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:

Two distinct mechanisms for experience-dependent homeostasis

Abstract

Models of firing rate homeostasis such as synaptic scaling and the sliding synaptic plasticity modification threshold predict that decreasing neuronal activity (for example, by sensory deprivation) will enhance synaptic function. Manipulations of cortical activity during two forms of visual deprivation, dark exposure (DE) and binocular lid suture, revealed that, contrary to expectations, spontaneous firing in conjunction with loss of visual input is necessary to lower the threshold for Hebbian plasticity and increase miniature excitatory postsynaptic current (mEPSC) amplitude. Blocking activation of GluN2B receptors, which are upregulated by DE, also prevented the increase in mEPSC amplitude, suggesting that DE potentiates mEPSCs primarily through a Hebbian mechanism, not through synaptic scaling. Nevertheless, NMDA-receptor-independent changes in mEPSC amplitude consistent with synaptic scaling could be induced by extreme reductions of activity. Therefore, two distinct mechanisms operate within different ranges of neuronal activity to homeostatically regulate synaptic strength.

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

Fig. 1: Deprivation-induced metaplasticity does not depend on reduced firing rates.
Fig. 2: Deprivation-induced upregulation of mEPSC amplitude does not depend on reduced firing rates.
Fig. 3: Manipulations of PV-IN activity regulate the deprivation-induced increase in mEPSC amplitude.
Fig. 4: Spontaneous activity is required for the DE-mediated increase in GluN2B, which in turn is necessary for increased mEPSC amplitude.
Fig. 5: Synaptic scaling is engaged by extreme reductions in neuronal activity.
Fig. 6: Model depicting two distinct mechanisms that increase mEPSC amplitude in response to decreased neuronal activity.

Similar content being viewed by others

References

  1. Marder, E., O’Leary, T. & Shruti, S. Neuromodulation of circuits with variable parameters: single neurons and small circuits reveal principles of state-dependent and robust neuromodulation. Annu. Rev. Neurosci. 37, 329–346 (2014).

    Article  PubMed  CAS  Google Scholar 

  2. Davis, G. W. Homeostatic control of neural activity: from phenomenology to molecular design. Annu. Rev. Neurosci. 29, 307–323 (2006).

    Article  PubMed  CAS  Google Scholar 

  3. Turrigiano, G. G. The self-tuning neuron: synaptic scaling of excitatory synapses. Cell 135, 422–435 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Cooper, L. N. & Bear, M. F. The BCM theory of synapse modification at 30: interaction of theory with experiment. Nat. Rev. Neurosci. 13, 798–810 (2012).

    Article  PubMed  CAS  Google Scholar 

  5. Abraham, W. C. & Bear, M. F. Metaplasticity: the plasticity of synaptic plasticity. Trends Neurosci. 19, 126–130 (1996).

    Article  PubMed  CAS  Google Scholar 

  6. Abraham, W. C. Metaplasticity: tuning synapses and networks for plasticity. Nat. Rev. Neurosci. 9, 387–399 (2008).

    Article  PubMed  CAS  Google Scholar 

  7. Kirkwood, A., Rioult, M. C. & Bear, M. F. Experience-dependent modification of synaptic plasticity in visual cortex. Nature 381, 526–528 (1996).

    Article  PubMed  CAS  Google Scholar 

  8. Philpot, B. D., Espinosa, J. S. & Bear, M. F. Evidence for altered NMDA receptor function as a basis for metaplasticity in visual cortex. J. Neurosci. 23, 5583–5588 (2003).

    Article  PubMed  CAS  Google Scholar 

  9. Guo, Y. et al. Dark exposure extends the integration window for spike-timing-dependent plasticity. J. Neurosci. 32, 15027–15035 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Philpot, B. D., Sekhar, A. K., Shouval, H. Z. & Bear, M. F. Visual experience and deprivation bidirectionally modify the composition and function of NMDA receptors in visual cortex. Neuron 29, 157–169 (2001).

    Article  PubMed  CAS  Google Scholar 

  11. Chen, W. S. & Bear, M. F. Activity-dependent regulation of NR2B translation contributes to metaplasticity in mouse visual cortex. Neuropharmacology 52, 200–214 (2007).

    Article  PubMed  CAS  Google Scholar 

  12. He, H.-Y., Hodos, W. & Quinlan, E. M. Visual deprivation reactivates rapid ocular dominance plasticity in adult visual cortex. J. Neurosci. 26, 2951–2955 (2006).

    Article  PubMed  CAS  Google Scholar 

  13. Turrigiano, G. G., Leslie, K. R., Desai, N. S., Rutherford, L. C. & Nelson, S. B. Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature 391, 892–896 (1998).

    Article  PubMed  CAS  Google Scholar 

  14. Goel, A. & Lee, H. K. Persistence of experience-induced homeostatic synaptic plasticity through adulthood in superficial layers of mouse visual cortex. J. Neurosci. 27, 6692–6700 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. He, K., Petrus, E., Gammon, N. & Lee, H.-K. Distinct sensory requirements for unimodal and cross-modal homeostatic synaptic plasticity. J. Neurosci. 32, 8469–8474 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Gao, M. et al. A specific requirement of Arc/Arg3.1 for visual experience-induced homeostatic synaptic plasticity in mouse primary visual cortex. J. Neurosci. 30, 7168–7178 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Goel, A. et al. Phosphorylation of AMPA receptors is required for sensory deprivation-induced homeostatic synaptic plasticity. PLoS One 6, e18264 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Desai, N. S., Cudmore, R. H., Nelson, S. B. & Turrigiano, G. G. Critical periods for experience-dependent synaptic scaling in visual cortex. Nat. Neurosci. 5, 783–789 (2002).

    Article  PubMed  CAS  Google Scholar 

  19. Hengen, K. B., Lambo, M. E., Van Hooser, S. D., Katz, D. B. & Turrigiano, G. G. Firing rate homeostasis in visual cortex of freely behaving rodents. Neuron 80, 335–342 (2013).

    Article  PubMed  CAS  Google Scholar 

  20. Keck, T. et al. Synaptic scaling and homeostatic plasticity in the mouse visual cortex in vivo. Neuron 80, 327–334 (2013).

    Article  PubMed  CAS  Google Scholar 

  21. Kotak, V. C. et al. Hearing loss raises excitability in the auditory cortex. J. Neurosci. 25, 3908–3918 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Glazewski, S., Greenhill, S. & Fox, K. Time-course and mechanisms of homeostatic plasticity in layers 2/3 and 5 of the barrel cortex. Philos. Trans. R. Soc. Lond. B Biol. Sci. 372, 20160150 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Greenhill, S. D., Ranson, A. & Fox, K. Hebbian and homeostatic plasticity mechanisms in regular spiking and intrinsic bursting cells of cortical layer 5. Neuron 88, 539–552 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Fong, M. F., Newman, J. P., Potter, S. M. & Wenner, P. Upward synaptic scaling is dependent on neurotransmission rather than spiking. Nat. Commun. 6, 6339 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Barnes, S. J. et al. Deprivation-induced homeostatic spine scaling in vivo is localized to dendritic branches that have undergone recent spine loss. Neuron 96, 871–882.e5 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Gianfranceschi, L. et al. Visual cortex is rescued from the effects of dark rearing by overexpression of BDNF. Proc. Natl Acad. Sci. USA 100, 12486–12491 (2003).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Benevento, L. A., Bakkum, B. W., Port, J. D. & Cohen, R. S. The effects of dark-rearing on the electrophysiology of the rat visual cortex. Brain Res. 572, 198–207 (1992).

    Article  PubMed  CAS  Google Scholar 

  28. Gu, Y. et al. Neuregulin-dependent regulation of fast-spiking interneuron excitability controls the timing of the critical period. J. Neurosci. 36, 10285–10295 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Kuhlman, S. J. et al. A disinhibitory microcircuit initiates critical-period plasticity in the visual cortex. Nature 501, 543–546 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Huang, S., Hokenson, K., Bandyopadhyay, S., Russek, S. J. & Kirkwood, A. Brief dark exposure reduces tonic inhibition in visual cortex. J. Neurosci. 35, 15916–15920 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Hensch, T. K. et al. Local GABA circuit control of experience-dependent plasticity in developing visual cortex. Science 282, 1504–1508 (1998).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Gu, Y. et al. Obligatory role for the immediate early gene NARP in critical period plasticity. Neuron 79, 335–346 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Greenblatt, D. J. & Sethy, V. H. Benzodiazepine concentrations in brain directly reflect receptor occupancy: studies of diazepam, lorazepam, and oxazepam. Psychopharmacology (Berl.) 102, 373–378 (1990).

    Article  CAS  Google Scholar 

  34. Rittenhouse, C. D. et al. Stimulus for rapid ocular dominance plasticity in visual cortex. J. Neurophysiol. 95, 2947–2950 (2006).

    Article  PubMed  Google Scholar 

  35. Rittenhouse, C. D., Shouval, H. Z., Paradiso, M. A. & Bear, M. F. Monocular deprivation induces homosynaptic long-term depression in visual cortex. Nature 397, 347–350 (1999).

    Article  PubMed  CAS  Google Scholar 

  36. Frenkel, M. Y. & Bear, M. F. How monocular deprivation shifts ocular dominance in visual cortex of young mice. Neuron 44, 917–923 (2004).

    Article  PubMed  CAS  Google Scholar 

  37. Christian, C. A. et al. Endogenous positive allosteric modulation of GABAA receptors by diazepam binding inhibitor. Neuron 78, 1063–1074 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Wen, L. et al. Neuregulin 1 regulates pyramidal neuron activity via ErbB4 in parvalbumin-positive interneurons. Proc. Natl Acad. Sci. USA 107, 1211–1216 (2010).

    Article  PubMed  Google Scholar 

  39. Cho, K. K. A., Khibnik, L., Philpot, B. D. & Bear, M. F. The ratio of NR2A/B NMDA receptor subunits determines the qualities of ocular dominance plasticity in visual cortex. Proc. Natl Acad. Sci. USA 106, 5377–5382 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Drasbek, K. R. & Jensen, K. THIP, a hypnotic and antinociceptive drug, enhances an extrasynaptic GABAA receptor-mediated conductance in mouse neocortex. Cereb. Cortex 16, 1134–1141 (2006).

    Article  PubMed  Google Scholar 

  41. Sun, Y. et al. Neuregulin-1/ErbB4 signaling regulates visual cortical plasticity. Neuron 92, 160–173 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Pratt, K. G. & Aizenman, C. D. Homeostatic regulation of intrinsic excitability and synaptic transmission in a developing visual circuit. J. Neurosci. 27, 8268–8277 (2007).

    Article  PubMed  CAS  Google Scholar 

  43. Gambrill, A. C., Storey, G. P. & Barria, A. Dynamic regulation of NMDA receptor transmission. J. Neurophysiol. 105, 162–171 (2011).

    Article  PubMed  CAS  Google Scholar 

  44. Lee, K. F. H., Soares, C. & Béïque, J.-C. Tuning into diversity of homeostatic synaptic plasticity. Neuropharmacology 78, 31–37 (2014).

    Article  PubMed  CAS  Google Scholar 

  45. Petrus, E. et al. Vision loss shifts the balance of feedforward and intracortical circuits in opposite directions in mouse primary auditory and visual cortices. J. Neurosci. 35, 8790–8801 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Kaneko, M., Stellwagen, D., Malenka, R. C. & Stryker, M. P. Tumor necrosis factor-alpha mediates one component of competitive, experience-dependent plasticity in developing visual cortex. Neuron 58, 673–680 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Ranson, A., Cheetham, C. E. J., Fox, K. & Sengpiel, F. Homeostatic plasticity mechanisms are required for juvenile, but not adult, ocular dominance plasticity. Proc. Natl Acad. Sci. USA 109, 1311–1316 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Deeg, K. E. & Aizenman, C. D. Sensory modality-specific homeostatic plasticity in the developing optic tectum. Nat. Neurosci. 14, 548–550 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Frémaux, N. & Gerstner, W. Neuromodulated spike-timing-dependent plasticity, and theory of three-factor learning rules. Front. Neural Circuits 9, 85 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Zenke, F., Agnes, E. J. & Gerstner, W. Diverse synaptic plasticity mechanisms orchestrated to form and retrieve memories in spiking neural networks. Nat. Commun. 6, 6922 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Tallaksen-Greene, S. J., Janiszewska, A., Benton, K., Ruprecht, L. & Albin, R. L. Lack of efficacy of NMDA receptor-NR2B selective antagonists in the R6/2 model of Huntington disease. Exp. Neurol. 225, 402–407 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Rumbaugh, G. & Vicini, S. Distinct synaptic and extrasynaptic NMDA receptors in developing cerebellar granule neurons. J. Neurosci. 19, 10603–10610 (1999).

    Article  PubMed  CAS  Google Scholar 

  53. Murase, S. et al. Matrix metalloproteinase-9 regulates neuronal circuit development and excitability. Mol. Neurobiol. 53, 3477–3493 (2016).

    Article  PubMed  CAS  Google Scholar 

  54. Niell, C. M. & Stryker, M. P. Highly selective receptive fields in mouse visual cortex. J. Neurosci. 28, 7520–7536 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Chen, L., Deng, Y., Luo, W., Wang, Z. & Zeng, S. Detection of bursts in neuronal spike trains by the mean inter-spike interval method. Prog. Nat. Sci. 19, 229–235 (2009).

    Article  Google Scholar 

Download references

Acknowledgements

Research reported in this article was supported by the National Eye Institute of the National Institutes of Health under award number R01EY012124 (to A.K.), R01EY016431 (to E.Q.), R01EY025922 (to E.Q., A.K. and H.-K. L.) and R01-EY014882 (to H.-K.L.). A.K. was also supported by NIH grant P01 AG009973. M.C.D.B. was supported by grants T32EY007143 and T32HL110952.

Author information

Authors and Affiliations

Authors

Contributions

M.C.D.B., R.d.P., T.T., K.H. and S.Z.H. collected slice electrophysiology data. M.C.D.B. and R.d.P. analyzed slice electrophysiology data. S.-Y.C. provided initial data for the inception of the project. C.L.L., Y.G. and A.B. collected and analyzed the in vivo unit recording data. A.D. and H.-K.L. provided code for analysis of single-unit firing properties. M.C.D.B., E.M.Q. and A.K. wrote the manuscript.

Corresponding authors

Correspondence to Elizabeth M. Quinlan or Alfredo Kirkwood.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Single-unit recordings in awake, head-fixed animals.

(A,B) Representative waveforms from units recorded before and after DE (A) and BS (B) (see Figure 1A, D). Thin lines represent individual waveforms; black lines represent the average waveform. (C) Single unit firing rates in naïve mice at baseline and following DE with diazepam treatment. There was no significant difference between groups (P=0.676, 2-tailed Wilcoxon rank sum test), confirming that diazepam prevents the DE-mediated increase in spontaneous firing (see Figure 1A). Data are shown as individual points; dashed lines and error bars represent mean±SEM. Sample size is indicated in parentheses as (cells, mice)

Supplementary Figure 2 Effects of visual deprivation and modulation of inhibition on single-unit firing properties.

(A) Spontaneous bursting increased after DE and after BS with flumazenil treatment. *P≤0.001 (U(60)=211.0), #P=0.047 (U(60)=235.0), †P=0.006 (U(64)=328.0), ‡P=0.013 (U(49)=193.0), 2-tailed Wilcoxon rank sum test. Sample size is shown as (units, mice). (B) Non-burst firing rates increased after DE. *P≤0.001 (U(60)=235.0), 2-tailed Wilcoxon rank sum test. No significant differences were observed between the recovery and DE+diazepam (U(60)=363.0, P=0.100), baseline and BS (U(64)=403.0, P=0.071), or recovery and BS+flumazenil (U(49)=238, P=0.103) conditions. Sample size as shown in panel (A). Data are shown as individual points; dashed lines and error bars represent mean±SEM. (C) Raster plots of two example units from awake, head-fixed mice viewing a grey screen. Spikes were classified as non-burst (blue) or part of a burst (magenta) (see Methods). A portion of the raster plots is enlarged (below) to show more detail

Supplementary Figure 3 Effects of acute drug treatment on spontaneous single-unit firing properties.

(A) Single unit recordings were obtained during presentation of a grey screen in awake, head-fixed mice. V1 firing rates were measured at baseline (open circles) and 20 minutes after drug administration i.p. (filled circles). Diazepam and THIP decreased, and flumazenil increased, spontaneous firing rate. *P≤0.001, #P=0.002, 2-tailed Wilcoxon signed rank test. Vehicle injection did not alter spontaneous firing rate (Z(15)=0.646, P=0.528). Sample size is indicated as (units, mice). (B) To demonstrate the long-lasting effects of a single drug injection i.p., single unit recordings were made from V1 under isoflurane anesthesia. Diazepam decreased and flumazenil increased (4 units, 4 mice per group) spontaneous neuronal firing rates for at least 2h after the injection. The dotted gray line indicates baseline firing rate for reference. *P≤0.001 vs. −30 minute time point, One-way repeated measures ANOVA with Holm-Sidak post-hoc test. Sample size is shown as (units, mice). (C,D) Single unit firing properties during acute drug administration. Diazepam and THIP decreased, and flumazenil increased, both spontaneous bursting (C; 2-tailed paired t test) and the inter-burst firing rate (D, 2-tailed Wilcoxon signed rank test). *P≤0.001, #P=0.002. Vehicle injection did not alter these parameters (bursting, 2-tailed paired t test t(15)=−0.313, P=0.759; inter-burst firing rate, 2-tailed Wilcoxon signed rank test Z(15)=0.267, P=0.831. Sample size as shown in (A). For all panels, individual unit firing parameters at each time point are connected by grey lines. Colored symbols and lines indicate the mean and error bars represent SEM

Supplementary Figure 4 Diazepam administered i.c.v. prevents the DE-mediated increase in mEPSC amplitude.

(A) Average mEPSC traces after diazepam or vehicle infusion directly into the lateral ventricle via osmotic minipump during DE. (B) mEPSC amplitude was larger in vehicle- than diazepam-infused animals. *P=0.013, 2-tailed t test. Dashed lines and error bars indicate mean±SEM. For clarity, some data points are displaced horizontally. Sample size is shown as (cells, mice)

Supplementary Figure 5 Activation of Gi-DREADD in PV cells elevates spontaneous activity in neighboring regular-spiking neurons.

(A) CNO (5 mg/kg i.p.) increased the firing rate of regular spiking neurons in hemispheres where Gi-DREADD was expressed in PV cells. *P≤0.001, 2-tailed paired t test. Data are from 31 units recorded from 3 awake, head-fixed animals viewing a grey screen. Grey lines represent individual units before and after CNO administration; green line and error bars represent mean ±SEM. (B) For mEPSC recordings (Fig. 3C, D), patch electrodes were targeted to pyramidal cells neighboring mCherry-labeled, Gi-DREADD expressing PV cells

Supplementary Figure 6 Diazepam administered i.c.v. prevents the DE-mediated increase in GluN2B function.

(A) DE increased the percentage of NMDAR EPSC blocked by ifenprodil (*P=0.021, 2-tailed t test) and the NMDA receptor decay constant (τw) (†P=0.002, 2-tailed t test) in animals with no drug administration. (B) In DE animals, diazepam prevented the GluN2B increase compared to vehicle (*P=0.039; †P=0.030; 2-tailed t test). Solid lines: average baseline NMDAR current; dashed lines: after ifenprodil wash-in. Traces are normalized to baseline. Data are displayed as individual points; dashed lines indicate average (±SEM); sample size is shown as (neurons, mice)

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 and Supplementary Tables 1–5

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bridi, M.C.D., de Pasquale, R., Lantz, C.L. et al. Two distinct mechanisms for experience-dependent homeostasis. Nat Neurosci 21, 843–850 (2018). https://doi.org/10.1038/s41593-018-0150-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0150-0

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