Skip to main content

Homeostatic plasticity in neural development

Abstract

Throughout life, neural circuits change their connectivity, especially during development, when neurons frequently extend and retract dendrites and axons, and form and eliminate synapses. In spite of their changing connectivity, neural circuits maintain relatively constant activity levels. Neural circuits achieve functional stability by homeostatic plasticity, which equipoises intrinsic excitability and synaptic strength, balances network excitation and inhibition, and coordinates changes in circuit connectivity. Here, we review how diverse mechanisms of homeostatic plasticity stabilize activity in developing neural circuits.

Background

Nervous systems face a constant challenge: how to maintain flexibility and stability at the same time. Neural circuits must stay flexible to allow for changes in connectivity and synaptic strength during development and learning. As changes in connectivity push neural circuits away from equilibrium, they need to maintain activity within a working range and avoid extremes of quiescence and saturation. Functional stability is maintained by homeostatic plasticity, which is defined broadly as a set of neuronal changes that restore activity to a setpoint following perturbation [1,2,3]. Recent studies have identified diverse homeostatic plasticity mechanisms triggered by a variety of perturbations. These mechanisms regulate dendritic and axonal connectivity of a neuron, as well as its intrinsic excitability (Fig. 1). In addition to maintaining the activity of individual neurons, homeostatic plasticity can act at a network level to coordinate changes in connectivity and excitability across multiple neurons to stabilize circuit function [4] (Fig. 2). Several recent reviews have covered the function of homeostatic plasticity in the mature nervous system [5,6,7,8]. Here, we focus on homeostatic plasticity in developing circuits.

Fig. 1
figure 1

Diverse homeostatic plasticity mechanisms stabilize the activity of developing neurons. When the activity of individual neurons decreases below (1 and 2) or increases above (3 and 4) a setpoint, homeostatic regulation of synaptic strength (1 and 3) and/or intrinsic excitability (2 and 4) acts to restore normal activity. By increasing (1) or decreasing (3) synaptic input (e.g., changes in mEPSC amplitude or frequency), a neuron’s output firing rate can be shifted up or down to the target activity (grey area). By increasing (2) or decreasing (4) intrinsic excitability (e.g., changes in the length and location of AIS), a neuron’s input/output relationship can be modified

Fig. 2
figure 2

Network-level homeostatic plasticity stabilizes activity of developing circuits. Network activity homeostasis is achieved by balancing excitation (red) and inhibition (blue). Synaptic strength and connectivity can be regulated in a cell-type-specific manner to maintain network homeostasis. Upward/downward red arrows: increased/decreased excitatory drive; upward/downward blue arrows: increased/decreased inhibitory drive

Homeostatic regulation of intrinsic excitability

Neuronal intrinsic excitability is determined by the density, distribution, and function of ion channels, and controls how synaptic inputs are converted into action potential outputs [9]. Several studies have found a reciprocal relationship between intrinsic excitability and synaptic inputs across development, which stabilizes activity [10,11,12]. As synaptic inputs increase in developing Xenopus retinotectal circuits, Na+ currents decrease, reducing intrinsic excitability [12]. Conversely, silencing synaptic inputs to developing Xenopus tectal neurons and Drosophila motorneurons increases Na+ currents and intrinsic excitability [10, 12, 13]. Several mechanisms mediate homeostatic changes in Na+ currents. Translational repression and post-translational phosphorylation reduce the density and open probability, respectively, of voltage-gated Na+ channels in Drosophila motorneurons and rat cortical neurons in response to elevated synaptic activity [11, 14,15,16,17].

Multiple ion channels in the same neuron can balance each other to stabilize activity [2, 18, 19]. For example, the A-type K+ channels shal and shaker are reciprocally regulated in motorneurons of Drosophila larvae: shaker is up-regulated in shal mutants, and shal is up-regulated in shaker mutants [20]. However, compensatory expression is not always a two-way street; in Drosophila mutants of the delayed rectifier K+ channel shab, increased expression of the Ca2+-dependent K+ channel slo prevents motorneuron hyperactivity, but, loss of slo does not increase expression of shab [21]. Neurons can synergistically regulate ion channels with opposite effects on excitability to restore activity. Silencing of pyramidal neurons cultured from visual cortex of rat pups with TTX increases Na+ currents and decreases K+ currents [22]. Finally, neurons of the same type with similar excitability can vary significantly in their membrane conductances, which may reflect the complex homeostatic interactions between ion channels [23,24,25] (for more discussion, see [26, 27]).

Detailed examination of the distribution of ion channels revealed an important role of the axon-initial-segment (AIS) in intrinsic homeostatic plasticity. Changes in length and location of the AIS, a specialized region with clusters of voltage-gated Na+ and K+ channels involved in spike generation, can counter the effects of sensory deprivation or photostimulation [28,29,30,31]. In mice, eye opening at postnatal day 13–14 shortens the AIS of pyramidal neurons in visual cortex [32, 33]. Together, adjustments in ion channel density, distribution, and function, resulting from changes in transcription, translation, post-translational modifications, and trafficking, can alter intrinsic excitability and balance changes in synaptic input to maintain activity homeostasis [9, 34,35,36].

Homeostatic regulation of synapse strength and number

Homeostatic plasticity can regulate synaptic strength pre- and postsynaptically, and its dominant expression site can shift during development. In the early stages of network formation, miniature excitatory postsynaptic current (mEPSC) amplitudes increase when spike generation is blocked in cortical and hippocampal neuron cultures (i.e., suppression of intrinsic excitability), indicative of postsynaptic changes in AMPA receptor accumulation [37]. At later stages, presynaptic regulation of vesicle release and recycling is added, and mEPSC frequencies increase along with mEPSC amplitudes when spike generation is blocked [37, 38]. This suggests a developmental shift in the capacity for pre- and postsynaptic homeostatic plasticity [37]. Homeostatic control of synaptic strength has also been observed in vivo [39, 40]. The extent and expression site of this control depends on circuit maturation [41,42,43,44,45]. Homeostatic synaptic plasticity in layers 4 and 6 of primary visual cortex elicited by visual deprivation is restricted to an early critical period (postnatal day 16 to 21) [42, 43]. Later, homeostatic regulation of mEPSC amplitudes shifts to layers 2/3, where it persists into adulthood [42, 44]. The purpose of this shift in homeostatic plasticity across cortical layers remains unknown [41]. Chronic activity suppression by intracranial infusion of the Na+ channel blocker TTX or NMDA receptor blockers increases spine densities of developing thalamocortical neurons in the dorsolateral geniculate nucleus of cats and ferrets [46, 47]. Thus, homeostatic plasticity can regulate synapse number as well as strength [48,49,50].

In addition to homeostatic synaptic changes elicited by experimental perturbations, Desai et al. showed that across development, mEPSC amplitudes in layers 2/3 and 4 of rat primary visual cortex decrease as mEPSC frequencies and synapse numbers increase [42]. Retinogeniculate circuits provide another example of developmental homeostatic co-regulation [51,52,53]. Initially, many retinal ganglion cells converge onto thalamocortical cells, each forming weak connections. Then, for up to 3 weeks after eye opening, thalamocortical cells prune inputs, retaining synapses from fewer ganglion cells, which strengthen their connections [53, 54]. Thus, presynaptic neurotransmitter release, postsynaptic receptor abundance, and synapse number are homeostatically co-regulated during normal development and after activity perturbations. In several systems, the expression sites and the combination of mechanisms engaged shift across development [2, 3, 55,56,57].

Homeostatic regulation of network activity

Homeostatic plasticity can stabilize the activity of individual neurons [54, 58, 59]. Neurons connect to each other in a cell-type-specific manner, forming circuits that perform specific functions. In the following sections, we discuss how homeostatic mechanisms are coordinated across neurons to stabilize circuit function [4, 60].

Homeostatic regulation of network excitation and inhibition

Network activity is determined by the ratio of excitation and inhibition (E/I ratio) [1, 4, 61]. In response to perturbations, developing circuits can differentially adjust inhibitory and excitatory connectivity to alter the E/I ratio and restore activity [62,63,64,65]. In developing hippocampal and organotypic cerebellar cultures, TTX or glutamate receptor antagonists decrease inhibitory synapse densities and strengths, whereas blocking GABAergic transmission with bicuculline increases the density of inhibitory synapses. Similarly, brain slice recordings in barrel cortex layer 4 showed that sensory deprivation selectively reduces inhibitory input to layer 4 spiny neurons in young but not in adult animals [66, 67]. Activity-dependent changes in inhibitory synaptic transmission appear to be regulated non-cell autonomously, as activity suppression of individual presynaptic or postsynaptic cells failed to elicit compensatory changes observed after global application of TTX in neonatal cultured hippocampal neurons [65]. It has been suggested that inhibitory interneurons may sacrifice their own firing rate homeostasis to stabilize spiking of cortical pyramidal neurons after global activity blockade [4, 68]. Another example of network homeostasis comes from studies of monocular deprivation during the critical period [4]. Here, homeostatic plasticity adjusts recurrent and feedforward connections between layer 4 circuits and layer 2/3 circuits in primary visual cortex. Visual deprivation via intraocular TTX injection increases the excitatory drive and reduces inhibitory drive from layer 4 to layer 2/3, compensating for the lost excitatory sensory input [4, 69, 70]. Intriguingly, in another deprivation paradigm (i.e., lid suture), increased intrinsic excitability and decreased E/I ratios stabilize activity in layer 2/3, indicating the same circuit can use different combinations of homeostatic mechanisms to compensate for sensory deprivation.

In addition to regulating excitatory and inhibitory synapse strength and number, homeostatic plasticity can switch the transmitter phenotype of neurons from glutamate to GABA or vice versa to adjust the E/I ratio of developing circuits [71,72,73]. In the embryonic Xenopus spinal cord, the fractions of neurons expressing excitatory transmitters increase and decrease, respectively, when network activity is pharmacologically suppressed and enhanced. These switches in transmitter phenotype occur without changes in the expression of cell identity markers [74]. Similar to homeostatic regulation of inhibitory synapses, the activity-dependent transmitter switch is non-cell autonomous and depends on network activity, evidenced by the reciprocal relationship between the number of silenced cells and the ratio of neurons expressing GABA vs. glutamate [75]. Whether switches in transmitter phenotypes contribute to network homeostasis during normal development remains to be investigated [71].

Homeostatic regulation of cell-type-specific connectivity

Recent advances in single-cell RNA sequencing together with large-scale morphological and functional surveys have revealed a great diversity of excitatory and inhibitory cell types, which serve distinct circuit functions [76,77,78,79]. This raises the questions whether, beyond categorical differences between excitatory and inhibitory neurons, homeostatic plasticity may act in a cell-type-specific manner to stabilize circuit function [80]. In the developing dentate gyrus, loss of excitatory drive by tetanus toxin expression results in reduced inhibitory input to granule cells [81]. This reduction is cell-type specific, affecting somatic innervation by parvalbumin-positive basket cells, but not dendritic innervation by calretinin- and somatostatin-expressing interneurons. Selective reduction of somatic inhibition efficiently restores the firing of granule cells [82, 83]. Similarly, monocular deprivation during a pre-critical period was shown to regulate feedback but not feedforward inhibition to layer 4 pyramidal cells in rat primary visual cortex [84]; and early hearing loss weakens inhibitory synapses from fast-spiking interneurons but not from low-threshold spiking interneurons onto pyramidal cells [85, 86].

Homeostatic regulation of excitatory connectivity can also be cell type specific [87]. In the developing mouse retina, following removal of their dominant B6 bipolar cell input, ONα retinal ganglion cells up-regulate connectivity with XBC, B7, and rod bipolar cells, but leave input from B8 bipolar cells unchanged. This cell-type-specific rewiring not only maintains the sustained activity of ONα retinal ganglion cells, but also precisely preserves their light responses. Thus, homeostatic plasticity can regulate inhibitory and excitatory connectivity in a cell-type-specific manner to maintain the activity and sensory function of developing circuits.

Homeostatic regulation of patterned spontaneous activity

Throughout the nervous system, developing circuits spontaneously generate activity patterns that help refine their connectivity [88, 89]. Before eye opening, waves of activity originating in the retina propagate through the visual system and dominate activity up to primary visual cortex [90,91,92]. Retinal waves mature in three stages (I-III), in which different circuit mechanisms generate distinct activity patterns that serve specific functions in visual system refinement [88]. In mice, stage I waves, which are mediated by gap-junctional coupling of retinal ganglion cells, were first observed at embryonic day 17. Around birth, the wave generation switches to networks of cholinergic amacrine cells (stage II, postnatal day 1–10) followed in the second postnatal week by glutamatergic input from bipolar cells (stage III, postnatal day 10–14). The transitions between stages appear to be homeostatically regulated. When stage II (i.e., cholinergic) waves are disrupted by genetic deletion or pharmacological blockade of ß2 nicotinic acetylcholine receptors nAChRs, stage I waves persist until premature stage III waves take over [93,94,95,96]. Similarly, in VGluT1 knockout mice, in which stage III waves are abolished, stage II waves persist until eye opening [97]. Studies of developing spinal networks revealed an important role of excitatory GABAergic currents in homeostatic regulation of patterned spontaneous activity [98]. During development, GABA switches from excitatory to inhibitory as initially high intracellular Cl concentrations are lowered by the developmentally regulated expression of cation-chloride cotransporters [99, 100]. When spontaneous network activity in chick embryos was reduced by injection of a sodium channel blocker, excitatory GABAergic mEPSC amplitudes were found to increase because of an increased Cl driving force due to intracellular Cl accumulation [101, 102].

Although homeostatic mechanisms can restore spontaneous activity patterns following perturbations, the extent to which these activity patterns support normal circuit refinement varies depending on age and means of perturbation and needs to be further investigated [103,104,105].

Conclusions

Developing circuits undergo profound changes in connectivity that threaten to destabilize their activity. Recent research has revealed a diverse set of homeostatic plasticity mechanisms, which safeguard activity of developing circuits. Different combinations of these mechanisms are recruited by different perturbations in different neuronal cell types at different stages of development. What signals control the recruitment of specific combinations of mechanisms is unclear and an interesting topic for future studies [41, 55].

Another important and mostly unanswered question is how activity setpoints are determined [2, 106,107,108]. Recent evidence suggests that this may occur during specific critical periods of development [109, 110]. Altering network activity in wild-type Drosophila during a critical period induces subsequent seizures, whereas correcting abnormal activity in mutant flies during the same period is sufficient to suppress seizures for life. Importantly, in the seizure-prone flies, homeostatic plasticity mechanisms are intact, but working toward the “wrong” setpoints. Insights into critical period timing and determinants of activity setpoints could have significant implications for the treatment of neurodevelopmental diseases including epilepsy and autisms [111,112,113,114].

Abbreviations

AIS:

Axon-initial-segment

E/I:

Excitation/Inhibition

mEPSC:

miniature excitatory postsynaptic current

References

  1. Turrigiano GG, Nelson SB. Homeostatic plasticity in the developing nervous system. Nat Rev Neurosci. 2004;5(2):97–107.

    Article  PubMed  CAS  Google Scholar 

  2. Davis GW. Homeostatic signaling and the stabilization of neural function. Neuron. 2013;80(3):718–28.

    Article  PubMed  CAS  Google Scholar 

  3. Wefelmeyer W, Puhl CJ, Burrone J. Homeostatic plasticity of subcellular neuronal structures: from inputs to outputs. Trends Neurosci. 2016;39(10):656–67.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Maffei A, Fontanini A. Network homeostasis: a matter of coordination. Curr Opin Neurobiol. 2009;19(2):168–73.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  5. Fox K, Stryker M. Integrating Hebbian and homeostatic plasticity: introduction. Philos Trans R Soc Lond Ser B Biol Sci. 2017;372(1715). https://doi.org/10.1098/rstb.2016.0413.

  6. Turrigiano GG. The dialectic of Hebb and homeostasis. Philos Trans R Soc Lond Ser B Biol Sci. 2017;372(1715). https://doi.org/10.1098/rstb.2016.0258.

  7. Turrigiano GG. The self-tuning neuron: synaptic scaling of excitatory synapses. Cell. 2008;135(3):422–35.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Vitureira N, Goda Y. Cell biology in neuroscience: the interplay between Hebbian and homeostatic synaptic plasticity. J Cell Biol. 2013;203(2):175–86.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Beck H, Yaari Y. Plasticity of intrinsic neuronal properties in CNS disorders. Nat Rev Neurosci. 2008;9(5):357–69.

    Article  PubMed  CAS  Google Scholar 

  10. Baines RA, Uhler JP, Thompson A, Sweeney ST, Bate M. Altered electrical properties in Drosophila neurons developing without synaptic transmission. J Neurosci. 2001;21(5):1523–31.

    Article  PubMed  CAS  Google Scholar 

  11. Baines RA. Postsynaptic protein kinase a reduces neuronal excitability in response to increased synaptic excitation in the Drosophila CNS. J Neurosci. 2003;23(25):8664–72.

    Article  PubMed  CAS  Google Scholar 

  12. Pratt KG, Aizenman CD. Homeostatic regulation of intrinsic excitability and synaptic transmission in a developing visual circuit. J Neurosci. 2007;27(31):8268–77.

    Article  PubMed  CAS  Google Scholar 

  13. Hamodi AS, Pratt KG. Region-specific regulation of voltage-gated intrinsic currents in the developing optic tectum of the Xenopus tadpole. J Neurophysiol. 2014;112(7):1644–55.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Mee CJ, Pym EC, Moffat KG, Baines RA. Regulation of neuronal excitability through pumilio-dependent control of a sodium channel gene. J Neurosci. 2004;24(40):8695–703.

    Article  PubMed  CAS  Google Scholar 

  15. Muraro NI, Weston AJ, Gerber AP, Luschnig S, Moffat KG, Baines RA. Pumilio binds Para mRNA and requires Nanos and brat to regulate sodium current in Drosophila motoneurons. J Neurosci. 2008;28(9):2099–109.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Driscoll HE, Muraro NI, He M, Baines RA. Pumilio-2 regulates translation of Nav1.6 to mediate homeostasis of membrane excitability. J Neurosci. 2013;33(23):9644–54.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Scheuer T. Regulation of sodium channel activity by phosphorylation. Semin Cell Dev Biol. 2011;22(2):160–5.

    Article  PubMed  CAS  Google Scholar 

  18. Turrigiano G, LeMasson G, Marder E. Selective regulation of current densities underlies spontaneous changes in the activity of cultured neurons. J Neurosci. 1995;15(5 Pt 1):3640–52.

    Article  PubMed  CAS  Google Scholar 

  19. Marder E, Prinz AA. Modeling stability in neuron and network function: the role of activity in homeostasis. Bioessays. 2002;24(12):1145–54.

    Article  PubMed  CAS  Google Scholar 

  20. Bergquist S, Dickman DK, Davis GW. A hierarchy of cell intrinsic and target-derived homeostatic signaling. Neuron. 2010;66(2):220–34.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Kim EZ, Vienne J, Rosbash M, Griffith LC. Nonreciprocal homeostatic compensation in Drosophila potassium channel mutants. J Neurophysiol. 2017;117(6):2125–36.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  22. Desai NS, Rutherford LC, Turrigiano GG. Plasticity in the intrinsic excitability of cortical pyramidal neurons. Nat Neurosci. 1999;2(6):515–20.

    Article  PubMed  CAS  Google Scholar 

  23. Schulz DJ, Goaillard JM, Marder E. Variable channel expression in identified single and electrically coupled neurons in different animals. Nat Neurosci. 2006;9(3):356–62.

    Article  PubMed  CAS  Google Scholar 

  24. Golowasch J, Goldman MS, Abbott LF, Marder E. Failure of averaging in the construction of a conductance-based neuron model. J Neurophysiol. 2002;87(2):1129–31.

    Article  PubMed  Google Scholar 

  25. Schulz DJ. Plasticity and stability in neuronal output via changes in intrinsic excitability: it's what's inside that counts. J Exp Biol. 2006;209(Pt 24):4821–7.

    Article  PubMed  Google Scholar 

  26. O'Leary T, Williams AH, Caplan JS, Marder E. Correlations in ion channel expression emerge from homeostatic tuning rules. Proc Natl Acad Sci U S A. 2013;110(28):E2645–54.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Hudson AE, Prinz AA. Conductance ratios and cellular identity. PLoS Comput Biol. 2010;6(7):e1000838.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Grubb MS, Burrone J. Activity-dependent relocation of the axon initial segment fine-tunes neuronal excitability. Nature. 2010;465(7301):1070–4.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Kuba H, Oichi Y, Ohmori H. Presynaptic activity regulates Na(+) channel distribution at the axon initial segment. Nature. 2010;465(7301):1075–8.

    Article  PubMed  CAS  Google Scholar 

  30. Evans MD, Dumitrescu AS, Kruijssen DLH, Taylor SE, Grubb MS. Rapid modulation of axon initial segment length influences repetitive spike firing. Cell Rep. 2015;13(6):1233–45.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Yin J, Yuan Q. Structural homeostasis in the nervous system: a balancing act for wiring plasticity and stability. Front Cell Neurosci. 2014;8:439.

    PubMed  Google Scholar 

  32. Gutzmann A, Ergul N, Grossmann R, Schultz C, Wahle P, Engelhardt M. A period of structural plasticity at the axon initial segment in developing visual cortex. Front Neuroanat. 2014;8:11.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Petersen AV, Cotel F, Perrier JF. Plasticity of the axon initial segment: fast and slow processes with multiple functional roles. Neuroscientist. 2016;23(4):364–73.

  34. Temporal S, Lett KM, Schulz DJ. Activity-dependent feedback regulates correlated ion channel mRNA levels in single identified motor neurons. Current biology : CB. 2014;24(16):1899–904.

    Article  PubMed  CAS  Google Scholar 

  35. Zhang W, Linden DJ. The other side of the engram: experience-driven changes in neuronal intrinsic excitability. Nat Rev Neurosci. 2003;4(11):885–900.

    Article  PubMed  CAS  Google Scholar 

  36. Frick A, Johnston D. Plasticity of dendritic excitability. J Neurobiol. 2005;64(1):100–15.

    Article  PubMed  CAS  Google Scholar 

  37. Wierenga CJ, Walsh MF, Turrigiano GG. Temporal regulation of the expression locus of homeostatic plasticity. J Neurophysiol. 2006;96(4):2127–33.

    Article  PubMed  CAS  Google Scholar 

  38. Burrone J, O'Byrne M, Murthy VN. Multiple forms of synaptic plasticity triggered by selective suppression of activity in individual neurons. Nature. 2002;420(6914):414–8.

    Article  PubMed  CAS  Google Scholar 

  39. Kaneko M, Stryker MP. Homeostatic plasticity mechanisms in mouse V1. Philos Trans R Soc Lond Ser B Biol Sci. 2017;372(1715). https://doi.org/10.1098/rstb.2016.0504.

  40. Johnson RE, Tien NW, Shen N, Pearson JT, Soto F, Kerschensteiner D. Homeostatic plasticity shapes the visual system's first synapse. Nat Commun. 2017;8(1):1220.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Turrigiano G. Homeostatic synaptic plasticity: local and global mechanisms for stabilizing neuronal function. Cold Spring Harb Perspect Biol. 2012;4(1):a005736.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Desai NS, Cudmore RH, Nelson SB, Turrigiano GG. Critical periods for experience-dependent synaptic scaling in visual cortex. Nat Neurosci. 2002;5(8):783–9.

    Article  PubMed  CAS  Google Scholar 

  43. Petrus E, Anguh TT, Pho H, Lee A, Gammon N, Lee HK. Developmental switch in the polarity of experience-dependent synaptic changes in layer 6 of mouse visual cortex. J Neurophysiol. 2011;106(5):2499–505.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Goel A, Lee HK. Persistence of experience-induced homeostatic synaptic plasticity through adulthood in superficial layers of mouse visual cortex. J Neurosci. 2007;27(25):6692–700.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Espinosa JS, Stryker MP. Development and plasticity of the primary visual cortex. Neuron. 2012;75(2):230–49.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Dalva MB, Ghosh A, Shatz CJ. Independent control of dendritic and axonal form in the developing lateral geniculate nucleus. J Neurosci. 1994;14(6):3588–602.

    Article  PubMed  CAS  Google Scholar 

  47. Rocha M, Sur M. Rapid acquisition of dendritic spines by visual thalamic neurons after blockade of N-methyl-D-aspartate receptors. Proc Natl Acad Sci U S A. 1995;92(17):8026–30.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Pak DT, Sheng M. Targeted protein degradation and synapse remodeling by an inducible protein kinase. Science. 2003;302(5649):1368–73.

    Article  PubMed  CAS  Google Scholar 

  49. Seeburg DP, Sheng M. Activity-induced polo-like kinase 2 is required for homeostatic plasticity of hippocampal neurons during epileptiform activity. J Neurosci. 2008;28(26):6583–91.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Lee KJ, Lee Y, Rozeboom A, Lee JY, Udagawa N, Hoe HS, Pak DT. Requirement for Plk2 in orchestrated ras and rap signaling, homeostatic structural plasticity, and memory. Neuron. 2011;69(5):957–73.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Chen C, Regehr WG. Developmental remodeling of the retinogeniculate synapse. Neuron. 2000;28(3):955–66.

    Article  PubMed  CAS  Google Scholar 

  52. Jaubert-Miazza L, Green E, Lo FS, Bui K, Mills J, Guido W. Structural and functional composition of the developing retinogeniculate pathway in the mouse. Vis Neurosci. 2005;22(5):661–76.

    Article  PubMed  Google Scholar 

  53. Hooks BM, Chen C. Distinct roles for spontaneous and visual activity in remodeling of the retinogeniculate synapse. Neuron. 2006;52(2):281–91.

    Article  PubMed  CAS  Google Scholar 

  54. Lin DJ, Kang E, Chen C. Changes in input strength and number are driven by distinct mechanisms at the retinogeniculate synapse. J Neurophysiol. 2014;112(4):942–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Davis GW, Muller M. Homeostatic control of presynaptic neurotransmitter release. Annu Rev Physiol. 2015;77:251–70.

    Article  PubMed  CAS  Google Scholar 

  56. Burrone J, Murthy VN. Synaptic gain control and homeostasis. Curr Opin Neurobiol. 2003;13(5):560–7.

    Article  PubMed  CAS  Google Scholar 

  57. Oleskevich S, Walmsley B. Synaptic transmission in the auditory brainstem of normal and congenitally deaf mice. J Physiol. 2002;540(Pt 2):447–55.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Echegoyen J, Neu A, Graber KD, Soltesz I. Homeostatic plasticity studied using in vivo hippocampal activity-blockade: synaptic scaling, intrinsic plasticity and age-dependence. PLoS One. 2007;2(8):e700.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Lambo ME, Turrigiano GG. Synaptic and intrinsic homeostatic mechanisms cooperate to increase L2/3 pyramidal neuron excitability during a late phase of critical period plasticity. J Neurosci. 2013;33(20):8810–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Slomowitz E, Styr B, Vertkin I, Milshtein-Parush H, Nelken I, Slutsky M, Slutsky I. Interplay between population firing stability and single neuron dynamics in hippocampal networks. elife. 2015;4:e04378.

  61. Pozo K, Goda Y. Unraveling mechanisms of homeostatic synaptic plasticity. Neuron. 2010;66(3):337–51.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Chattopadhyaya B, Di Cristo G, Higashiyama H, Knott GW, Kuhlman SJ, Welker E, Huang ZJ. Experience and activity-dependent maturation of perisomatic GABAergic innervation in primary visual cortex during a postnatal critical period. J Neurosci. 2004;24(43):9598–611.

    Article  PubMed  CAS  Google Scholar 

  63. Marty S, Wehrle R, Sotelo C. Neuronal activity and brain-derived neurotrophic factor regulate the density of inhibitory synapses in organotypic slice cultures of postnatal hippocampus. J Neurosci. 2000;20(21):8087–95.

    Article  PubMed  CAS  Google Scholar 

  64. Seil FJ, Drake-Baumann R. Reduced cortical inhibitory synaptogenesis in organotypic cerebellar cultures developing in the absence of neuronal activity. J Comp Neurol. 1994;342(3):366–77.

    Article  PubMed  CAS  Google Scholar 

  65. Hartman KN, Pal SK, Burrone J, Murthy VN. Activity-dependent regulation of inhibitory synaptic transmission in hippocampal neurons. Nat Neurosci. 2006;9(5):642–9.

    Article  PubMed  CAS  Google Scholar 

  66. Jiao Y, Zhang C, Yanagawa Y, Sun QQ. Major effects of sensory experiences on the neocortical inhibitory circuits. J Neurosci. 2006;26(34):8691–701.

    Article  PubMed  CAS  Google Scholar 

  67. Flores CE, Mendez P. Shaping inhibition: activity dependent structural plasticity of GABAergic synapses. Front Cell Neurosci. 2014;8:327.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Rutherford LC, Nelson SB, Turrigiano GG. BDNF has opposite effects on the quantal amplitude of pyramidal neuron and interneuron excitatory synapses. Neuron. 1998;21(3):521–30.

    Article  PubMed  CAS  Google Scholar 

  69. Maffei A, Nataraj K, Nelson SB, Turrigiano GG. Potentiation of cortical inhibition by visual deprivation. Nature. 2006;443(7107):81–4.

    Article  PubMed  CAS  Google Scholar 

  70. Maffei A, Turrigiano GG. Multiple modes of network homeostasis in visual cortical layer 2/3. J Neurosci. 2008;28(17):4377–84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  71. Borodinsky LN, Belgacem YH, Swapna I, Sequerra EB. Dynamic regulation of neurotransmitter specification: relevance to nervous system homeostasis. Neuropharmacology. 2014;78:75–80.

    Article  PubMed  CAS  Google Scholar 

  72. Spitzer NC. Neurotransmitter switching? No surprise. Neuron. 2015;86(5):1131–44.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Spitzer NC. Neurotransmitter switching in the developing and adult brain. Annu Rev Neurosci. 2017;40:1–19.

    Article  PubMed  CAS  Google Scholar 

  74. Borodinsky LN, Root CM, Cronin JA, Sann SB, Gu X, Spitzer NC. Activity-dependent homeostatic specification of transmitter expression in embryonic neurons. Nature. 2004;429(6991):523–30.

    Article  PubMed  CAS  Google Scholar 

  75. Guemez-Gamboa A, Xu L, Meng D, Spitzer NC. Non-cell-autonomous mechanism of activity-dependent neurotransmitter switching. Neuron. 2014;82(5):1004–16.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  76. Tasic B, Menon V, Nguyen TN, Kim TK, Jarsky T, Yao Z, Levi B, Gray LT, Sorensen SA, Dolbeare T, et al. Adult mouse cortical cell taxonomy revealed by single cell transcriptomics. Nat Neurosci. 2016;19(2):335–46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Cembrowski MS, Wang L, Sugino K, Shields BC, Spruston N. Hipposeq: a comprehensive RNA-seq database of gene expression in hippocampal principal neurons. elife. 2016;5:e14997.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Toledo-Rodriguez M, Markram H. Single-cell RT-PCR, a technique to decipher the electrical, anatomical, and genetic determinants of neuronal diversity. Methods Mol Biol. 2007;403:123–39.

    Article  PubMed  CAS  Google Scholar 

  79. Bloodgood BL, Sharma N, Browne HA, Trepman AZ, Greenberg ME. The activity-dependent transcription factor NPAS4 regulates domain-specific inhibition. Nature. 2013;503(7474):121–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  80. Bartley AF, Huang ZJ, Huber KM, Gibson JR. Differential activity-dependent, homeostatic plasticity of two neocortical inhibitory circuits. J Neurophysiol. 2008;100(4):1983–94.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Pieraut S, Gounko N, Sando R 3rd, Dang W, Rebboah E, Panda S, Madisen L, Zeng H, Maximov A. Experience-dependent remodeling of basket cell networks in the dentate gyrus. Neuron. 2014;84(1):107–22.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Miles R, Toth K, Gulyas AI, Hajos N, Freund TF. Differences between somatic and dendritic inhibition in the hippocampus. Neuron. 1996;16(4):815–23.

    Article  PubMed  CAS  Google Scholar 

  83. Pouille F, Scanziani M. Enforcement of temporal fidelity in pyramidal cells by somatic feed-forward inhibition. Science. 2001;293(5532):1159–63.

    Article  PubMed  CAS  Google Scholar 

  84. Maffei A, Nelson SB, Turrigiano GG. Selective reconfiguration of layer 4 visual cortical circuitry by visual deprivation. Nat Neurosci. 2004;7(12):1353–9.

    Article  PubMed  CAS  Google Scholar 

  85. Takesian AE, Kotak VC, Sanes DH. Presynaptic GABA(B) receptors regulate experience-dependent development of inhibitory short-term plasticity. J Neurosci. 2010;30(7):2716–27.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  86. Takesian AE, Kotak VC, Sharma N, Sanes DH. Hearing loss differentially affects thalamic drive to two cortical interneuron subtypes. J Neurophysiol. 2013;110(4):999–1008.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Tien NW, Soto F, Kerschensteiner D. Homeostatic plasticity shapes cell-type-specific wiring in the retina. Neuron. 2017;94(3):656–65. e654

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  88. Blankenship AG, Feller MB. Mechanisms underlying spontaneous patterned activity in developing neural circuits. Nat Rev Neurosci. 2010;11(1):18–29.

    Article  PubMed  CAS  Google Scholar 

  89. Kerschensteiner D. Spontaneous network activity and synaptic development. J Neurosci. 2014;20(3):272–90.

    CAS  Google Scholar 

  90. Ackman JB, Burbridge TJ, Crair MC. Retinal waves coordinate patterned activity throughout the developing visual system. Nature. 2012;490(7419):219–25.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Assali A, Gaspar P, Rebsam A. Activity dependent mechanisms of visual map formation--from retinal waves to molecular regulators. Semin Cell Dev Biol. 2014;35:136–46.

    Article  PubMed  CAS  Google Scholar 

  92. Arroyo DA, Feller MB. Spatiotemporal features of retinal waves instruct the wiring of the visual circuitry. Front Neural Circuits. 2016;10:54.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Stacy RC, Demas J, Burgess RW, Sanes JR, Wong RO. Disruption and recovery of patterned retinal activity in the absence of acetylcholine. J Neurosci. 2005;25(41):9347–57.

    Article  PubMed  CAS  Google Scholar 

  94. Sun C, Warland DK, Ballesteros JM, van der List D, Chalupa LM: Retinal waves in mice lacking the beta2 subunit of the nicotinic acetylcholine receptor. Proc Natl Acad Sci U S A 2008, 105(36):13638–13643.

  95. Stafford BK, Sher A, Litke AM, Feldheim DA. Spatial-temporal patterns of retinal waves underlying activity-dependent refinement of retinofugal projections. Neuron. 2009;64(2):200–12.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Bansal A, Singer JH, Hwang BJ, Xu W, Beaudet A, Feller MB. Mice lacking specific nicotinic acetylcholine receptor subunits exhibit dramatically altered spontaneous activity patterns and reveal a limited role for retinal waves in forming ON and OFF circuits in the inner retina. J Neurosci. 2000;20(20):7672–81.

    Article  PubMed  CAS  Google Scholar 

  97. Blankenship AG, Ford KJ, Johnson J, Seal RP, Edwards RH, Copenhagen DR, Feller MB. Synaptic and extrasynaptic factors governing glutamatergic retinal waves. Neuron. 2009;62(2):230–41.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  98. Wenner P. Homeostatic synaptic plasticity in developing spinal networks driven by excitatory GABAergic currents. Neuropharmacology. 2014;78:55–62.

    Article  PubMed  CAS  Google Scholar 

  99. Ben-Ari Y, Gaiarsa JL, Tyzio R, Khazipov R. GABA: a pioneer transmitter that excites immature neurons and generates primitive oscillations. Physiol Rev. 2007;87(4):1215–84.

    Article  PubMed  CAS  Google Scholar 

  100. Blaesse P, Airaksinen MS, Rivera C, Kaila K. Cation-chloride cotransporters and neuronal function. Neuron. 2009;61(6):820–38.

    Article  PubMed  CAS  Google Scholar 

  101. Gonzalez-Islas C, Wenner P. Spontaneous network activity in the embryonic spinal cord regulates AMPAergic and GABAergic synaptic strength. Neuron. 2006;49(4):563–75.

    Article  PubMed  CAS  Google Scholar 

  102. Gonzalez-Islas C, Chub N, Garcia-Bereguiain MA, Wenner P. GABAergic synaptic scaling in embryonic motoneurons is mediated by a shift in the chloride reversal potential. J Neurosci. 2010;30(39):13016–20.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  103. Torborg CL, Feller MB. Spontaneous patterned retinal activity and the refinement of retinal projections. Prog Neurobiol. 2005;76(4):213–35.

    Article  PubMed  Google Scholar 

  104. Huberman AD, Wang GY, Liets LC, Collins OA, Chapman B, Chalupa LM. Eye-specific retinogeniculate segregation independent of normal neuronal activity. Science. 2003;300(5621):994–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Torborg CL, Hansen KA, Feller MB. High frequency, synchronized bursting drives eye-specific segregation of retinogeniculate projections. Nat Neurosci. 2005;8(1):72–8.

    Article  PubMed  CAS  Google Scholar 

  106. Giachello CN, Baines RA. Regulation of motoneuron excitability and the setting of homeostatic limits. Curr Opin Neurobiol. 2017;43:1–6.

    Article  PubMed  CAS  Google Scholar 

  107. Hobert O. Terminal selectors of neuronal identity. Curr Top Dev Biol. 2016;116:455–75.

    Article  PubMed  Google Scholar 

  108. Marder E, Goaillard JM. Variability, compensation and homeostasis in neuron and network function. Nat Rev Neurosci. 2006;7(7):563–74.

    Article  PubMed  CAS  Google Scholar 

  109. Giachello CN, Baines RA. Inappropriate neural activity during a sensitive period in embryogenesis results in persistent seizure-like behavior. Current biology : CB. 2015;25(22):2964–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  110. Truszkowski TL, Aizenman CD. Neurobiology: setting the set point for neural homeostasis. Current biology : CB. 2015;25(23):R1132–3.

    Article  PubMed  CAS  Google Scholar 

  111. Ramocki MB, Zoghbi HY. Failure of neuronal homeostasis results in common neuropsychiatric phenotypes. Nature. 2008;455(7215):912–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  112. Meredith RM, Dawitz J, Kramvis I. Sensitive time-windows for susceptibility in neurodevelopmental disorders. Trends Neurosci. 2012;35(6):335–44.

    Article  PubMed  CAS  Google Scholar 

  113. Mullins C, Fishell G, Tsien RW. Unifying views of autism Spectrum disorders: a consideration of autoregulatory feedback loops. Neuron. 2016;89(6):1131–56.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  114. Vislay RL, Martin BS, Olmos-Serrano JL, Kratovac S, Nelson DL, Corbin JG, Huntsman MM. Homeostatic responses fail to correct defective amygdala inhibitory circuit maturation in fragile X syndrome. J Neurosci. 2013;33(17):7548–58.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Funding

Work of the authors was supported by funding from the National Institutes of Health (NIH EY023441, EY026978, and EY027411).

Availability of data and materials

Data sharing not applicable to this review article.

Author information

Authors and Affiliations

Authors

Contributions

N-WT and DK wrote the manuscript. Both authors read and approved the final manuscript.

Corresponding authors

Correspondence to Nai-Wen Tien or Daniel Kerschensteiner.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tien, NW., Kerschensteiner, D. Homeostatic plasticity in neural development. Neural Dev 13, 9 (2018). https://doi.org/10.1186/s13064-018-0105-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13064-018-0105-x

Keywords