Open Access
26 September 2023 Capturing a rising star: the emerging role of astrocytes in neural circuit wiring and plasticity–lessons from the visual system
David Foubert, Finnley Cookson, Edward S. Ruthazer
Author Affiliations +
Abstract

The increasingly widespread use of calcium imaging to explore the nature of neuronal activity and circuits has unexpectedly revealed the ubiquitous presence and significance of astrocytic activity. Here, we present a brief review of visual system development, placing it in the context of recently identified roles of astrocytes in the modulation of neuronal responses and circuit plasticity, through their responses to sensory stimuli and the release of gliotransmitters.

The orderly representation of sensory space in the form of retinotopic maps in the visual system makes it an accessible and intuitive model system in which to investigate the molecular and experience-dependent mechanisms responsible for establishing precise circuit wiring in the brain. Increasingly, it has become clear that glial cells, long-neglected in early studies, are in fact fundamental partners in this process. Our well-established understanding of the key mechanisms that underlie visual system development therefore offers a useful starting point for investigating the involvement of glia in circuit refinement and function.

The initial pathfinding of retinal ganglion cell (RGC) axons from the eye relies on a diverse set of molecular guidance cues, including slits, netrins, semaphorins, and ephrins to ensure projecting neurons reach their targets and synapse with their proper partners.1,2 Subsequent retinotopic projection refinement exploits patterned neural activity, which in mammals takes the form of spontaneous retinal waves prior to eye-opening.3 The importance of retinal waves in refining RGC axon arbor termination zones in the superior colliculus is underscored by studies in ß2-nicotinic acetylcholine receptor knockout mice, which lack retinal waves. RGC projections in these mice fail to fully refine, occupying a broader area with less dense local arborization.4 Moreover, retinal waves have been found to propagate throughout the visual system where they may provide further wiring guidance in higher visual areas during critical periods for topographic specification.5,6

By contrast, RGCs in most externally fertilized species, including common experimental models such as zebrafish (Danio rerio) and African claw-toed frogs (Xenopus laevis), respond to visual input starting from much earlier developmental timepoints, obviating the need for spontaneous retinal waves to provide patterned activity.7 The developing retinotectal projection in these animals is primed to undergo experience-expectant refinement driven by optic flow-generated retinal activity.8 This property of early visual responsiveness in the retinotectal systems of small, transparent aquatic organisms makes them particularly well-suited for studies of circuit refinement in early development, and much work to date has focused on structural remodeling of single RGC axon arbors or tectal dendritic fields, reviewed in Kutsarova et al.9 The translucency and small size of larval zebrafish and Xenopus tadpoles also account for their having been adopted as subjects in some of the earliest intravital imaging studies of the vertebrate brain, ranging from in vivo time-lapse imaging of single neuron morphology to whole-brain calcium imaging.

1.

Calcium Imaging Reveals Sensory Maps in Both Neurons and Astrocytes

Until recently, it was unclear whether the initial functional organization of RGC inputs, especially to the immature Xenopus tadpole optic tectum, was sufficiently organized to manifest as orderly retinotopic maps. Initial extracellular electrophysiological tectal mapping studies in Xenopus tadpoles indicated remarkably poor retinotopic organization at these early developmental stages.10 While electrophysiological methods have the benefit of reporting cellular responses at high temporal resolution, they are poorly suited for mapping neural connectivity at fine sub-cellular spatial resolution in the small brains of developing organisms (Fig. 1). Interestingly, more recent investigation utilizing in vivo calcium imaging has indeed revealed an impressive degree of fine-scale topographic order at subcellular resolution within the three-dimensional structure of the Xenopus tectal neuropil and a role for neuronal activity in further refining these retinotopic maps.11

Fig. 1

Comparing the application of (a) optical and (b) electrophysiological methods to the X. laevis optic tectum. Electrophysiological methods, although superior in temporal resolution and in detecting spiking, cannot achieve the subcellular spatial resolution, cell-type specificity, and spatial coverage provided by calcium imaging (images generated using BioRender).

NPh_10_4_044408_f001.png

In addition to improved spatial resolution, the possibility of imaging calcium dynamics over large volumes of brain tissue has also heralded a shift from traditional electrophysiological recordings from single or small numbers of neurons, to observing network-wide response properties.12 The use of chemical fluorescent calcium indicators first revealed that astrocytes, far from being passive support cells, are themselves capable of exhibiting robust responses to external stimuli in the form of calcium transients.13 Genetically encoded calcium indicators (GECIs) have largely replaced their chemical predecessors due to the ability to obtain broad expression in defined cell types with minimal invasiveness.14

One surprising set of findings to emerge from in vivo calcium imaging is the degree to which astrocytes can exhibit sensory-evoked calcium transients that are highly selective for stimulus features, and reflect a degree of functional selectivity rivaling that of neurons. In ferret primary visual cortex, astrocytes respond to visual stimuli with defined receptive fields and sharp tuning to stimulus features such as orientation preference.15 These responses, which are detected primarily in the astrocytic somata, are delayed by up to several seconds relative to their neighboring neurons.16 Similarly, in the Xenopus optic tectum, radial astrocytes have been shown to respond to visual loom stimuli with a latency of 2 to 3 s later than nearby neurons.17 Interestingly, even after full pharmacological blockade of all glutamatergic receptors on tectal neurons, these glial responses persist in full, suggesting that they are driven, not by signals released by local tectal neurons, but directly by RGC axon glutamate release. This was confirmed by live calcium imaging in tadpoles expressing the GECI GCaMP6s, in which the visually evoked responses of radial astrocytes were abolished by blockade either of excitatory amino acid transporters (EAATs) expressed in astrocytes or of the sodium–calcium exchanger that pumps calcium into radial astrocytes while exchanging the sodium that EAATs import into the cell together with glutamate for calcium but not by blocking glutamate receptors on postsynaptic tectal neurons. This finding demonstrates a critical role for the sodium–calcium exchanger in mediating direct signaling by glutamate from RGC axon terminals to drive sensory-driven Ca2+ events in glia during early development.

In addition to sensory-driven calcium transients, astrocytes in the brains of mice, as well as fish and frogs, can respond robustly to neuromodulatory influences, such as norepinephrine (NE) released during locomotion.18 Inhibition of alpha1-adrenergic receptors significantly reduces neuromodulator-driven global activity in primary visual cortex, unmasking the smaller visually-driven responses in microdomains.16 As opposed to somatic calcium elevations, astrocyte microdomains are mostly found locally within fine peripheral processes near neuronal synapses and can respond to single synaptic events with rapid local calcium elevations on a timescale more similar to neuronal events.19,20 Using a clever anterograde viral transduction method to target expression of a GECI to astrocytes that interact with thalamocortical axons in the whisker cortex in mouse, Georgiou et al.21 showed that sensory-evoked calcium microdomains in astrocytes could be observed as hotspots where calcium elevation was strongly elevated during volitional locomotor behavior, presumably via NE release, and that the locations and properties of these behaviorally modulated hotspots were stable over days. Taken together, these findings suggest that astrocytes do not only read out local glutamate release but also modulate their responses to these stimuli depending on behavioral state and neuromodulation. How might such differences in glial activation, based on behavioral state and neuromodulatory input, impact the mechanisms that underlie developmental circuit refinement in the retinotectal system?

2.

Gliotransmission Mediates Metaplasticity

In the developing retinotectal (retinocollicular) system, N-methyl-D-aspartate receptor (NMDAR) blockade prevents normal circuit refinement, measured both functionally and anatomically.11,2226 A useful set of experimental models for studying NMDAR-dependent mechanisms in Xenopus development are retinotectal spike-timing-dependent long-term potentiation and depression.27 Spike-timing-dependent plasticity has been shown to induce receptive field shifts, suggesting its relevance for circuit refinement.28 Here too, optical approaches have proven useful for studying network-scale plasticity. Using calcium imaging to quantify in parallel the response changes of dozens of tectal neurons while imaging single-cell morphological remodeling, it was shown that plasticity-inducing stimuli that strengthen functional responses also promote dendritic growth and stabilization; conversely, stimuli that depress tectal responsiveness lead to dendritic branch loss.29,30 Similarly, on the presynaptic side, RGC axons that fire synchronously with neighboring inputs engage an NMDAR-dependent Hebbian branch stabilization mechanism, whereas the axons that fire out of synchrony with other inputs exhibit destabilization and exploratory branch induction, referred to as Stentian plasticity.9,31

Astrocytes in many brain areas are able to modulate synaptic efficacy through their calcium transient-driven release of gliotransmitters, including D-serine, adenosine triphosphate (ATP), and glutamate. D-serine is a co-agonist of the NMDAR that greatly enhances calcium currents through NMDARs and thus has been observed to act as a metaplasticity factor to favor LTP over LTD.32 Electrophysiological experiments have shown that calcium-mediated D-serine release from astrocytes promotes NMDAR-dependent LTP in rodent cornu Ammonis 1.3335 In the retinotectal system, D-serine has been shown to promote synaptic maturation through the addition of aminomethylphosphonic acid type glutamate receptor, and Hebbian axonal stabilization through its enhancement of NMDAR function.36

Gliotransmitters ATP and glutamate have also been linked to changes in synaptic efficacy and thus may contribute to the plasticity of the retinotectal system. ATP can either act directly on purinergic receptors or can be degraded to adenosine to act on a distinct family of G-protein-coupled receptors. Although not yet shown in the optic tectum, activation of astrocytes by NE was shown to trigger ATP release and activation of P2X(7) receptors on neurons to mediate an increase in postsynaptic efficacy in rat hypothalamus.37 Furthermore, glutamate itself is believed to act as a gliotransmitter in some circumstances. For example, it has also been demonstrated that glutamate-induced glutamate release from Bergmann glia in the cerebellum amplifies excitatory signals by contributing to synaptically mediated currents.38

3.

Modulation of Brain States through Astrocyte Signaling

Given the large, tiled domains serviced by the fine processes of astrocytes, bringing each in contact with about 100,000 synapses, as well as their potential to communicate with one another via gap junctions, they are well-positioned to regulate circuit-wide neural activity and brain states. It therefore makes sense that neuromodulatory inputs might leverage the intimate access that astrocytes have with synapses across a brain region to both expand and better target their influence on the network.

In mice, stimulation of nucleus basalis, whose activity normally correlates with elevated attentional states, drives modulatory acetylcholine release that activates muscarinic receptors on astrocytes in primary visual cortex.39 This, in turn, causes slow currents in cortical neurons that enhance NMDAR-mediated signaling, an outcome that is prevented when calcium transients in the astrocytes are blocked. Impressively, the pairing of cholinergic stimulation with presentation of visual stimuli at a given orientation specifically enhanced the response of cortical neurons to the paired orientation.

Whole-brain calcium imaging in actively swimming zebrafish revealed that radial astrocytes accumulate evidence of swimming futility from noradrenergic inputs, which fire when motor activity is experimentally mismatched with visual feedback.40 The radial astrocytes exhibit bursts of intense pan-network calcium transients that appear to prompt gamma-aminobutyric acid neurons to arrest motor output in the hindbrain.

Astrocyte activity also appears to precede transitions between slow wave sleep and wakefulness in the mammalian brain. Disruption of calcium signaling in astrocytes impairs slow wave sleep and results in sleep pattern anomalies.41 Astrocytic calcium signaling in cortex is strongly correlated with changes in arousal state. Arousal-associated NE release precedes astrocyte calcium activity, and the blockade of alpha1a-adrenergic (Adra1a) receptors reduces the correlation between astrocytes and cortical state.42 Conditional knockout of Adra1a in astrocytes increased arousal-related neuronal activity and asynchronous neuronal activity. In contrast, activation of glial Adra1a receptors increased cortical synchrony, suggesting that NE-driven astrocyte signaling, in contrast to its direct effects on neurons, mediates the transition back from high to low arousal states.

Recent work from our lab indicates that NE is similarly effective in driving radial astrocyte activation in the optic tectum of developing Xenopus tadpoles and that much like mammalian cortex this glial activity profoundly alters neuronal excitability and selectivity for visual stimuli.43 Moving forward, small transparent models such as the Xenopus visual system can offer a powerful platform in which to combine opto- or chemogenetic control of glial activity with classic neurophotonic assays for structural and functional plasticity of the developing retinotectal system. The traditional view of the experience-dependent fine-tuning of developing brain circuits as an exclusively neuronal phenomenon is due to undergo significant modulation in coming years.

Disclosure

The authors declare no conflict of interest.

References

1. 

S. A. Russell and G. J. Bashaw, “Axon guidance pathways and the control of gene expression,” Dev. Dyn., 247 (4), 571 –580 https://doi.org/10.1002/dvdy.24609 DEDYEI 1097-0177 (2018). Google Scholar

2. 

F. Mann, W. A. Harris and C. E. Holt, “New views on retinal axon development: a navigation guide,” Int. J. Dev. Biol., 48 (8–9), 957 –964 https://doi.org/10.1387/ijdb.041899fm IJDBE5 0214-6282 (2004). Google Scholar

3. 

T. McLaughlin et al., “Retinotopic map refinement requires spontaneous retinal waves during a brief critical period of development,” Neuron, 40 (6), 1147 –1160 https://doi.org/10.1016/S0896-6273(03)00790-6 NERNET 0896-6273 (2003). Google Scholar

4. 

O. S. Dhande et al., “Development of single retinofugal axon arbors in normal and β2 knock-out mice,” J. Neurosci., 31 (9), 3384 –3399 https://doi.org/10.1523/JNEUROSCI.4899-10.2011 JNRSDS 0270-6474 (2011). Google Scholar

5. 

J. B. Ackman, T. J. Burbridge and M. C. Crair, “Retinal waves coordinate patterned activity throughout the developing visual system,” Nature, 490 (7419), 219 –225 https://doi.org/10.1038/nature11529 (2012). Google Scholar

6. 

T. Murakami et al., “Modular strategy for development of the hierarchical visual network in mice,” Nature, 608 (7923), 578 –585 https://doi.org/10.1038/s41586-022-05045-w (2022). Google Scholar

7. 

K. G. Pratt, M. Hiramoto and H. T. Cline, “An evolutionarily conserved mechanism for activity-dependent visual circuit development,” Front. Neural Circuits, 10 79 https://doi.org/10.3389/fncir.2016.00079 (2016). Google Scholar

8. 

M. Hiramoto and H. T. Cline, “Optic flow instructs retinotopic map formation through a spatial to temporal to spatial transformation of visual information,” Proc. Natl. Acad. Sci. U. S. A., 111 (47), E5105 –E5113 https://doi.org/10.1073/pnas.1416953111 (2014). Google Scholar

9. 

E. Kutsarova, M. Munz and E. S. Ruthazer, “Rules for shaping neural connections in the developing brain,” Front. Neural Circuits, 10 111 https://doi.org/10.3389/fncir.2016.00111 (2016). Google Scholar

10. 

R. M. Gaze, M. J. Keating and S. H. Chung, “The evolution of the retinotectal map during development in Xenopus,” Proc. R. Soc. Lond. B Biol. Sci., 185 (1080), 301 –330 https://doi.org/10.1098/rspb.1974.0021 (1974). Google Scholar

11. 

V. J. Li, A. Schohl and E. S. Ruthazer, “Topographic map formation and the effects of NMDA receptor blockade in the developing visual system,” Proc. Natl. Acad. Sci. U. S. A., 119 (8), e2107899119 https://doi.org/10.1073/pnas.2107899119 (2022). Google Scholar

12. 

M. B. Ahrens et al., “Whole-brain functional imaging at cellular resolution using light-sheet microscopy,” Nat. Methods, 10 (5), 413 –420 https://doi.org/10.1038/nmeth.2434 1548-7091 (2013). Google Scholar

13. 

A. H. Cornell-Bell et al., “Glutamate induces calcium waves in cultured astrocytes: long-range glial signaling,” Science, 247 (4941), 470 –473 https://doi.org/10.1126/science.1967852 SCIEAS 0036-8075 (1990). Google Scholar

14. 

T.-W. Chen et al., “Ultra-sensitive fluorescent proteins for imaging neuronal activity,” Nature, 499 (7458), 295 –300 https://doi.org/10.1038/nature12354 (2013). Google Scholar

15. 

J. Schummers, H. Yu and M. Sur, “Tuned responses of astrocytes and their influence on hemodynamic signals in the visual cortex,” Science, 320 (5883), 1638 –1643 https://doi.org/10.1126/science.1156120 SCIEAS 0036-8075 (2008). Google Scholar

16. 

K. Sonoda et al., “Astrocytes in the mouse visual cortex reliably respond to visual stimulation,” Biochem. Biophys. Res. Commun., 505 (4), 1216 –1222 https://doi.org/10.1016/j.bbrc.2018.10.027 BBRCA9 0006-291X (2018). Google Scholar

17. 

N. J. Benfey et al., “Sodium-calcium exchanger mediates sensory-evoked glial calcium transients in the developing retinotectal system,” Cell Rep., 37 (1), 109791 https://doi.org/10.1016/j.celrep.2021.109791 (2021). Google Scholar

18. 

M. Slezak et al., “Distinct mechanisms for visual and motor-related astrocyte responses in mouse visual cortex,” Curr. Biol., 29 (18), 3120 –3127.e5 https://doi.org/10.1016/j.cub.2019.07.078 CUBLE2 0960-9822 (2019). Google Scholar

19. 

E. Bindocci et al., “Three-dimensional Ca2+ imaging advances understanding of astrocyte biology,” Science, 356 (6339), eaai8185 https://doi.org/10.1126/science.aai8185 SCIEAS 0036-8075 (2017). Google Scholar

20. 

J. L. Stobart et al., “Cortical circuit activity evokes rapid astrocyte calcium signals on a similar timescale to neurons,” Neuron, 98 (4), 726 –735.e4 https://doi.org/10.1016/j.neuron.2018.03.050 NERNET 0896-6273 (2018). Google Scholar

21. 

L. Georgiou et al., “Ca2+ activity maps of astrocytes tagged by axoastrocytic AAV transfer,” Sci. Adv., 8 (6), eabe5371 https://doi.org/10.1126/sciadv.abe5371 STAMCV 1468-6996 (2022). Google Scholar

22. 

L. Huang and S. L. Pallas, “NMDA antagonists in the superior colliculus prevent developmental plasticity but not visual transmission or map compression,” J. Neurophysiol., 86 (3), 1179 –1194 https://doi.org/10.1152/jn.2001.86.3.1179 JONEA4 0022-3077 (2001). Google Scholar

23. 

K. O. Johnson, L. Harel and J. W. Triplett, “Postsynaptic NMDA receptor expression is required for visual corticocollicular projection refinement in the mouse superior colliculus,” J. Neurosci., 43 (8), 1310 –1320 https://doi.org/10.1523/JNEUROSCI.1473-22.2022 JNRSDS 0270-6474 (2023). Google Scholar

24. 

K. G. Pratt, W. Dong and C. D. Aizenman, “Development and spike timing–dependent plasticity of recurrent excitation in the Xenopus optic tectum,” Nat. Neurosci., 11 (4), 467 –475 https://doi.org/10.1038/nn2076 NANEFN 1097-6256 (2008). Google Scholar

25. 

E. S. Ruthazer, C. J. Akerman and H. T. Cline, “Control of axon branch dynamics by correlated activity in vivo,” Science, 301 (5629), 66 –70 https://doi.org/10.1126/science.1082545 SCIEAS 0036-8075 (2003). Google Scholar

26. 

J. T. Schmidt, M. R. Fleming and B. Leu, “Presynaptic protein kinase C controls maturation and branch dynamics of developing retinotectal arbors: possible role in activity-driven sharpening,” J. Neurobiol., 58 (3), 328 –340 https://doi.org/10.1002/neu.10286 JNEUBZ 0022-3034 (2004). Google Scholar

27. 

L. I. Zhang et al., “A critical window for cooperation and competition among developing retinotectal synapses,” Nature, 395 (6697), 37 –44 https://doi.org/10.1038/25665 (1998). Google Scholar

28. 

R. L. Vislay-Meltzer, A. R. Kampff and F. Engert, “Spatiotemporal specificity of neuronal activity directs the modification of receptive fields in the developing retinotectal system,” Neuron, 50 (1), 101 –114 https://doi.org/10.1016/j.neuron.2006.02.016 NERNET 0896-6273 (2006). Google Scholar

29. 

S. X. Chen et al., “The transcription factor MEF2 directs developmental visually driven functional and structural metaplasticity,” Cell, 151 (1), 41 –55 https://doi.org/10.1016/j.cell.2012.08.028 CELLB5 0092-8674 (2012). Google Scholar

30. 

D. Dunfield and K. Haas, “Metaplasticity governs natural experience-driven plasticity of nascent embryonic brain circuits,” Neuron, 64 (2), 240 –250 https://doi.org/10.1016/j.neuron.2009.08.034 NERNET 0896-6273 (2009). Google Scholar

31. 

M. Munz et al., “Rapid Hebbian axonal remodeling mediated by visual stimulation,” Science, 344 (6186), 904 –909 https://doi.org/10.1126/science.1251593 SCIEAS 0036-8075 (2014). Google Scholar

32. 

M. Van Horn, M. Sild and E. Ruthazer, “D-serine as a gliotransmitter and its roles in brain development and disease,” Front. Cell. Neurosci., 7 39 https://doi.org/10.3389/fncel.2013.00039 (2013). Google Scholar

33. 

C. Henneberger et al., “Long-term potentiation depends on release of D-serine from astrocytes,” Nature, 463 (7278), 232 –236 https://doi.org/10.1038/nature08673 (2010). Google Scholar

34. 

A. Panatier et al., “Glia-derived d-serine controls NMDA receptor activity and synaptic memory,” Cell, 125 (4), 775 –784 https://doi.org/10.1016/j.cell.2006.02.051 CELLB5 0092-8674 (2006). Google Scholar

35. 

M. W. Sherwood et al., “Astrocytic IP3Rs: contribution to Ca2+ signalling and hippocampal LTP,” Glia, 65 (3), 502 –513 https://doi.org/10.1002/glia.23107 GLIAEJ 1098-1136 (2017). Google Scholar

36. 

M. R. Van Horn et al., “The gliotransmitter d-serine promotes synapse maturation and axonal stabilization in vivo,” J. Neurosci., 37 (26), 6277 –6288 https://doi.org/10.1523/JNEUROSCI.3158-16.2017 JNRSDS 0270-6474 (2017). Google Scholar

37. 

G. R. J. Gordon et al., “Norepinephrine triggers release of glial ATP to increase postsynaptic efficacy,” Nat. Neurosci., 8 (8), 1078 –1086 https://doi.org/10.1038/nn1498 NANEFN 1097-6256 (2005). Google Scholar

38. 

K. Beppu, N. Kubo and K. Matsui, “Glial amplification of synaptic signals,” J. Physiol., 599 (7), 2085 –2102 https://doi.org/10.1113/JP280857 JPHYA7 0022-3751 (2021). Google Scholar

39. 

N. Chen et al., “Nucleus basalis-enabled stimulus-specific plasticity in the visual cortex is mediated by astrocytes,” Proc. Natl. Acad. Sci. U. S. A., 109 (41), E2832 –E2841 https://doi.org/10.1073/pnas.1206557109 (2012). Google Scholar

40. 

Y. Mu et al., “Glia accumulate evidence that actions are futile and suppress unsuccessful behavior,” Cell, 178 (1), 27 –43.e19 https://doi.org/10.1016/j.cell.2019.05.050 CELLB5 0092-8674 (2019). Google Scholar

41. 

L. Bojarskaite et al., “Astrocytic Ca2+ signaling is reduced during sleep and is involved in the regulation of slow wave sleep,” Nat. Commun., 11 (1), 3240 https://doi.org/10.1038/s41467-020-17062-2 NCAOBW 2041-1723 (2020). Google Scholar

42. 

M. E. Reitman et al., “Norepinephrine links astrocytic activity to regulation of cortical state,” Nat. Neurosci., 26 (4), 579 –593 https://doi.org/10.1038/s41593-023-01284-w NANEFN 1097-6256 (2023). Google Scholar

43. 

E. S. Ruthazer, “Functional plasticity in developing retinotopic maps,” in 6th Annu. Front. in Neurophotonics Symp., (2022). Google Scholar

Biographies of the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
David Foubert, Finnley Cookson, and Edward S. Ruthazer "Capturing a rising star: the emerging role of astrocytes in neural circuit wiring and plasticity–lessons from the visual system," Neurophotonics 10(4), 044408 (26 September 2023). https://doi.org/10.1117/1.NPh.10.4.044408
Received: 15 May 2023; Accepted: 8 September 2023; Published: 26 September 2023
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
KEYWORDS
Calcium

Neurons

Visual system

Brain

Visualization

Axons

Modulation

RELATED CONTENT


Back to Top