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:

Regulation of interneuron excitability by gap junction coupling with principal cells

Abstract

Electrical coupling of inhibitory interneurons can synchronize activity across multiple neurons, thereby enhancing the reliability of inhibition onto principal cell targets. It is unclear whether downstream activity in principal cells controls the excitability of such inhibitory networks. Using paired patch-clamp recordings, we show that excitatory projection neurons (fusiform cells) and inhibitory stellate interneurons of the dorsal cochlear nucleus form an electrically coupled network through gap junctions containing connexin36 (Cxc36, also called Gjd2). Remarkably, stellate cells were more strongly coupled to fusiform cells than to other stellate cells. This heterologous coupling was functionally asymmetric, biasing electrical transmission from the principal cell to the interneuron. Optogenetically activated populations of fusiform cells reliably enhanced interneuron excitability and generated GABAergic inhibition onto the postsynaptic targets of stellate cells, whereas deep afterhyperpolarizations following fusiform cell spike trains potently inhibited stellate cells over several hundred milliseconds. Thus, the excitability of an interneuron network is bidirectionally controlled by distinct epochs of activity in principal cells.

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

Figure 1: Asymmetric electrical coupling between DCN fusiform and stellate cells.
Figure 2: Fusiform cells generate spikelets in electrically coupled stellate cells.
Figure 3: Fusiform cell action potentials propagate into the distal apical dendrites.
Figure 4: Control of the stellate cell membrane potential by fusiform cells and auditory nerve activity.
Figure 5: Optogenetic activation of fusiform cells depolarizes stellate cells.
Figure 6: Fusiform cells control the spike rate of stellate cells.
Figure 7: The timing of fusiform cell activity bidirectionally controls stellate cell spike output.
Figure 8: Fusiform cells generate inhibition in the DCN.

Similar content being viewed by others

References

  1. Bell, C.C., Han, V. & Sawtell, N.B. Cerebellum-like structures and their implications for cerebellar function. Annu. Rev. Neurosci. 31, 1–24 (2008).

    Article  CAS  Google Scholar 

  2. Dean, P., Porrill, J., Ekerot, C.-F. & Jörntell, H. The cerebellar microcircuit as an adaptive filter: experimental and computational evidence. Nat. Rev. Neurosci. 11, 30–43 (2010).

    Article  CAS  Google Scholar 

  3. Requarth, T. & Sawtell, N.B. Neural mechanisms for filtering self-generated sensory signals in cerebellum-like circuits. Curr. Opin. Neurobiol. 21, 602–608 (2011).

    Article  CAS  Google Scholar 

  4. Oertel, D. & Young, E.D. What's a cerebellar circuit doing in the auditory system? Trends Neurosci. 27, 104–110 (2004).

    Article  CAS  Google Scholar 

  5. Roberts, M.T. & Trussell, L.O. Molecular layer inhibitory interneurons provide feedforward and lateral inhibition in the dorsal cochlear nucleus. J. Neurophysiol. 104, 2462–2473 (2010).

    Article  Google Scholar 

  6. Bennett, M.V.L. & Zukin, R.S. Electrical coupling and neuronal synchronization in the mammalian brain. Neuron 41, 495–511 (2004).

    Article  CAS  Google Scholar 

  7. Hestrin, S. & Galarreta, M. Electrical synapses define networks of neocortical GABAergic neurons. Trends Neurosci. 28, 304–309 (2005).

    Article  CAS  Google Scholar 

  8. Christie, M.J., Williams, J.T. & North, R.A. Electrical coupling synchronizes subthreshold activity in locus coeruleus neurons in vitro from neonatal rats. J. Neurosci. 9, 3584–3589 (1989).

    Article  CAS  Google Scholar 

  9. Maher, B.J., McGinley, M.J. & Westbrook, G.L. Experience-dependent maturation of the glomerular microcircuit. Proc. Natl. Acad. Sci. USA 106, 16865–16870 (2009).

    Article  CAS  Google Scholar 

  10. Bennett, M.V. Physiology of electrotonic junctions. Ann. NY Acad. Sci. 137, 509–539 (1966).

    Article  CAS  Google Scholar 

  11. Gibson, J.R., Beierlein, M. & Connors, B.W. Two networks of electrically coupled inhibitory neurons in neocortex. Nature 402, 75–79 (1999).

    Article  CAS  Google Scholar 

  12. Wouterlood, F.G., Mugnaini, E., Osen, K.K. & Dahl, A.L. Stellate neurons in rat dorsal cochlear nucleus studies with combined Golgi impregnation and electron microscopy: synaptic connections and mutual coupling by gap junctions. J. Neurocytol. 13, 639–664 (1984).

    Article  CAS  Google Scholar 

  13. Mann-Metzer, P. & Yarom, Y. Electrotonic coupling interacts with intrinsic properties to generate synchronized activity in cerebellar networks of inhibitory interneurons. J. Neurosci. 19, 3298–3306 (1999).

    Article  CAS  Google Scholar 

  14. Leao, R.M., Li, S., Doiron, B. & Tzounopoulos, T. Diverse levels of an inwardly rectifying potassium conductance generate heterogeneous neuronal behavior in a population of dorsal cochlear nucleus pyramidal neurons. J. Neurophysiol. 107, 3008–3019 (2012).

    Article  CAS  Google Scholar 

  15. Davis, K.A. & Young, E.D. Pharmacological evidence of inhibitory and disinhibitory neuronal circuits in dorsal cochlear nucleus. J. Neurophysiol. 83, 926–940 (2000).

    Article  CAS  Google Scholar 

  16. Molitor, S.C. & Manis, P.B. Dendritic Ca2+ transients evoked by action potentials in rat dorsal cochlear nucleus pyramidal and cartwheel neurons. J. Neurophysiol. 89, 2225–2237 (2003).

    Article  CAS  Google Scholar 

  17. Tzounopoulos, T., Kim, Y., Oertel, D. & Trussell, L.O. Cell-specific, spike timing–dependent plasticities in the dorsal cochlear nucleus. Nat. Neurosci. 7, 719–725 (2004).

    Article  CAS  Google Scholar 

  18. Zhang, S. & Oertel, D. Neuronal circuits associated with the output of the dorsal cochlear nucleus through fusiform cells. J. Neurophysiol. 71, 914–930 (1994).

    Article  CAS  Google Scholar 

  19. Dugué, G.P. et al. Electrical coupling mediates tunable low-frequency oscillations and resonance in the cerebellar Golgi cell network. Neuron 61, 126–139 (2009).

    Article  Google Scholar 

  20. Vervaeke, K. et al. Rapid desynchronization of an electrically coupled interneuron network with sparse excitatory synaptic input. Neuron 67, 435–451 (2010).

    Article  CAS  Google Scholar 

  21. Gibson, J.R., Beierlein, M. & Connors, B.W. Functional properties of electrical synapses between inhibitory interneurons of neocortical layer 4. J. Neurophysiol. 93, 467–480 (2005).

    Article  Google Scholar 

  22. Ma, W.-L.D. & Brenowitz, S.D. Single-neuron recordings from unanesthetized mouse dorsal cochlear nucleus. J. Neurophysiol. 107, 824–835 (2012).

    Article  Google Scholar 

  23. Rhode, W.S., Smith, P.H. & Oertel, D. Physiological response properties of cells labeled intracellularly with horseradish peroxidase in cat dorsal cochlear nucleus. J. Comp. Neurol. 213, 426–447 (1983).

    Article  CAS  Google Scholar 

  24. Hancock, K.E. & Voigt, H.F. Intracellularly labeled fusiform cells in dorsal cochlear nucleus of the gerbil. I. Physiological response properties. J. Neurophysiol. 87, 2505–2519 (2002).

    Article  Google Scholar 

  25. Merchan, M.A., Collia, F.P., Merchan, J.A. & Saldana, E. Distribution of primary afferent fibres in the cochlear nuclei. A silver and horseradish peroxidase (HRP) study. J. Anat. 141, 121–130 (1985).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Arenkiel, B.R. et al. In vivo light-induced activation of neural circuitry in transgenic mice expressing channelrhodopsin-2. Neuron 54, 205–218 (2007).

    Article  CAS  Google Scholar 

  27. Hägglund, M., Borgius, L., Dougherty, K.J. & Kiehn, O. Activation of groups of excitatory neurons in the mammalian spinal cord or hindbrain evokes locomotion. Nat. Neurosci. 13, 246–252 (2010).

    Article  Google Scholar 

  28. Ito, T. & Oliver, D.L. Origins of glutamatergic terminals in the inferior colliculus identified by retrograde transport and expression of VGLUT1 and VGLUT2 genes. Front. Neuroanat. 4, 135 (2010).

    Article  Google Scholar 

  29. Mugnaini, E. GABA neurons in the superficial layers of the rat dorsal cochlear nucleus: light and electron microscopic immunocytochemistry. J. Comp. Neurol. 235, 61–81 (1985).

    Article  CAS  Google Scholar 

  30. Nagel, G. et al. Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. USA 100, 13940–13945 (2003).

    Article  CAS  Google Scholar 

  31. Mitchell, S.J. & Silver, R.A. Shunting inhibition modulates neuronal gain during synaptic excitation. Neuron 38, 433–445 (2003).

    Article  CAS  Google Scholar 

  32. Rubio, M.E. & Juiz, J.M. Differential distribution of synaptic endings containing glutamate, glycine, and GABA in the rat dorsal cochlear nucleus. J. Comp. Neurol. 477, 253–272 (2004); erratum 485, 266 (2005).

    Article  CAS  Google Scholar 

  33. Isaacson, J.S. & Scanziani, M. How inhibition shapes cortical activity. Neuron 72, 231–243 (2011).

    Article  CAS  Google Scholar 

  34. Osen, K.K., Ottersen, O.P. & Storm-Mathisen, J. Colocalization of glycine-like and GABA-like immunoreactivities: a semi-quantitative study of individual neurons in the dorsal cochlear nucleus of cat. in Glycine Neurotransmission (eds. Ottersen, O.P. & Storm-Mathisen, J.): 417–451 (J. Wiley and Sons, New York, 1990).

  35. Bell, C.C., Han, V.Z., Sugawara, Y. & Grant, K. Synaptic plasticity in a cerebellum-like structure depends on temporal order. Nature 387, 278–281 (1997).

    Article  CAS  Google Scholar 

  36. Han, V.Z., Grant, K. & Bell, C.C. Reversible associative depression and nonassociative potentiation at a parallel fiber synapse. Neuron 27, 611–622 (2000).

    Article  CAS  Google Scholar 

  37. Bell, C.C., Caputi, A. & Grant, K. Physiology and plasticity of morphologically identified cells in the mormyrid electrosensory lobe. J. Neurosci. 17, 6409–6423 (1997).

    Article  CAS  Google Scholar 

  38. Sawtell, N.B. & Williams, A. Transformations of electrosensory encoding associated with an adaptive filter. J. Neurosci. 28, 1598–1612 (2008).

    Article  CAS  Google Scholar 

  39. Hansel, C. et al. αCaMKII Is essential for cerebellar LTD and motor learning. Neuron 51, 835–843 (2006).

    Article  CAS  Google Scholar 

  40. Szapiro, G. & Barbour, B. Multiple climbing fibers signal to molecular layer interneurons exclusively via glutamate spillover. Nat. Neurosci. 10, 735–742 (2007).

    Article  CAS  Google Scholar 

  41. Mathews, P.J., Lee, K.H., Peng, Z., Houser, C.R. & Otis, T.S. Effects of climbing fiber driven inhibition on Purkinje neuron spiking. J. Neurosci. 32, 17988–17997 (2012).

    Article  CAS  Google Scholar 

  42. Coddington, L.T., Rudolph, S., Vande Lune, P., Overstreet-Wadiche, L. & Wadiche, J.I. Spillover-mediated feedforward inhibition functionally segregates interneuron activity. Neuron 78, 1050–1062 (2013).

    Article  CAS  Google Scholar 

  43. Zhao, S. et al. Cell type–specific channelrhodopsin-2 transgenic mice for optogenetic dissection of neural circuitry function. Nat. Methods 8, 745–752 (2011).

    Article  CAS  Google Scholar 

  44. Kuo, S.P. & Trussell, L.O. Spontaneous spiking and synaptic depression underlie noradrenergic control of feed-forward inhibition. Neuron 71, 306–318 (2011).

    Article  CAS  Google Scholar 

  45. Apostolides, P.F. & Trussell, L.O. Rapid, Activity-independent turnover of vesicular transmitter content at a mixed glycine/GABA synapse. J. Neurosci. 33, 4768–4781 (2013).

    Article  CAS  Google Scholar 

  46. Bender, K.J., Ford, C.P. & Trussell, L.O. Dopaminergic modulation of axon initial segment calcium channels regulates action potential initiation. Neuron 68, 500–511 (2010).

    Article  CAS  Google Scholar 

  47. Devor, A. & Yarom, Y. Electrotonic coupling in the inferior olivary nucleus revealed by simultaneous double patch recordings. J. Neurophysiol. 87, 3048–3058 (2002).

    Article  Google Scholar 

Download references

Acknowledgements

We thank M. Roberts and S. Kuo for preliminary observations that led us to search for electrical coupling in the DCN; M. Bateschell and R. Larisch for help with mouse colony management and genotyping; C. Borges-Merjane (Vollum Institute/Oregon Hearing Research Center) for initially genotyping the Thy1-ChR2 mice and providing the cerebellum slices used in Supplementary Figure 2; H.-W. Lu for writing the macros used to analyze the calcium imaging data; S. Foster for performing the auditory brainstem response measurements; and K. Bender, W. Giardino and N. Sawtell for critical comments on the manuscript. Funding was provided by US National Institutes of Health grants R01DC004450 to L.O.T., F31DC012222 to P.F.A. and P30 DC005983 to the Oregon Hearing Research Center.

Author information

Authors and Affiliations

Authors

Contributions

P.F.A. collected the data. P.F.A. and L.O.T. designed the study, analyzed and interpreted the results and wrote the manuscript.

Corresponding author

Correspondence to Laurence O Trussell.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 The gap junction blocker MFA abolishes electrical coupling.

a) Average traces from a coupled fusiform-stellate cell pair (upper and lower traces, respectively). Black traces are the baseline period and red traces are after 15-30 min of incubation in the gap junction blocker MFA (100-200 μM). The increased amplitude in the stellate cell prejunctional waveform after MFA likely reflects a contribution of gap junction channels to the basal leak conductance. b) Summary of the coupling coefficients from six pairs before and after application of MFA. Open circles represent individual experiments, red points are mean±1SEM. Left and right panels represent fusiform-to-stellate and stellate-to-fusiform directions, respectively. MFA significantly reduced the coupling coefficient in the fusiform-to-stellate cell direction from 0.10±10.02 to 0.008±1 0.002 (t(5)=4.74, p=0.005), whereas the stellate-to-fusiform coupling dropped from 0.018±10.003 to 0.002±10.001 (t(5)=4.63, 0.006, paired t-tests).

Supplementary Figure 2 DCN stellate cells are weakly electrically-coupled.

a) Top: Example average traces from an electrically-coupled pair of DCN stellate cells. Lower panel: Summary data plotting the coupling coefficients for 9 coupled pairs. The apparent directional asymmetry in some pairs is probably due to intrinsic variability in basal input resistance, as the group average (red point) falls on the unity line. The homologous coupling coefficient between DCN stellate cells (0.024±0.005, average of both directions) was significantly weaker than fusiform-to-stellate cell coupling (F(2,123)=23.82, p=0.0006, One way ANOVA + Dunnett's multiple comparisons tests) but not significantly different from stellate-to-fusiform coupling (p=0.998). The calculated junctional conductance for homologous coupling between DCN stellate cells was 0.14±0.04 nS. b) Top: Example average traces from a pair of cerebellar molecular layer interneurons. 9/18 pairs showed coupling, and the coupling coefficient (0.15±0.02) was significantly greater than in DCN stellate cells (t(17)=6.83, p<0.0001, unpaired t-test). These values are in line with previous reports13. The junctional conductance for coupling between cerebellar stellate cells was 0.52± 0.07 nS.

Supplementary Figure 3 Fusiform cells are weakly electrically coupled to one another.

a) Example average traces from a pair of fusiform cells. Current injection in one cell results in a small hyperpolarization of the other cell. Black and red traces are before and after addition of the gap junction blocker MFA, confirming that this effect is due to gap junction coupling. b) Summary data showing the coupling coefficient of 20 similar pairs. Red symbol is mean±1SEM. Dotted gray line is unity. Homologous coupling between fusiform cells (0.014±10.001, average of both directions) was significantly weaker than fusiform-to-stellate cell coupling (F(3,143)=22.41, p<0.0001, one way ANOVA and Dunnett's multiple comparison test), but not significantly different than stellate-to-fusiform or stellate-to-stellate coupling (p=0.74 and p=0.95, respectively). The junctional conductance for this data set was 0.50±10.07 nS. c) Summary data showing that 100-200 μM MFA significantly reduced electrical coupling between fusiform cells (n=3 pairs; t(2)=9.5, p=0.01, paired t-test.) Open symbols are individual experiments, red is mean±1SEM. Asterisks denote statistical significance. d) Average traces from a fusiform cell pair in a Cx36-/- mouse. Electrical coupling was absent in fusiform cell pairs recorded in Cx36-/- mice (0/20 connected. c2(1)=22.64,p<0.0001 compared to wild-type mice.)

Supplementary Figure 4 Stellate cells do not express ChR2

a) Voltage-clamped fusiform (left) or stellate cell (right) at different membrane potentials. The -67 and +53 mV traces are highlighted for comparison in black and mauve, respectively. Peak and steady-state photocurrents in fusiform cells are outward at +53 mV, while stellate cell responses remain inward at all potentials. Traces from a Thy1-ChR2 mouse; similar results were observed in VGluT2-ChR2 mice. b) Average IV relationship for photocurrent responses in fusiform and stellate cells from ChR2 mice (n=10 cells each, solid and open symbols, respectively). Stellate cell responses were always inward, whereas fusiform cell photocurrents showed reversal. The positive reversal potential in fusiform cells may be due in part to electrically-coupled fusiform cells that are also depolarized during the light stimulus (Supplemental Figure 3). Moreover, the dendritic arbor of fusiform cells (see example cell in Figure 3, main text) may result in space-clamp limitations. c) Differential block of fusiform and stellate cell light responses by the gap junction blocker MFA. Average photocurrent responses in a simultaneously recorded (uncoupled) fusiform/stellate cell pair from a Thy1-ChR2 mouse during baseline (black) and after 200 μM MFA (red). The peak photocurrent in the fusiform cell was unaffected, whereas the stellate cell response was entirely abolished. MFA caused a small reduction in the steady-state fusiform cell response, and blocked the small outward current upon light offset. This may result from electrical coupling of adjacent fusiform cells activated during the light stimulus. d) Summary data showing the effects of 100-200 μM MFA on photocurrents in fusiform and stellate cells. The leftmost points show the fraction remaining in MFA (normalized to a pre-drug baseline) for the peak and steady-state amplitudes of the fusiform cell photocurrents. Black lines connect individual experiments, red points are mean±1SEM. MFA had no effect on peak photocurrents (98±1 4% remaining, n=7. t(6)=0.34, p=0.7, paired t-test). MFA slightly reduced the steady state photocurrent in 4/7 experiments, although this value was not statistically significant in grouped data (82±1 8% remaining. t(6)=2.09, p=0.08, paired t-test). The rightmost data points represent the fraction remaining of the steady-state postjunctional photocurrent in stellate cells. MFA completely abolished photocurrents in stellate cells (3±1 1% remaining, n=9. t(8)=3.93, p=0.004, paired t-test).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–4 (PDF 717 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Apostolides, P., Trussell, L. Regulation of interneuron excitability by gap junction coupling with principal cells. Nat Neurosci 16, 1764–1772 (2013). https://doi.org/10.1038/nn.3569

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3569

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