistical significance was calculated by repeated measure two-way ANOVA with Bonferroni’s multiple comparisons test.

Supplementary Figure 12 Hypersensitivity to auditory stimuli.

(a) Experimental design: Mice were placed in a restraining tube positioned over a motion-sensitive platform inside a sound proof recording chamber. The whole-body startle response (of Fmr1–/y or WT mice) to brief auditory stimuli at 71 (+ 6), 77 (+ 12), 83 (+ 18), and 89 (+ 24) dB over a white background noise (65 dB) was measured. (b) The startle responses of Fmr1–/y mice were significantly increased compared to WT mice (n = 10 mice per genotype, p < 0.01). This effect was observed at all stimulus intensities. Data are shown as the mean ± s.e.m. **p < 0.01 (Fmr1–/y versus WT mice). Statistical analysis was performed using a 2 x 4 (genotype x stimulus level) ANOVA.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–12 and Supplementary Tables 1–3 (PDF 1572 kb)

Supplementary Methods Checklist

(PDF 518 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, Y., Bonnan, A., Bony, G. et al. Dendritic channelopathies contribute to neocortical and sensory hyperexcitability in Fmr1−/y mice . Nat Neurosci 17, 1701–1709 (2014). https://doi.org/10.1038/nn.3864

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3864

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