Next Article in Journal
Recent Advances in Pancreatic Cancer: Novel Prognostic Biomarkers and Targeted Therapy—A Review of the Literature
Previous Article in Journal
DJ-1 Acts as a Scavenger of α-Synuclein Oligomers and Restores Monomeric Glycated α-Synuclein
Previous Article in Special Issue
Inhibition of Astrocytic Histamine N-Methyltransferase as a Possible Target for the Treatment of Alzheimer’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Calcium Influx Pathways in Astrocyte Calcium Microdomain Physiology

College of Pharmacy, Rady Faculty of Health Sciences, University of Manitoba, 750 McDermot Avenue, Winnipeg, MG R3E 0T5, Canada
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(10), 1467; https://doi.org/10.3390/biom11101467
Submission received: 28 August 2021 / Revised: 25 September 2021 / Accepted: 1 October 2021 / Published: 6 October 2021
(This article belongs to the Special Issue Astroglia in Physiology, Pathology and Therapy)

Abstract

:
Astrocytes are complex glial cells that play many essential roles in the brain, including the fine-tuning of synaptic activity and blood flow. These roles are linked to fluctuations in intracellular Ca2+ within astrocytes. Recent advances in imaging techniques have identified localized Ca2+ transients within the fine processes of the astrocytic structure, which we term microdomain Ca2+ events. These Ca2+ transients are very diverse and occur under different conditions, including in the presence or absence of surrounding circuit activity. This complexity suggests that different signalling mechanisms mediate microdomain events which may then encode specific astrocyte functions from the modulation of synapses up to brain circuits and behaviour. Several recent studies have shown that a subset of astrocyte microdomain Ca2+ events occur rapidly following local neuronal circuit activity. In this review, we consider the physiological relevance of microdomain astrocyte Ca2+ signalling within brain circuits and outline possible pathways of extracellular Ca2+ influx through ionotropic receptors and other Ca2+ ion channels, which may contribute to astrocyte microdomain events with potentially fast dynamics.

1. Introduction

Astrocytes are brain glial cells that contact nearby neurons and enwrap blood vessels with their highly branched processes. Physiologically, astrocytes are critical for brain homeostasis [1]. They buffer extracellular ions [2], they remove and recycle neurotransmitters [3,4,5], and they supply neurons with energy substrates [6,7,8,9]. However, astrocytes also express a plethora of neurotransmitter receptors, ion channels, and metabolite transporters that respond to nearby neuronal activity and integrate astrocytes into neural networks [1]. Many of these receptors and ion channels induce transient increases in intracellular Ca2+ [10] that are required for various astrocyte functions, as discussed below [10,11,12,13,14,15]. Recently, localized Ca2+ transients in fine astrocytic structures, such as processes and endfeet around blood vessels, have been identified using genetically encoded Ca2+ indicators (GECIs), such as GCaMP6f [16,17,18,19,20,21,22,23,24,25]. Here, we refer to these small, localized Ca2+ transients as astrocyte microdomain Ca2+ events (MCEs).
Astrocyte MCEs are heterogenous; they vary in amplitude and duration, and occur within astrocytes at rest (i.e., in the absence of nearby synaptic activity) [17,18]. The dynamics of astrocyte Ca2+ transients are dictated by the resting, basal intracellular Ca2+ concentration [26], which is higher in fine processes compared to the soma [27]. The number of astrocyte MCEs, their volume, and their amplitude increases [17,18,19,28,29] following nearby neuronal responses evoked by physiological stimuli, such as whisker stimulation-induced somatosensory activation [17,18,30,31], visual stimulation of the visual cortex [29], or odor presentation in the olfactory bulb [28]. The majority of astrocyte somatic Ca2+ events [32,33,34] and MCEs [17,18] activated during local circuit activity have a delayed signal onset latency (for example: MCEs arise 5 s after the start of whisker stimulation). Compared to neuronal Ca2+ signal onset timescales (a few milliseconds after the start of stimulation), this astrocytic Ca2+ signalling was deemed too slow to modulate rapid processes such as synaptic activity or blood flow [32,33,34]. However, fast onset Ca2+ dynamics have recently been described within fine astrocyte structures in response to physiological stimuli in vivo [17,28,30,31,35]. In particular, a subset of astrocyte MCEs near the plasma membrane of astrocyte processes, have a fast signal onset that closely follows neuronal activity (within 100 ms) and are reproducibly evoked within the same regions during repeated whisker stimulation [17]. This suggests that astrocytes have the necessary temporal and spatial Ca2+ signalling to play a rapid role in fine-tuning circuits as discussed below.

2. Functional Roles of Astrocyte Microdomain Ca2+ Events

Astrocytes are active contributors to brain processes through the release of gliotransmitters or vasoactive molecules that modulate the nearby neuronal activity or blood flow [10,11,12]. The gliotransmitters released by astrocytes include glutamate [36], GABA [37,38], ATP [39,40], and possibly D-serine [41,42] (though this remains controversial, as there is evidence of D-serine release from neurons [43,44]). These molecules act on neuronal receptors or nearby astrocyte receptors as a form of glial communication [11]. The release of these molecules is Ca2+ dependent, suggesting that astrocyte Ca2+ events are a key component of bidirectional astrocyte-neuron interactions [11,19]. Specifically, MCEs may play a critical role in confined, localized delivery of gliotransmitters that influence local synaptic activity [39,40,45,46,47,48,49,50], and the recruitment of larger Ca2+ domains or more global astrocyte Ca2+ signals may modulate neuronal networks and dictate animal behaviour [51,52,53,54,55] as outlined more specifically below.
At the synaptic level, astrocyte Ca2+ signalling and gliotransmitter release influences basal synaptic activity, excitatory and inhibitory neurotransmission, and synaptic plasticity (Figure 1) [36,39,40,41,45,50,56,57,58,59]. Some specific examples include, first, astrocytes modulate basal synaptic transmission in the hippocampus [39,45,60] through adenosine that is likely produced from the metabolism of astrocyte ATP released during gliotransmission. Adenosine activates presynaptic A2A [39] or A1 receptors [60] to encourage or reduce neurotransmitter release, respectively. Second, hippocampal pyramidal neuron inhibition is enhanced by astrocyte ATP/adenosine gliotransmission at inhibitory interneuron synapses [40]. Third, glutamate released from astrocytes at excitatory synapses can increase synaptic release [59], boost synaptic strength [57], and elevate neuronal synchrony [36]. Finally, astrocyte glutamate [50,56,61] and D-serine [41,62] also contribute to long-term potentiation (LTP) and long-term depression (LTD) that are important for synaptic plasticity. This may include cholinergic-induced synaptic plasticity following activation of the nucleus basalis [50,63,64].
These examples highlight the diversity of astrocyte-neuron interactions at different synapses and via distinct gliotransmitters; however, a link between localized MCEs and gliotransmission has not been proven. The majority of these studies described above demonstrated a requirement of astrocyte Ca2+ signalling for the modulation of synaptic processes by utilizing Ca2+ chelator BAPTA [39,40,45,56,57] or clamping intracellular Ca2+ levels [41]. These approaches effectively silence all astrocytic intracellular Ca2+ events from microdomains to somatic transients to global Ca2+ waves, irrespective of their cellular location. Future studies that decode the effect of MCEs in astrocytic processes by targeting specific pathways will help to better disentangle the roles of astrocytes in gliotransmission and neuronal modulation.
In addition to gliotransmission, Ca2+ events can induce morphological remodeling of fine astrocytic processes at synapses [65,66,67]. This has the potential to change their synaptic coverage, affecting gliotransmission and synaptic function [68], and suggests that localized astrocytic MCEs within perisynaptic processes may regulate the stability of individual synapses.
During periods of neuronal activity, astrocyte Ca2+ signalling increases both in the number of MCEs within each cell as well as an enlargement of the MCE spatial area [17,18,19,28,29,30,31]. It has been suggested that a scaling of astrocyte Ca2+ signalling may induce heterosynaptic modulation where astrocytes integrate information from multiple synapses to influence additional neighbouring connections, or modulate an entire territory or neuronal network depending on the level of the evoked Ca2+ response [11]. For example, astrocytes play a regulatory role in neocortical slow oscillations that underlie resting brain waves [69,70], since Ca2+ signalling in astrocytes precedes a shift to slow-wave oscillations [70] and induces cortical UP states, where multiple neurons are synchronized [69]. Additionally, multiple studies have shown that robust, global Ca2+ events in astrocytes occur when norepinephrine is released from the locus coeruleus [15,24,33,71,72], suggesting that astrocytes have an important role in network modulation during arousal. Astrocytes have also been linked to animal behaviour, since increased Ca2+ in the hippocampus enhances memory formation [52], while mouse models with reduced astrocyte Ca2+ events (by targeting specific pathways in different brain regions) have repetitive [53], depressive [54], or autistic-like behaviours [55]. Thus, astrocytes may “sense” nearby neuronal activity through Ca2+ events that locally regulate circuit activity, modulate the processing of information in large networks and impact animal behaviour. Fast onset MCEs evoked by neuronal activity could be of critical importance for rapidly tuning changes at single synapses that amount to alterations in activity over larger circuits. Again, future studies specifically targeting pathways that contribute directly to astrocyte MCEs will help to link MCEs to the modulation of single synapses, but will also help determine how the scaling of astrocyte Ca2+ signalling and the recruitment of MCEs influence larger neuronal networks and behaviour.
Astrocytes may also regulate local blood flow through the Ca2+-dependent release of vasoactive molecules, such as arachidonic acid metabolites (Figure 1) [12]. This is important for tonic blood vessel tone [13], particularly during vasomotion [73]. However, a fast, dynamic role for astrocytes in regulating vasodilation during neurovascular coupling remains controversial. Early studies in brain slices ex vivo linked astrocyte Ca2+ to changes in vascular tone [12,74,75,76,77], but this has not translated to in vivo experiments where astrocyte Ca2+ events, particularly in endfeet microdomains, may [28,30,31] or may not [32,72,78] rapidly precede vasodilatory responses during neurovascular coupling. Several of these recent in vivo studies suggest astrocyte Ca2+ events are not essential for vasodilation [32,72,79]; however, when astrocyte endfoot Ca2+ signals are evoked by brief, local circuit activity, the magnitude of the hemodynamic response is enhanced [79]. During prolonged sensory stimulation [79] or the postictal epileptic period [80], slow, sustained astrocyte Ca2+ signals are induced, which correlate with vasoconstriction [81]. Therefore, while astrocytes and MCEs may not rapidly evoke blood flow changes during neurovascular coupling, they provide important, complex homeostatic and modulatory effects on blood flow that are relevant for both vasodilation and vasoconstriction at rest and during periods of brain activity [82].

3. Pathways Underlying Fast Astrocyte MCEs

A number of mechanisms are known to contribute to localized astrocyte MCEs [10,15,20,25]. Spontaneous astrocyte MCEs that occur in the absence of synaptic activity have been shown to be mediated by mitochondrial Ca2+ release [14] via the opening of mitochondrial permeability transition pore [15] and by extracellular Ca2+ influx through transient receptor potential cation channel A1 (TRPA1) [20,25]. It should be noted that other TRP channels such as TRPV1, TRPV4, TRPC1, TRPC3, TRPC4, and TRPC5 may also mediate Ca2+ influx in astrocytes [83,84,85,86,87,88], but there is limited evidence that these channels are directly activated during synaptic transmission.
The most extensively studied astrocyte pathway that contributes to Ca2+ events is the release of Ca2+ from the endoplasmic reticulum following inositol-1,4,5-trisphosphate receptor (IP3R) and upstream Gq-G-protein coupled receptor (GPCR) activation (Figure 2) [1]. This mechanism has been targeted in astrocytes using an IP3R2 knockout mouse [17,24,32,55,89,90], since IP3R2 is believed to be the principal isoform in astrocytes [91]. Knockout of endoplasmic reticulum IP3R2 reduces the number of astrocyte MCEs [17,18,24], but does not prevent increased astrocyte MCE responses in fine processes to arousal [24] or sensory stimulation [18], nor does it reduce the number of fast onset MCEs evoked by nearby synaptic activity [17]. Metabotropic glutamate receptors (mGluRs) were one of the first Gq-GPCR pathways found to elevate Ca2+ in astrocytes [77,92,93]. However, these receptors are potentially more important during development because mature, adult astrocytes have low mGluR mRNA expression [94] and reduced calcium responses to mGluR agonists [95], though this does not exclude mGluR expression and signalling in the fine processes of adult astrocytes [10,96]. Several other GPCR pathways that evoke IP3 signalling in astrocytes are activated by neuromodulators, such as norepinephrine and acetylcholine. These cause astrocyte Ca2+ transients during behavioural arousal states [17,24,71,72], but contribute more to large, delayed onset MCEs [17,24]. This suggests that fast onset MCEs are mediated by mechanisms other than GPCR activity, such as extracellular Ca2+ influx. Here, we discuss key pathways for rapid astrocyte Ca2+ influx through ionotropic receptors and ion channels that are activated during neurotransmission and may play important physiological roles in brain circuits (Figure 2).

3.1. Ionotropic Glutamate Receptors (NMDA, AMPA, and Kainate Receptors)

3.1.1. Astrocyte iGluR Expression

Ionotropic glutamate receptors (iGluRs) are ligand-gated ion channels that conduct cations (Na+, Ca2+ and K+) when activated by synaptic glutamate (Figure 2), and this mediates fast excitatory synaptic transmission. Based on their selective agonists, iGluRs are categorized into three classes, including α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors, kainate receptors, and N-methyl-D-aspartate (NMDA) receptors [97]. AMPA receptors are tetramers formed from four possible subunits (GluA1-GluA4), which dictate the functional properties of the receptor, including their calcium permeability [98]. These receptors also generally have rapid deactivation kinetics [99]. Classical NMDA receptors are hetero-tetramers formed from two GluN1 subunits and two GluN2 subunits (of four possible types, A—D) [100]. There are also less-common GluN3A and 3B subunits, which can join GluN1 and GluN2 to form “non-conventional” receptors or possibly trimeric GluN1/GluN3 receptors [101,102]. The GluN2 subunit composition of conventional NMDA receptors confers functional properties such as a sensitivity to blockade by Mg2+, deactivation kinetics, and Ca2+ permeability [100]. Receptors containing GluN2C/2D are insensitive to blockade by Mg2+ and therefore do not require membrane depolarization to be activated. These subunits are also less permeable to Ca2+ than GluN2A/2B receptors, and they have slower deactivation kinetics [100].
Astrocytes express the genes of iGluRs, albeit at lower levels than neurons. All four AMPA receptor subunits (GluA1-GluA4) have been detected in astrocytes [1,103], although with some regional differences in expression [104,105]. For example, GluA1 and GluA4 are the most common subunits in cortical astrocytes and potentially localize to astrocyte processes [104]. Hippocampal astrocytes may also express GluA2 [106], which reduces calcium permeability through heteromeric receptors [105]. At early developmental stages (before postnatal day 5), astrocyte AMPA receptors deactivate slower than more mature stages (over postnatal day 10) [107]. This suggests that AMPA receptors on mature astrocytes may contribute to brief Ca2+ transients before deactivation. At the mRNA and protein level, NMDA receptor subunits GluN1 and GluN2A/B have been identified in astrocytes [1]. However, pharmacological studies suggest that functional NMDAR in astrocytes contain GluN2C/D and are most probably a heteromeric composition of GluN1, GluN2C/D, and GluN3 [108,109,110]. This explains the low sensitivity of astrocyte NMDA receptors to blockage by Mg2+ within the channel pore, and suggests that these receptors are resistant to deactivation, but have reduced Ca2+ permeability (compared to GluN2A/B receptors). There is evidence of astrocyte kainate receptor subunit expression at the mRNA and protein levels [111,112]; however, the functionality of these receptors remains controversial [1,113,114,115,116,117].
While astrocytes express iGluRs, the functionality of these receptors, particularly regarding Ca2+ permeability and their contribution to Ca2+ signalling, has been controversial. Early Ca2+ imaging studies were conducted in primary astrocyte cultures (Table 1), with several possible issues that could influence the interpretation of the results. First, some of these studies failed to detect NMDA-induced Ca2+ transients in astrocytes [113,114,115,118], but they used 100 μM NMDA, which is over the toxicity concentration threshold (50 μM) [119,120]. When 20 μM NMDA was applied, astrocytic Ca2+ responses were evoked [121]. Second, quisqualate (QA) was used as an agonist in some studies to identify functional AMPA and kainate- iGluRs [113,114,115,122]. However, quisqualate is not an iGluR-specific agonist and can activate metabotropic glutamate receptor I (mGluR I), which may have contributed to the mixed findings that QA-evoked Ca2+ responses have an internal Ca2+ store component [114,115,122]. Application of more specific agonists, such as AMPA, confirmed the presence of functional AMPARs on cultured hippocampal, cortical, and cerebellar astrocytes [122,123] as well as astrocytes in isolated optic nerve [124]. Third, astrocytes were cultured from different brain regions including the cortex, cerebellum, and hippocampus in these studies. Recent evidence suggests that there are regional iGluR expression differences in astrocytes [104,105,108,109,110], which may alter the Ca2+ permeability of the receptor and make it harder to compare results between studies [105,125]. Finally, the main limitation of astrocyte culture studies is that cells are isolated from neonatal animals and maintained for weeks in culture before the experiment. Thus, cultured cells may not reflect the mechanisms and receptor-activated effects of in situ astrocytes [126].
More recent studies have examined iGluR-mediated Ca2+ dynamics in ex vivo brain slices or acutely isolated astrocytes (Table 2). These preparations used astrocytes that are similar to in situ cells and often combined Ca2+ imaging with electrophysiological recordings. The majority of these studies reported a contribution of AMPA [28,108,124,127] and/or NMDA receptor [93,108,109,110,117,124,128,129,130,131] signalling to astrocyte Ca2+ transients; however, there are a few points to note when considering this work. First, the use of iGluR pharmacology can make it difficult to disentangle the activity of astrocyte receptors from neuronal receptors, particularly when the drugs are bath applied to brain slices where both cell populations are present (Table 2). Studies that have patch-clamped astrocytes and applied iGluR agonists or antagonists directly through the patch pipette have provided more convincing evidence of functional astrocyte AMPA and NMDA receptors that induce cell depolarization and Ca2+ events [108,109,128,129,131,132]. Second, as mentioned above, astrocyte iGluR ion fluxes may vary between different brain regions. For example, astrocyte NMDAR in the hippocampus and cortex permits different levels of Na+ influx [125]. Also, Na+ influx through cortical astrocyte NMDAR, induces reversal of the Na+/Ca2+ exchanger that can further evoke Ca2+ events, as described below [125,133]. This raises the interesting possibility that astrocyte iGluR signalling is fine-tuned to modulate the needs of the local circuit, but it makes it challenging to compare and interpret results from studies of astrocytes in different brain regions (Table 2). Finally, both juvenile [28,92,93,124,127,131,134] and adult animals [108,109,110,117,129,130] have been used to study astrocyte iGluR receptors ex vivo and this may provide conflicting results regarding astrocyte NMDARs because the subunit composition and their characteristics change during development [107,132]. Specifically, AMPAR deactivates faster in more mature astrocytes [107], while adult mice over 3 months of age display larger NMDAR-mediated currents and calcium transients than younger animals [132].
Although iGluR agonists evoke Ca2+ transients in astrocytes in culture and brain slices, most studies have focussed on somatic Ca2+ events. It is still unclear if these receptors contribute to astrocyte MCEs within fine processes, particularly during local circuit activity. Several studies have distinguished between Ca2+ responses in different cellular compartments (processes versus soma) by combining Ca2+ imaging dyes with GFAP-eGFP transgenic mice to better label astrocytes [110,128,129]. However, GECIs are now the most reliable way to detect astrocyte Ca2+ events in fine structures. Using GCaMP3 and GCaMP6f, Haustein et al. [135] showed that NMDAR blocker, D-AP5, did not change spontaneous astrocyte MCEs in the hippocampus, which indicates that astrocyte NMDAR may only be activated during nearby synaptic activity. Topical superfusion of AMPA or NMDA receptor antagonists on the brain, significantly reduced slow-onset MCEs in astrocyte endfeet evoked by whisker-stimulation, suggesting that iGluR signalling contributes to these Ca2+ events [72]. In similar studies, fast onset MCEs in astrocyte fine processes and endfeet were identified in response to stimulation of the contralateral ramus infraorbitalis of the trigeminal nerve [30,31], which is physiologically similar to sensory stimulation. The fast astrocyte Ca2+ responses happened on the same time scale as neurons and preceded local vasodilation. Blockers for AMPA or NMDA receptors were applied directly to the brain and both drugs reduced fast Ca2+ events in astrocyte processes, but only CNQX reduced fast Ca2+ events in endfeet [30]. This suggests that iGluR signalling may mediate rapid astrocyte MCEs that have the capacity to contribute to blood flow. The main drawback of all these studies of iGluRs and MCEs is that the pharmacological approaches employed likely affected both neuron and astrocyte receptors [28,30], making it unclear whether the drugs have direct effects on astrocyte iGluRs or if the impact on MCE activity was merely caused by decreased neuronal activity. Future work specifically targeting astrocyte iGluRs by genetic approaches will help to tease apart a role for these receptors in astrocyte MCE signalling, including fast onset events.

3.1.2. Functional Roles of Astrocyte iGluRs

While it is clear that AMPA receptor activation can cause an elevation in astrocyte Ca2+ in the soma, limited studies have found a functional role for astrocyte AMPAR. In the cerebellum, Bergmann glia astrocytes express the GluA1 and GluA4 subunits [136]. When Bergmann glial AMPAR activity is inhibited by (a) expression of the GluA2 subunit that renders AMPAR Ca2+ impermeable [137] or (b) the conditional knockout of GluA1 and GluA4 [136], structural changes occur within the molecular layer of the cerebellum. Glial fine processes retract from Purkinje cell dendritic spines, which leads to delayed glutamate uptake at synapses [137] and deficits in fine motor control [136]. Clearly, Bergmann glia AMPAR are essential components of cerebellar circuits. Further work is required to determine the functional relevance of astrocyte AMPA receptors in other circuits (cortex, hippocampus, etc.), but also to determine if the fast deactivation kinetics of AMPA receptors limit their contribution to astrocyte MCEs and other signalling.
Astrocyte NMDA receptors have functional roles in maintaining astrocyte Ca2+ stores [131], antioxidant protection [121], gliotransmission [130], and the regulation of synaptic strength (Figure 3) [49]. First, pharmacological intervention during theta-burst cortical stimulation suggests that NMDA receptor activity decreases free Ca2+ in astrocytes through elevation of store uptake [131]. Thus, NMDA receptors may regulate basal astrocyte Ca2+ concentrations, which has implications for Ca2+ microdomain activity and their dynamics [26,27]. Second, NMDA-induced somatic Ca2+ transients in cultured cortical astrocytes upregulate the Cdk5/Nrf2 pathway, a key regulator of genes for cell antioxidant machinery [121]. This increases the release of glutathione precursors from astrocytes, which are used by nearby neurons to synthesize glutathione, an important antioxidant. Therefore, activation of astrocytic NMDA receptors may contribute to neuronal protection against oxidative stress. NMDA receptor antagonists cause neurotoxicity [138], and conceivably, a loss of astrocyte NMDA receptor activity by receptor blockade, may remove their antioxidant effects, contributing to neuronal damage. Third, cultured cortical astrocytes release ATP in response to NMDA treatment, which may decrease synaptic inhibition of pyramidal neurons in the cortex [130].
As discussed in Section 2, astrocytic ATP modulates both autistic-like and depressive-like behaviors in mice [54,55], and ATP-derived adenosine regulates basal synaptic transmission [39] and upregulates somatostatin interneuron synaptic activity [40]. Taken together, gliotransmitter release evoked by astrocyte NMDA receptor activity has the potential to alter nearby neuronal responses or play a role in behaviour. If NMDA receptors induce fast onset Ca2+ events in astrocytes, then gliotransmitter release may happen in a temporal realm that can rapidly tune synaptic activity. Finally, astrocyte NMDA receptors may also regulate synaptic strength in the hippocampus by maintaining paired-pulse ratio heterogeneity, which is seen between two presynaptic neurons that target the same postsynaptic cell [49]. Paired-pulse ratio heterogeneity was reduced when: (a) astrocytes were patch-loaded with BAPTA, (b) astrocytes were patch-loaded with NMDA receptor antagonist, MK-801, or (c) the GluN1 subunit of NMDAR was knocked out specifically in astrocytes. This strongly underlines the importance of astrocytic Ca2+ transients and astrocytic NMDA receptor signalling in diversifying the presynaptic strengths and potentially elevating the network dynamics of dendrites [49]. This type of regulation may be highly specialized within specific circuits. For example, the diversity of presynaptic strengths in the stratum radiatum of the hippocampus is specifically maintained by astrocyte NMDA receptors containing the GluN2C subunit [139]. While there is some evidence of a functional role for astrocyte NMDA receptors regarding gliotransmission, antioxidant protection, and synaptic modulation, further studies that selectively target NMDA receptors, such as knock-out of the GluN1 subunit in astrocytes, will advance the concepts of Ca2+ signalling mediated by these receptors and their physiological roles.

3.2. P2X Receptors

3.2.1. Astrocyte P2X Receptor Expression

Astrocytes express ionotropic P2X purinergic receptors (Figure 2), likely composed of heterotrimeric P2X1/5 [140] or homotrimeric P2X7 subunits [1,141]. These ligand-gated ion channels bind synaptic ATP and conduct Ca2+, K+, and Na+ into the cell. The subunit composition confers ATP binding affinity and Ca2+ permeability [1,142,143]. P2X7 receptors are only activated by high extracellular ATP levels and have been linked to pathology and astrocyte reactivity [144,145]. Therefore, P2X1/5, with its higher affinity for ATP and good Ca2+ permeability, is more likely to be involved in astrocyte MCEs, particularly with a fast onset during local circuit activity. So far, the contribution of P2X1/5 activity to astrocyte MCEs has not been explored with GECIs, but P2X activation causes astrocyte Ca2+ transients (primarily somatic) in brain slices and acutely isolated astrocytes, as measured with Ca2+ dyes [109,146].

3.2.2. Functional Roles of Astrocyte P2XRs

Coincidently, astrocyte P2X receptor activation enhances purinergic signalling in different brain regions. In the cortex, astrocyte P2X receptors increase ATP release [147], which modulates nearby synapses. Further, ATP release by astrocytes in the brain stem is evoked by decreased pH, and propagated and amplified by neighbouring astrocytes via P2X receptor activation [148]. This induces the respiratory reflex and increases the breathing rate [148]. Additionally, astrocyte P2X1 receptors have been linked to endfoot Ca2+ transients and capillary dilation during neurovascular coupling, suggesting that these ionotropic receptors induce the release of vasoactive molecules that specifically act on capillaries and not arterioles [146]. Astrocyte P2X receptor activity also decreases with age [132,147], which leads to an increase in inhibitory and a decrease in excitatory neurotransmission [147] as well as impaired LTP [149]. These effects can be mitigated in aged mice through environmental enrichment and caloric restriction [147], which has important implications for the plasticity of astrocyte activity, and the modulation of synaptic transmission and neurovascular coupling by astrocytes via purinergic signalling. Further functional roles of astrocyte P2X receptors will be identified by future studies selectively targeting these receptors by genetic approaches (i.e., astrocyte P2X receptor knockouts).

3.3. Nicotinic Receptors

3.3.1. Astrocyte Nicotinic Receptor Expression

Nicotinic receptors are pentameric ionotropic acetylcholine receptors that conduct Ca2+, Na+ and K+ and are made up of 16 possible subunits. Astrocytes express homomeric alpha-7 nicotinic acetylcholine receptors (α7nAChRs; Figure 2), and activation of these astrocyte receptors in culture or in hippocampal slices induces intracellular Ca2+ transients [150,151]. Based on their subunit composition, α7nAChRs have high Ca2+ permeability, but are rapidly deactivated [152], suggesting they may cause more brief Ca2+ events in astrocytes. α7nAChRs Ca2+ transients are further amplified in astrocytes by Ca2+ release from intracellular Ca2+ stores through ryanodine receptors [150]. At this point, α7nAChR activation has not yet been linked to localized astrocyte MCEs.

3.3.2. Functional Roles of Astrocyte Nicotinic Receptors

Functionally, astrocyte α7nAChRs activation in the hippocampus by acetylcholine from medial septal projections induces D-serine release, leading to nearby neuronal NMDA receptor modulation [153]. This is notably activated by wakeful acetylcholine levels and oscillates throughout the day, creating a rhythmic pattern of gliotransmission [153]. Nicotinic receptor activation also induces morphological changes in the processes of cultured astrocytes [154], which has implications for perisynaptic astrocyte process coverage and remodeling in intact circuits. Finally, α7nAChRs activation in cultured astrocytes upregulates Nrf2 antioxidant genes during inflammation, suggesting astrocyte nAChRs are neuroprotective and decrease oxidative stress [155]. Future studies with GECIs and specific genetic approaches to selectively target astrocyte α7nAChRs will further determine the role of nicotinic receptors in astrocyte physiology and MCE dynamics.

3.4. Na+-Ca2+ Exchanger

3.4.1. Astrocyte Na+-Ca2+ Exchanger Expression

Astrocytes express the Na+/Ca2+ exchanger (NCX), which has an important role in buffering intracellular Ca2+ in exchange for Na+ influx (Figure 2) [156,157,158]. Increased intracellular Na+ levels can cause NCX to reverse direction where it brings extracellular Ca2+ in for Na+ efflux and this creates Ca2+ events in astrocytes [115,125]. Importantly, NCX is primarily confined to fine peri-synaptic astrocyte processes where it is frequently localized with the Na+/K+ ATPase and glutamate transporters that work together to take up glutamate and buffer ion gradients [159,160,161]. This creates an insular compartment for Ca2+ and Na+ signalling that is potentially ideal for the localization of MCEs [158].
Several possible mechanisms increase intracellular astrocyte Na+ and trigger NCX reversal, including (a) glutamate activation of Na+-permeable ionotropic kainate or NMDA receptors [125,162,163], (b) excitatory amino acid transporters which utilize the extracellular Na+ gradient to drive synaptic glutamate uptake [14,164,165], or (c) GABA transporter (GAT-3), which also conducts Na+ into the cell during GABA uptake [46,166]. Ca2+ events due to NCX reversal may also trigger Ca2+-induced Ca2+ release from intracellular Ca2+ stores, suggesting NCX reverse function amplifies agonist-induced Ca2+ events in astrocytes [164,166].

3.4.2. Functional Roles of Astrocyte NCX Reversal

Astrocyte NCX reversal and increased cellular Ca2+ may evoke gliotransmitter release, such as glutamate [167,168], ATP/adenosine [46], and homocysteic acid, the endogenous ligand for NMDA receptors [133]. An increase in extracellular adenosine as a result of GABA uptake and NCX reversal suppresses glutamatergic signalling by activating presynaptic adenosine receptors [46]. This is one way that NCX activity may cause astrocyte Ca2+ transients and regulate excitatory transmission. While a number of studies have attempted to model the contribution of NCX to astrocyte MCEs in fine processes [169,170,171], further work is required using GECIs to determine the role of NCX in astrocyte MCE formation and temporal dynamics, particularly regarding the rapid modulation of physiologically active circuits.

3.5. Voltage-Gated Calcium Channels

Astrocytes also express voltage-gated Ca2+ channels (VGCCs) [172,173,174], although at lower levels than neurons [175]. These channels open during membrane depolarization, permitting the influx of extracellular Ca2+. VGCCs do not contribute to spontaneous MCEs [176], but they are activated following astrocyte depolarization due to extracellular K+ uptake [174], as well as possible astrocyte NMDA receptor activity [49]. Thus, astrocyte VGCC activity is likely evoked during circuit stimulation when synaptic K+ and glutamate accumulate. Functionally, astrocyte VGCCs induce the release of glutamate [174] and possibly other gliotransmitters that maintain the heterogeneity of presynaptic strengths [49]. However, evidence linking VGCC activity with astrocyte MCEs and other functional roles of astrocytes is lacking.

4. Conclusions

Astrocytes have localized, rapid fluctuations in intracellular Ca2+ that grant them the potential to quickly regulate local processes such as synaptic activity and blood flow. By aiding the integration of information at multiple synapses and among different neuronal types in brain circuits, astrocytes may impact neuronal activity at the population level, thereby contributing to information processing and animal behaviour. Future experiments should focus on GECI tools to better identify MCEs, particularly with a fast onset during nearby synaptic activity. Additionally, pathways of extracellular calcium influx outlined here have important implications for astrocyte calcium physiology and are gaining interest in the field. Approaches to selectively target these pathways will help to better understand their contribution to rapid onset astrocyte MCEs and their functional relevance regarding neuronal network activity.

Author Contributions

Conceptualization, N.A., M.K. and J.L.S.; writing—original draft preparation, N.A. and M.K. writing—review and editing, N.A., M.K. and J.L.S.; visualization, N.A. and M.K.; supervision, J.L.S.; funding acquisition, J.L.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant and start-up funding from the University of Manitoba. The Stobart lab is also supported by grants from Brain Canada/Azrieli Foundation, Canadian Institutes for Health Research, Research Manitoba and the Manitoba Medical Service Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The figures in this paper were created with BioRender.com, accessed on 25 August 2021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Verkhratsky, A.; Nedergaard, M. Physiology of astroglia. Physiol. Rev. 2018, 98, 239–389. [Google Scholar] [CrossRef]
  2. Wallraff, A.; Köhling, R.; Heinemann, U.; Theis, M.; Willecke, K.; Steinhäuser, C. The impact of astrocytic gap junctional coupling on potassium buffering in the hippocampus. J. Neurosci. 2006, 26, 5438–5447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bröer, S.; Bröer, A.; Hansen, J.T.; Bubb, W.A.; Balcar, V.J.; Nasrallah, F.A.; Garner, B.; Rae, C. Alanine metabolism, transport, and cycling in the brain. J. Neurochem. 2007, 102, 1758–1770. [Google Scholar] [CrossRef] [PubMed]
  4. Andersen, J.V.; Markussen, K.H.; Jakobsen, E.; Schousboe, A.; Waagepetersen, H.S.; Rosenberg, P.A.; Aldana, B.I. Glutamate metabolism and recycling at the excitatory synapse in health and neurodegeneration. Neuropharmacology 2021, 196, 108719. [Google Scholar] [CrossRef] [PubMed]
  5. Weber, B.; Barros, L.F. The astrocyte: Powerhouse and recycling center. Cold Spring Harb. Perspect. Biol. 2015, 7, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Mächler, P.; Wyss, M.T.; Elsayed, M.; Stobart, J.; Gutierrez, R.; Von Faber-Castell, A.; Kaelin, V.; Zuend, M.; San Martín, A.; Romero-Gómez, I.; et al. In vivo evidence for a lactate gradient from astrocytes to neurons. Cell Metab. 2016, 23, 94–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Hertz, L.; Gibbs, M.E.; Dienel, G.A. Fluxes of lactate into, from, and among gap junction-coupled astrocytes and their interaction with noradrenaline. Front. Neurosci. 2014, 8, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Magistretti, P.J.; Allaman, I. A cellular perspective on brain energy metabolism and functional imaging. Neuron 2015, 86, 883–901. [Google Scholar] [CrossRef] [Green Version]
  9. Stobart, J.; Anderson, C.M. Multifunctional role of astrocytes as gatekeepers of neuronal energy supply. Front. Cell. Neurosci. 2013, 7, 38. [Google Scholar] [CrossRef] [Green Version]
  10. Bazargani, N.; Attwell, D. Astrocyte calcium signaling: The third wave. Nat. Neurosci. 2016, 19, 182–189. [Google Scholar] [CrossRef]
  11. Araque, A.; Carmignoto, G.; Haydon, P.G.; Oliet, S.H.R.; Robitaille, R.; Volterra, A. Gliotransmitters travel in time and space. Neuron 2014, 81, 728–739. [Google Scholar] [CrossRef] [Green Version]
  12. Attwell, D.; Buchan, A.M.; Charpak, S.; Lauritzen, M.J.; Macvicar, B.A.; Newman, E.A. Glial and neuronal control of brain blood flow. Nature 2010, 468, 232–243. [Google Scholar] [CrossRef] [Green Version]
  13. Rosenegger, D.G.; Tran, C.H.T.; Wamsteeker Cusulin, J.I.; Gordon, G.R. Tonic local brain blood flow control by astrocytes independent of phasic neurovascular coupling. J. Neurosci. 2015, 35, 13463–13474. [Google Scholar] [CrossRef]
  14. Jackson, J.G.; Robinson, M.B. Reciprocal regulation of mitochondrial dynamics and calcium signaling in astrocyte processes. J. Neurosci. 2015, 35, 15199–15213. [Google Scholar] [CrossRef] [PubMed]
  15. Agarwal, A.; Wu, P.H.; Hughes, E.G.; Fukaya, M.; Tischfield, M.A.; Langseth, A.J.; Wirtz, D.; Bergles, D.E. Transient opening of the mitochondrial permeability transition pore induces microdomain calcium transients in astrocyte processes. Neuron 2017, 93, 587–605.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yu, X.; Nagai, J.; Khakh, B.S. Improved tools to study astrocytes. Nat. Rev. Neurosci. 2020, 21, 121–138. [Google Scholar] [CrossRef] [PubMed]
  17. Stobart, J.L.; Ferrari, K.D.; Barrett, M.J.P.; Glück, C.; Stobart, M.J.; Zuend, M.; Weber, B. Cortical circuit activity evokes rapid astrocyte calcium signals on a similar timescale to neurons. Neuron 2018, 98, 726–735.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Stobart, J.L.; Ferrari, K.D.; Barrett, M.J.P.P.; Stobart, M.J.; Looser, Z.J.; Saab, A.S.; Weber, B. Long-term in vivo calcium imaging of astrocytes reveals distinct cellular compartment responses to sensory stimulation. Cereb. Cortex 2018, 28, 184–198. [Google Scholar] [CrossRef] [Green Version]
  19. Bindocci, E.; Savtchouk, I.; Liaudet, N.; Becker, D.; Carriero, G.; Volterra, A. Three-dimensional calcium imaging advances understanding of astrocyte biology. Science 2017, 356, eaai8185. [Google Scholar] [CrossRef] [Green Version]
  20. Shigetomi, E.; Tong, X.; Kwan, K.Y.; Corey, D.P.; Khakh, B.S. TRPA1 channels regulate astrocyte resting calcium and inhibitory synapse efficacy through GAT-3. Nat. Neurosci. 2011, 15, 70–80. [Google Scholar] [CrossRef] [Green Version]
  21. Arizono, M.; Inavalli, V.V.G.K.; Panatier, A.; Pfeiffer, T.; Angibaud, J.; Levet, F.; Ter Veer, M.J.T.; Stobart, J.; Bellocchio, L.; Mikoshiba, K.; et al. Structural basis of astrocytic Ca2+ signals at tripartite synapses. Nat. Commun. 2020, 11, 1906. [Google Scholar] [CrossRef] [Green Version]
  22. Shigetomi, E.; Kracun, S.; Sofroniew, M.V.; Khakh, B.S. A genetically targeted optical sensor to monitor calcium signals in astrocyte processes. Nat. Neurosci. 2010, 13, 759–766. [Google Scholar] [CrossRef] [Green Version]
  23. Shigetomi, E.; Kracun, S.; Khakh, B.S. Monitoring astrocyte calcium levels with improved membrane targeted GCaMP reporters. Neuron Glia Biol. 2010, 6, 183–191. [Google Scholar] [CrossRef] [Green Version]
  24. Srinivasan, R.; Huang, B.S.; Venugopal, S.; Johnston, A.D.; Chai, H.; Zeng, H.; Golshani, P.; Khakh, B.S. Ca(2+) signaling in astrocytes from Ip3r2(-/-) mice in brain slices and during startle responses in vivo. Nat. Neurosci. 2015, 18, 708–717. [Google Scholar] [CrossRef] [Green Version]
  25. Shigetomi, E.; Jackson-Weaver, O.; Huckstepp, R.T.; O’Dell, T.J.; Khakh, B.S. TRPA1 channels are regulators of astrocyte basal calcium levels and long-term potentiation via constitutive D-serine release. J. Neurosci. 2013, 33, 10143–10153. [Google Scholar] [CrossRef]
  26. King, C.M.; Bohmbach, K.; Minge, D.; Delekate, A.; Zheng, K.; Reynolds, J.; Rakers, C.; Zeug, A.; Petzold, G.C.; Rusakov, D.A.; et al. Local resting Ca2+ controls the scale of astroglial Ca2+ signals. Cell Rep. 2020, 30, 3466–3477.e4. [Google Scholar] [CrossRef] [Green Version]
  27. Zheng, K.; Bard, L.; Reynolds, J.P.; King, C.; Jensen, T.P.; Gourine, A.V.; Rusakov, D.A. Time-resolved imaging reveals heterogeneous landscapes of nanomolar Ca2+ in neurons and astroglia. Neuron 2015, 88, 277–288. [Google Scholar] [CrossRef] [Green Version]
  28. Otsu, Y.; Couchman, K.; Lyons, D.G.; Collot, M.; Agarwal, A.; Mallet, J.M.; Pfrieger, F.W.; Bergles, D.E.; Charpak, S. Calcium dynamics in astrocyte processes during neurovascular coupling. Nat. Neurosci. 2015, 18, 210–218. [Google Scholar] [CrossRef] [PubMed]
  29. Asada, A.; Ujita, S.; Nakayama, R.; Oba, S.; Ishii, S.; Matsuki, N.; Ikegaya, Y. Subtle modulation of ongoing calcium dynamics in astrocytic microdomains by sensory inputs. Physiol. Rep. 2015, 3, e12454. [Google Scholar] [CrossRef] [PubMed]
  30. Lind, B.L.; Jessen, S.B.; Lønstrup, M.; Joséphine, C.; Bonvento, G.; Lauritzen, M. Fast Ca2+ responses in astrocyte end-feet and neurovascular coupling in mice. Glia 2017, 66, 348–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Lind, B.L.; Brazhe, A.R.; Jessen, S.B.; Tan, F.C.C.; Lauritzen, M.J. Rapid stimulus-evoked astrocyte Ca2+ elevations and hemodynamic responses in mouse somatosensory cortex in vivo. Proc. Natl. Acad. Sci. USA 2013, 110, E4678–E4687. [Google Scholar] [CrossRef] [Green Version]
  32. Nizar, K.; Uhlirova, H.; Tian, P.; Saisan, P.a.; Cheng, Q.; Reznichenko, L.; Weldy, K.L.; Steed, T.C.; Sridhar, V.B.; MacDonald, C.L.; et al. In vivo stimulus-induced vasodilation occurs without IP3 receptor activation and may precede astrocytic calcium increase. J. Neurosci. 2013, 33, 8411–8422. [Google Scholar] [CrossRef] [PubMed]
  33. Ding, F.; O’Donnell, J.; Thrane, A.S.; Zeppenfeld, D.; Kang, H.; Xie, L.; Wang, F.; Nedergaard, M. α1-adrenergic receptors mediate coordinated Ca2+ signaling of cortical astrocytes in awake, behaving mice. Cell Calcium 2013, 54, 387–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, X.; Lou, N.; Xu, Q.; Tian, G.-F.; Peng, W.G.; Han, X.; Kang, J.; Takano, T.; Nedergaard, M. Astrocytic Ca2+ signaling evoked by sensory stimulation in vivo. Nat. Neurosci. 2006, 9, 816–823. [Google Scholar] [CrossRef] [PubMed]
  35. Winship, I.R.; Plaa, N.; Murphy, T.H. Rapid astrocyte calcium signals correlate with neuronal activity and onset of the hemodynamic response in vivo. J. Neurosci. 2007, 27, 6268–6272. [Google Scholar] [CrossRef] [Green Version]
  36. Fellin, T.; Pascual, O.; Gobbo, S.; Pozzan, T.; Haydon, P.G.; Carmignoto, G. Neuronal synchrony mediated by astrocytic glutamate through activation of extrasynaptic NMDA receptors. Neuron 2004, 43, 729–743. [Google Scholar] [CrossRef] [Green Version]
  37. Lee, S.; Yoon, B.; Berglund, K.; Oh, S.; Park, H.; Shin, H.; Augustine, G.J.; Lee, C.J. Channel-mediated tonic GABA release from glia. Science 2010, 330, 790–796. [Google Scholar] [CrossRef]
  38. Kozlov, A.S.; Angulo, M.C.; Audinat, E.; Charpak, S. Target cell-specific modulation of neuronal activity by astrocytes. Proc. Natl. Acad. Sci. USA 2006, 103, 10058–10063. [Google Scholar] [CrossRef] [Green Version]
  39. Panatier, A.; Vallée, J.; Haber, M.; Murai, K.K.; Lacaille, J.C.; Robitaille, R. Astrocytes are endogenous regulators of basal transmission at central synapses. Cell 2011, 146, 785–798. [Google Scholar] [CrossRef] [Green Version]
  40. Matos, M.; Bosson, A.; Riebe, I.; Reynell, C.; Vallée, J.; Laplante, I.; Panatier, A.; Robitaille, R.; Lacaille, C. Astrocytes detect and upregulate transmission at inhibitory synapses of somatostatin interneurons onto pyramidal cells. Nat. Commun. 2018, 9, 4254. [Google Scholar] [CrossRef]
  41. Henneberger, C.; Papouin, T.; Oliet, S.H.R.; Rusakov, D. a Long-term potentiation depends on release of D-serine from astrocytes. Nature 2010, 463, 232–236. [Google Scholar] [CrossRef]
  42. Papouin, T.; Henneberger, C.; Rusakov, D.A.; Oliet, S.H.R. Astroglial versus neuronal D-Serine: Fact checking. Trends Neurosci. 2017, 40, 517–520. [Google Scholar] [CrossRef] [Green Version]
  43. Wolosker, H.; Balu, D.T.; Coyle, J.T. Astroglial versus neuronal D-Serine: Check your controls! Trends Neurosci. 2017, 40, 520–522. [Google Scholar] [CrossRef]
  44. Wolosker, H.; Balu, D.T.; Coyle, J.T. The rise and fall of the D-Serine-mediated gliotransmission hypothesis. Trends Neurosci. 2016, 39, 712–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Di Castro, M.A.; Chuquet, J.; Liaudet, N.; Bhaukaurally, K.; Santello, M.; Bouvier, D.; Tiret, P.; Volterra, A. Local Ca2+ detection and modulation of synaptic release by astrocytes. Nat. Neurosci. 2011, 14, 1276–1284. [Google Scholar] [CrossRef] [PubMed]
  46. Boddum, K.; Jensen, T.P.; Magloire, V.; Kristiansen, U.; Rusakov, D.A.; Pavlov, I.; Walker, M.C. Astrocytic GABA transporter activity modulates excitatory neurotransmission. Nat. Commun. 2016, 7, 13572. [Google Scholar] [CrossRef] [PubMed]
  47. De Pittà, M.; Brunel, N.; Volterra, A. Astrocytes: Orchestrating synaptic plasticity? Neuroscience 2016, 323, 43–61. [Google Scholar] [CrossRef] [Green Version]
  48. Kastanenka, K.V.; Moreno-Bote, R.; De Pittà, M.; Perea, G.; Eraso-Pichot, A.; Masgrau, R.; Poskanzer, K.E.; Galea, E. A roadmap to integrate astrocytes into Systems Neuroscience. Glia 2020, 68, 5–26. [Google Scholar] [CrossRef] [Green Version]
  49. Letellier, M.; Park, Y.K.; Chater, T.E.; Chipman, P.H.; Gautam, S.G.; Oshima-Takago, T.; Goda, Y. Astrocytes regulate heterogeneity of presynaptic strengths in hippocampal networks. Proc. Natl. Acad. Sci. USA 2016, 113, E2685–E2694. [Google Scholar] [CrossRef] [Green Version]
  50. Navarrete, M.; Perea, G.; de Sevilla, D.F.; Gómez-Gonzalo, M.; Núñez, A.; Martín, E.D.; Araque, A. Astrocytes mediate in vivo cholinergic-induced synaptic plasticity. PLoS Biol. 2012, 10, e1001259. [Google Scholar] [CrossRef] [Green Version]
  51. Mu, Y.; Bennett, D.V.; Rubinov, M.; Narayan, S.; Yang, C.-T.; Tanimoto, M.; Mensh, B.D.; Looger, L.L.; Ahrens, M.B. Glia accumulate evidence that actions are futile and suppress unsuccessful behavior. Cell 2019, 178, 27–43.e19. [Google Scholar] [CrossRef] [Green Version]
  52. Adamsky, A.; Kol, A.; Kreisel, T.; Doron, A.; Ozeri-Engelhard, N.; Melcer, T.; Refaeli, R.; Horn, H.; Regev, L.; Groysman, M.; et al. Astrocytic activation generates de novo neuronal potentiation and memory enhancement. Cell 2018, 174, 59–71.e14. [Google Scholar] [CrossRef] [Green Version]
  53. Yu, X.; Taylor, A.M.W.; Nagai, J.; Golshani, P.; Evans, C.J.; Coppola, G.; Khakh, B.S. Reducing astrocyte calcium signaling in vivo alters striatal microcircuits and causes repetitive behavior. Neuron 2018, 99, 1170–1187. [Google Scholar] [CrossRef] [Green Version]
  54. Cao, X.; Li, L.P.; Wang, Q.; Wu, Q.; Hu, H.H.; Zhang, M.; Fang, Y.Y.; Zhang, J.; Li, S.J.; Xiong, W.C.; et al. Astrocyte-derived ATP modulates depressive-like behaviors. Nat. Med. 2013, 19, 773–777. [Google Scholar] [CrossRef]
  55. Wang, Q.; Kong, Y.; Wu, D.; Liu, J.-H.; Jie, W.; You, Q.; Huang, L.; Hu, J.; Chu, H.; Gao, F.; et al. Impaired calcium signaling in astrocytes modulates autism spectrum disorder-like behaviors in mice. Nat. Commun. 2021, 12, 3321. [Google Scholar] [CrossRef]
  56. Bonansco, C.; Couve, A.; Perea, G.; Ferradas, C.Á.; Roncagliolo, M.; Fuenzalida, M. Glutamate released spontaneously from astrocytes sets the threshold for synaptic plasticity. Eur. J. Neurosci. 2011, 33, 1483–1492. [Google Scholar] [CrossRef] [PubMed]
  57. Jourdain, P.; Bergersen, L.H.; Bhaukaurally, K.; Bezzi, P.; Santello, M.; Domercq, M.; Matute, C.; Tonello, F.; Gundersen, V.; Volterra, A. Glutamate exocytosis from astrocytes controls synaptic strength. Nat. Neurosci. 2007, 10, 331–339. [Google Scholar] [CrossRef] [PubMed]
  58. Gómez-Gonzalo, M.; Navarrete, M.; Perea, G.; Covelo, A.; Martín-Fernández, M.; Shigemoto, R.; Luján, R.; Araque, A. Endocannabinoids induce lateral long-term potentiation of transmitter release by stimulation of gliotransmission. Cereb. Cortex 2015, 25, 3699–3712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Perea, G.; Araque, A. Astrocytes potentiate transmitter release at single hippocampal synapses. Science 2007, 317, 1083–1086. [Google Scholar] [CrossRef] [PubMed]
  60. Pascual, O.; Casper, K.B.; Kubera, C.; Zhang, J.; Revilla-Sanchez, R.; Sul, J.-Y.Y.; Takano, H.; Moss, S.J.; McCarthy, K.; Haydon, P.G. Astrocytic purinergic signaling coordinates synaptic networks. Science 2005, 310, 113–116. [Google Scholar] [CrossRef]
  61. Min, R.; Nevian, T. Astrocyte signaling controls spike timing-dependent depression at neocortical synapses. Nat. Neurosci. 2012, 15, 746–753. [Google Scholar] [CrossRef]
  62. Fossat, P.; Turpin, F.R.; Sacchi, S.; Dulong, J.; Shi, T.; Rivet, J.M.; Sweedler, J.V.; Pollegioni, L.; Millan, M.J.; Oliet, S.H.R.; et al. Glial D-serine gates NMDA receptors at excitatory synapses in prefrontal cortex. Cereb. Cortex 2012, 22, 595–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chen, N.; Sugihara, H.; Sharma, J.; Perea, G.; Petravicz, J.; Le, C.; Sur, M. Nucleus basalis-enabled stimulus-specific plasticity in the visual cortex is mediated by astrocytes. Proc. Natl. Acad. Sci. USA 2012, 109, E2832–E2841. [Google Scholar] [CrossRef] [Green Version]
  64. Takata, N.; Mishima, T.; Hisatsune, C.; Nagai, T.; Ebisui, E.; Mikoshiba, K.; Hirase, H. Astrocyte calcium signaling transforms cholinergic modulation to cortical plasticity in vivo. J. Neurosci. 2011, 31, 18155–18165. [Google Scholar] [CrossRef] [PubMed]
  65. Bernardinelli, Y.; Randall, J.; Janett, E.; Nikonenko, I.; König, S.; Jones, E.V.; Flores, C.E.; Murai, K.K.; Bochet, C.G.; Holtmaat, A.; et al. Activity-dependent structural plasticity of perisynaptic astrocytic domains promotes excitatory synapse stability. Curr. Biol. 2014, 24, 1679–1688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Molotkov, D.; Zobova, S.; Arcas, J.M.; Khiroug, L. Calcium-induced outgrowth of astrocytic peripheral processes requires actin binding by Profilin-1. Cell Calcium 2013, 53, 338–348. [Google Scholar] [CrossRef] [PubMed]
  67. Tanaka, M.; Shih, P.-Y.; Gomi, H.; Yoshida, T.; Nakai, J.; Ando, R.; Furuichi, T.; Mikoshiba, K.; Semyanov, A.; Itohara, S. Astrocytic Ca2+ signals are required for the functional integrity of tripartite synapses. Mol. Brain 2013, 6, 6. [Google Scholar] [CrossRef] [Green Version]
  68. Bernardinelli, Y.; Muller, D.; Nikonenko, I. Astrocyte-synapse structural plasticity. Neural Plast. 2014, 2014, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Poskanzer, K.E.; Yuste, R. Astrocytic regulation of cortical UP states. Proc. Natl. Acad. Sci. USA 2011, 108, 18453–18458. [Google Scholar] [CrossRef] [Green Version]
  70. Poskanzer, K.E.; Yuste, R. Astrocytes regulate cortical state switching in vivo. Proc. Natl. Acad. Sci. USA 2016, 113, E2675–E2684. [Google Scholar] [CrossRef] [Green Version]
  71. Paukert, M.; Agarwal, A.; Cha, J.; Doze, V.A.; Kang, J.U.; Bergles, D.E. Norepinephrine controls astroglial responsiveness to local circuit activity. Neuron 2014, 82, 1263–1270. [Google Scholar] [CrossRef] [Green Version]
  72. Tran, C.H.T.; Peringod, G.; Gordon, G.R. Astrocytes integrate behavioral state and vascular signals during functional hyperemia. Neuron 2018, 100, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Haidey, J.N.; Peringod, G.; Institoris, A.; Gorzo, K.A.; Nicola, W.; Vandal, M.; Ito, K.; Liu, S.; Fielding, C.; Visser, F.; et al. Astrocytes regulate ultra-slow arteriole oscillations via stretch-mediated TRPV4-COX-1 feedback. Cell Rep. 2021, 36, 109405. [Google Scholar] [CrossRef] [PubMed]
  74. LeMaistre Stobart, J.L.; Lu, L.; Anderson, H.D.I.; Mori, H.; Anderson, C.M. Astrocyte-induced cortical vasodilation is mediated by D-serine and endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. USA 2013, 110, 3149–3154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Gordon, G.G.; Choi, H.B.; Rungta, R.L.; Ellis-Davies, G.C.R.; MacVicar, B.A. Brain metabolism dictates the polarity of astrocyte control over arterioles. Nature 2008, 456, 745–749. [Google Scholar] [CrossRef] [Green Version]
  76. Mulligan, S.J.; MacVicar, B.A. Calcium transients in astrocyte endfeet cause cerebrovascular constrictions. Nature 2004, 431, 195–199. [Google Scholar] [CrossRef] [PubMed]
  77. Zonta, M.; Angulo, M.C.; Gobbo, S.; Rosengarten, B.; Hossmann, K.-A.A.; Pozzan, T.; Carmignoto, G. Neuron-to-astrocyte signaling is central to the dynamic control of brain microcirculation. Nat. Neurosci. 2003, 6, 43–50. [Google Scholar] [CrossRef]
  78. Sharma, K.; Gordon, G.R.J.; Tran, C.H.T. Heterogeneity of sensory-induced astrocytic Ca2+ dynamics during functional hyperemia. Front. Physiol. 2020, 11, 611884. [Google Scholar] [CrossRef]
  79. Gu, X.; Chen, W.; Volkow, N.D.; Koretsky, A.P.; Du, C.; Pan, Y. Synchronized astrocytic Ca2+ responses in neurovascular coupling during somatosensory stimulation and for the resting state. Cell Rep. 2018, 23, 3878–3890. [Google Scholar] [CrossRef]
  80. Tran, C.H.T.; George, A.G.; Teskey, G.C.; Gordon, G.R. Seizures elevate gliovascular unit Ca2+ and cause sustained vasoconstriction. JCI Insight 2020, 5, e136469. [Google Scholar] [CrossRef]
  81. Volnova, A.; Tsytsarev, V.; Ptukha, M.; Inyushin, M. In vitro and in vivo study of the short-term vasomotor response during epileptic seizures. Brain Sci. 2020, 10, 942. [Google Scholar] [CrossRef] [PubMed]
  82. Grutzendler, J.; Nedergaard, M. Cellular control of brain capillary blood flow: In vivo imaging veritas. Trends Neurosci. 2019, 42, 528–536. [Google Scholar] [CrossRef]
  83. Dunn, K.M.; Hill-Eubanks, D.C.; Liedtke, W.B.; Nelson, M.T. TRPV4 channels stimulate Ca2+-induced Ca2+ release in astrocytic endfeet and amplify neurovascular coupling responses. Proc. Natl. Acad. Sci. USA 2013, 110, 6157–6162. [Google Scholar] [CrossRef] [Green Version]
  84. Malarkey, E.B.; Ni, Y.; Parpura, V. Ca2+ entry through TRPC1 channels contributes to intracellular Ca2+ dynamics and consequent glutamate release from rat astrocytes. Glia 2008, 56, 821–835. [Google Scholar] [CrossRef]
  85. Butenko, O.; Dzamba, D.; Benesova, J.; Honsa, P.; Benfenati, V.; Rusnakova, V.; Ferroni, S.; Anderova, M. The increased activity of TRPV4 channel in the astrocytes of the adult rat hippocampus after cerebral Hypoxia/Ischemia. PLoS ONE 2012, 7, e39959. [Google Scholar] [CrossRef] [Green Version]
  86. Benfenati, V.; Amiry-Moghaddam, M.; Caprini, M.; Mylonakou, M.N.; Rapisarda, C.; Ottersen, O.P.; Ferroni, S. Expression and functional characterization of transient receptor potential vanilloid-related channel 4 (TRPV4) in rat cortical astrocytes. Neuroscience 2007, 148, 876–892. [Google Scholar] [CrossRef]
  87. Pizzo, P.; Burgo, A.; Pozzan, T.; Fasolato, C. Role of capacitative calcium entry on glutamate-induced calcium influx in type-I rat cortical astrocytes. J. Neurochem. 2001, 79, 98–109. [Google Scholar] [CrossRef]
  88. Grimaldi, M.; Maratos, M.; Verma, A. Transient receptor potential channel activation causes a novel form of [Ca2+]i oscillations and is not involved in capacitative Ca2+ entry in glial cells. J. Neurosci. 2003, 23, 4737–4745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Agulhon, C.; Fiacco, T.A.; McCarthy, K.D. Hippocampal short- and long-term plasticity are not modulated by astrocyte Ca2+ signaling. Science 2010, 327, 1250–1254. [Google Scholar] [CrossRef] [Green Version]
  90. Takata, N.; Nagai, T.; Ozawa, K.; Oe, Y.; Mikoshiba, K.; Hirase, H. Cerebral blood flow modulation by basal forebrain or whisker stimulation can occur independently of large cytosolic Ca2+ signaling in astrocytes. PLoS ONE 2013, 8, 4–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Petravicz, J.; Fiacco, T.a.; McCarthy, K.D. Loss of IP3 receptor-dependent Ca2+ increases in hippocampal astrocytes does not affect baseline CA1 pyramidal neuron synaptic activity. J. Neurosci. 2008, 28, 4967–4973. [Google Scholar] [CrossRef] [Green Version]
  92. Porter, J.T.; McCarthy, K.D. Hippocampal astrocytes in situ respond to glutamate released from synaptic terminals. J. Neurosci. 1996, 16, 5073–5081. [Google Scholar] [CrossRef] [Green Version]
  93. Pasti, L.; Volterra, A.; Pozzan, T.; Carmignoto, G. Intracellular calcium oscillations in astrocytes: A highly plastic, bidirectional form of communication between neurons and astrocytes in situ. J. Neurosci. 1997, 17, 7817–7830. [Google Scholar] [CrossRef] [Green Version]
  94. Cahoy, J.D.; Emery, B.; Kaushal, A.; Foo, L.C.; Zamanian, J.L.; Christopherson, K.S.; Xing, Y.; Lubischer, J.L.; Krieg, P.A.; Krupenko, S.A.; et al. A transcriptome database for astrocytes, neurons, and oligodendrocytes: A new resource for understanding brain development and function. J. Neurosci. 2008, 28, 264–278. [Google Scholar] [CrossRef] [Green Version]
  95. Sun, W.; McConnell, E.; Pare, J.-F.F.; Xu, Q.; Chen, M.; Peng, W.; Lovatt, D.; Han, X.; Smith, Y.; Nedergaard, M.; et al. Glutamate-dependent neuroglial calcium signaling differs between young and adult brain. Science 2013, 339, 197–200. [Google Scholar] [CrossRef] [Green Version]
  96. Lavialle, M.; Aumann, G.; Anlauf, E.; Pröls, F.; Arpin, M.; Derouiche, A. Structural plasticity of perisynaptic astrocyte processes involves ezrin and metabotropic glutamate receptors. Proc. Natl. Acad. Sci. USA 2011, 108, 12915–12919. [Google Scholar] [CrossRef] [Green Version]
  97. Collingridge, G.L.; Olsen, R.W.; Peters, J.; Spedding, M. A nomenclature for ligand-gated ion channels. Neuropharmacology 2009, 56, 2–5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Guo, C.; Ma, Y.-Y. Calcium permeable-AMPA receptors and excitotoxicity in neurological disorders. Front. Neural Circuits 2021, 15, 711564. [Google Scholar] [CrossRef]
  99. Frandsen, A.; Schousboe, A. AMPA receptor-mediated neurotoxicity: Role of Ca2+ and desensitization. Neurochem. Res. 2003, 28, 1495–1499. [Google Scholar] [CrossRef] [PubMed]
  100. Paoletti, P.; Bellone, C.; Zhou, Q. NMDA receptor subunit diversity: Impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci. 2013, 14, 383–400. [Google Scholar] [CrossRef] [PubMed]
  101. Kehoe, L.A.; Bernardinelli, Y.; Muller, D. GluN3A: An NMDA receptor subunit with exquisite properties and functions. Neural Plast. 2013, 2013, 145387. [Google Scholar] [CrossRef] [Green Version]
  102. Grand, T.; Abi Gerges, S.; David, M.; Diana, M.A.; Paoletti, P. Unmasking GluN1/GluN3A excitatory glycine NMDA receptors. Nat. Commun. 2018, 9, 4769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Verkhratsky, A.; Steinhäuser, C. Ion channels in glial cells. Brain Res. Rev. 2000, 32, 380–412. [Google Scholar] [CrossRef]
  104. Conti, F.; Minelli, A.; Brecha, N. Cellular localization and laminar distribution of AMPA glutamate receptor subunits mRNAs and proteins in the rat cerebral cortex. J. Comp. Neurol. 1994, 350, 241–259. [Google Scholar] [CrossRef] [PubMed]
  105. Seifert, G.; Steinhäuser, C. Glial cells in the mouse hippocampus express AMPA receptors with an intermediate Ca2+ permeability. Eur. J. Neurosci. 1995, 7, 1872–1881. [Google Scholar] [CrossRef]
  106. Fan, D.; Grooms, S.Y.; Araneda, R.C.; Johnson, A.B.; Dobrenis, K.; Kessler, J.A.; Zukin, R.S. AMPA receptor protein expression and function in astrocytes cultured from hippocampus. J. Neurosci. Res. 1999, 57, 557–571. [Google Scholar] [CrossRef]
  107. Seifert, G.; Zhou, M.; Steinhäuser, C. Analysis of AMPA receptor properties during postnatal development of mouse hippocampal astrocytes. J. Neurophysiol. 1997, 78, 2916–2923. [Google Scholar] [CrossRef]
  108. Palygin, O.; Lalo, U.; Pankratov, Y. Distinct pharmacological and functional properties of NMDA receptors in mouse cortical astrocytes. Br. J. Pharmacol. 2011, 163, 1755–1766. [Google Scholar] [CrossRef] [Green Version]
  109. Palygin, O.; Lalo, U.; Verkhratsky, A.; Pankratov, Y. Ionotropic NMDA and P2X1/5 receptors mediate synaptically induced Ca2+signalling in cortical astrocytes. Cell Calcium 2010, 48, 225–231. [Google Scholar] [CrossRef]
  110. Dzamba, D.; Honsa, P.; Valny, M.; Kriska, J.; Valihrach, L.; Novosadova, V.; Kubista, M.; Anderova, M. Quantitative analysis of glutamate receptors in glial cells from the cortex of GFAP/EGFP mice following ischemic injury: Focus on NMDA receptors. Cell. Mol. Neurobiol. 2015, 35, 1187–1202. [Google Scholar] [CrossRef]
  111. García-Barcina, J.; Matute, C. Expression of kainate-selective glutamate receptor subunits in glial cells of the adult bovine white matter. Eur. J. Neurosci. 1996, 8, 2379–2387. [Google Scholar] [CrossRef] [PubMed]
  112. Brand-Schieber, E.; Lowery, S.; Werner, P. Select ionotropic glutamate AMPA/kainate receptors are expressed at the astrocyte-vessel interface. Brain Res. 2004, 1007, 178–182. [Google Scholar] [CrossRef] [PubMed]
  113. Pearce, B.; Albrecht, J.; Morrow, C.; Murphy, S. Astrocyte glutamate receptor activation promotes inositol phospholipid turnover and calcium flux. Neurosci. Lett. 1986, 72, 335–340. [Google Scholar] [CrossRef]
  114. Cornell-Bell, A.H.; Finkbeiner, S.M.; Cooper, M.S.; Smith, S.J. Glutamate induces calcium waves in cultured astrocytes: Long-range glial signaling. Science 1990, 247, 470–473. [Google Scholar] [CrossRef] [PubMed]
  115. Jensen, A.M.; Chiu, S.Y. Fluorescence measurement of changes in intracellular calcium induced by excitatory amino acids in cultured cortical astrocytes. J. Neurosci. 1990, 10, 1165–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Burnashev, N.; Khodorova, A.; Jonas, P.; Helm, P.; Wisden, W.; Monyer, H.; Seeburg, P.; Sakmann, B. Calcium-permeable AMPA-kainate receptors in fusiform cerebellar glial cells. Science 1992, 256, 1566–1570. [Google Scholar] [CrossRef]
  117. Lalo, U.; Pankratov, Y.; Parpura, V.; Verkhratsky, A. Ionotropic receptors in neuronal-astroglial signalling: What is the role of “excitable” molecules in non-excitable cells. Biochim. Biophys. Acta 2011, 1813, 992–1002. [Google Scholar] [CrossRef]
  118. Cai, Z.; Kimelberg, H.K. Glutamate receptor-mediated calcium responses in acutely isolated hippocampal astrocytes. Glia 1997, 21, 380–389. [Google Scholar] [CrossRef]
  119. Papadia, S.; Soriano, F.X.; Léveillé, F.; Martel, M.-A.; Dakin, K.A.; Hansen, H.H.; Kaindl, A.; Sifringer, M.; Fowler, J.; Stefovska, V.; et al. Synaptic NMDA receptor activity boosts intrinsic antioxidant defenses. Nat. Neurosci. 2008 114 2008, 11, 476–487. [Google Scholar] [CrossRef]
  120. Wroge, C.M.; Hogins, J.; Eisenman, L.; Mennerick, S. Synaptic NMDA receptors mediate hypoxic excitotoxic death. J. Neurosci. 2012, 32, 6732–6742. [Google Scholar] [CrossRef] [Green Version]
  121. Jimenez-Blasco, D.; Santofimia-Castanõ, P.; Gonzalez, A.; Almeida, A.; Bolanõs, J.P. Astrocyte NMDA receptors’ activity sustains neuronal survival through a Cdk5-Nrf2 pathway. Cell Death Differ. 2015, 22, 1877–1889. [Google Scholar] [CrossRef] [Green Version]
  122. Glaum, S.R.; Holzwarth, J.A.; Miller, R.J. Glutamate receptors activate Ca2+ mobilization and Ca2+ influx into astrocytes. Proc. Natl. Acad. Sci. USA 1990, 87, 3454–3458. [Google Scholar] [CrossRef] [Green Version]
  123. Kou, Z.W.; Mo, J.L.; Wu, K.W.; Qiu, M.H.; Huang, Y.L.; Tao, F.; Lei, Y.; Lv, L.L.; Sun, F.Y. Vascular endothelial growth factor increases the function of calcium-impermeable AMPA receptor GluA2 subunit in astrocytes via activation of protein kinase C signaling pathway. Glia 2019, 67, 1344–1358. [Google Scholar] [CrossRef]
  124. Hamilton, N.; Vayro, S.; Kirchhoff, F.; Verkhratsky, A.; Robbins, J.; Gorecki, D.C.; Butt, A.M. Mechanisms of ATP- and glutamate-mediated calcium signaling in white matter astrocytes. Glia 2008, 56, 734–749. [Google Scholar] [CrossRef] [PubMed]
  125. Ziemens, D.; Oschmann, F.; Gerkau, N.J.; Rose, C.R. Heterogeneity of activity-induced sodium transients between astrocytes of the mouse hippocampus and neocortex: Mechanisms and consequences. J. Neurosci. 2019, 39, 2620–2634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Lange, S.C.; Bak, L.K.; Waagepetersen, H.S.; Schousboe, A.; Norenberg, M.D. Primary cultures of astrocytes: Their value in understanding astrocytes in health and disease. Neurochem. Res. 2012, 37, 2569–2588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Shelton, M.K.; McCarthy, K. Mature hippocampal astrocytes exhibit functional metabotropic and ionotropic glutamate receptors in situ. Glia 1999, 26, 1–11. [Google Scholar] [CrossRef]
  128. Schipke, C.G.; Ohlemeyer, C.; Matyash, M.; Nolte, C.; Kettenmann, H.; Kirchhoff, F. Astrocytes of the mouse neocortex express functional N-methyl-D-aspartate receptors. FASEB J. 2001, 15, 1270–1272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Serrano, A.; Robitaille, R.; Lacaille, J.C. Differential NMDA-dependent activation of glial cells in mouse hippocampus. Glia 2008, 56, 1648–1663. [Google Scholar] [CrossRef]
  130. Lalo, U.; Palygin, O.; Rasooli-Nejad, S.; Andrew, J.; Haydon, P.G.; Pankratov, Y. Exocytosis of ATP from astrocytes modulates phasic and tonic inhibition in the neocortex. PLoS Biol. 2014, 12, e1001747. [Google Scholar] [CrossRef]
  131. Mehina, E.M.F.; Murphy-Royal, C.; Gordon, G.R. Steady-state free Ca2+ in astrocytes is decreased by experience and impacts arteriole tone. J. Neurosci. 2017, 37, 8150–8165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Lalo, U.; Palygin, O.; North, R.A.; Verkhratsky, A.; Pankratov, Y. Age-dependent remodelling of ionotropic signalling in cortical astroglia. Aging Cell 2011, 10, 392–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Benz, B.; Grima, G.; Do, K.Q. Glutamate-induced homocysteic acid release from astrocytes: Possible implication in glia-neuron signaling. Neuroscience 2004, 124, 377–386. [Google Scholar] [CrossRef] [PubMed]
  134. Droste, D.; Seifert, G.; Seddar, L.; Jädtke, O.; Steinhaüser, C.; Lohr, C. Ca2+-permeable AMPA receptors in mouse olfactory bulb astrocytes. Sci. Rep. 2017, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  135. Haustein, M.D.; Kracun, S.; Lu, X.H.; Shih, T.; Jackson-Weaver, O.; Tong, X.; Xu, J.; Yang, X.W.; O’Dell, T.J.; Marvin, J.S.; et al. Conditions and constraints for astrocyte calcium signaling in the hippocampal mossy fiber pathway. Neuron 2014, 82, 413–429. [Google Scholar] [CrossRef] [Green Version]
  136. Saab, A.S.; Neumeyer, A.; Jahn, H.M.; Cupido, A.; Šimek, A.A.M.; Boele, H.J.; Scheller, A.; Le Meur, K.; Götz, M.; Monyer, H.; et al. Bergmann glial AMPA receptors are required for fine motor coordination. Science 2012, 337, 749–753. [Google Scholar] [CrossRef] [Green Version]
  137. Iino, M.; Goto, K.; Kakegawa, W.; Okado, H.; Sudo, M.; Ishiuchi, S.; Miwa, A.; Takayasu, Y.; Saito, I.; Tsuzuki, K.; et al. Glia-synapse interaction through Ca2+-permeable AMPA receptors in Bergmann glia. Science 2001, 292, 926–929. [Google Scholar] [CrossRef]
  138. Olney, J.W. Neurotoxicity of NMDA receptor antagonists: An overiew. Psychopharmacol. Bull. 1994, 30, 533–540. [Google Scholar]
  139. Chipman, P.H.; Fernandez, A.P.; Fung, C.C.A.; Tedoldi, A.; Kawai, A.; Gautam, S.G.; Kurosawa, M.; Abe, M.; Sakimura, K.; Fukai, T.; et al. Astrocyte GluN2C NMDA receptors control basal synaptic strengths of hippocampal CA1 pyramidal neurons in the stratum radiatum. bioRxiv 2021. [Google Scholar] [CrossRef]
  140. Lalo, U.; Pankratov, Y.; Wichert, S.P.; Rossner, M.J.; North, R.A.; Kirchhoff, F.; Verkhratsky, A. P2X1 and P2X5 subunits form the functional P2X receptor in mouse cortical astrocytes. J. Neurosci. 2008, 28, 5473–5480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Fischer, W.; Appelt, K.; Grohmann, M.; Franke, H.; Nörenberg, W.; Illes, P. Increase of intracellular Ca2+ by P2X and P2Y receptor-subtypes in cultured cortical astroglia of the rat. Neuroscience 2009, 160, 767–783. [Google Scholar] [CrossRef]
  142. North, R.A. Molecular physiology of P2X receptors. Physiol. Rev. 2002, 82, 1013–1067. [Google Scholar] [CrossRef]
  143. Surprenant, A.; North, R.A. Signaling at purinergic P2X receptors. Annu. Rev. Physiol. 2009, 71, 333–359. [Google Scholar] [CrossRef] [Green Version]
  144. Franke, H.; Verkhratsky, A.; Burnstock, G.; Illes, P. Pathophysiology of astroglial purinergic signalling. Purinergic Signal. 2012, 8, 629–657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Franke, H.; Gunther, A.; Grosche, J.; Schmidt, R.; Rossner, S.; Reinhardt, R.; Faber-Zuschratter, H.; Schneider, D.; Illes, P. P2X 7 receptor expression after ischemia in the cerebral cortex of rats. J. Neuropathol. Exp. Neurol. 2004, 63, 686–699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Mishra, A.; Reynolds, J.P.; Chen, Y.; Gourine, A.V.; Rusakov, D.A.; Attwell, D. Astrocytes mediate neurovascular signaling to capillary pericytes but not to arterioles. Nat. Neurosci. 2016, 19, 1619–1627. [Google Scholar] [CrossRef] [Green Version]
  147. Lalo, U.; Bogdanov, A.; Pankratov, Y. Age- and experience-related plasticity of ATP-mediated signaling in the neocortex. Front. Cell. Neurosci. 2019, 13, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Gourine, A.V.; Kasymov, V.; Marina, N.; Tang, F.; Figueiredo, M.F.; Lane, S.; Teschemacher, A.G.; Spyer, K.M.; Deisseroth, K.; Kasparov, S.; et al. Astrocytes control breathing through pH-dependent release of ATP. Science 2010, 329, 571–575. [Google Scholar] [CrossRef] [Green Version]
  149. Lalo, U.; Bogdanov, A.; Pankratov, Y. Diversity of astroglial effects on aging- and experience-related cortical metaplasticity. Front. Mol. Neurosci. 2018, 11, 1–14. [Google Scholar] [CrossRef]
  150. Sharma, G.; Vijayaraghavan, S. Nicotinic cholinergic signaling in hippocampal astrocytes involves calcium-induced calcium release from intracellular stores. Proc. Natl. Acad. Sci. USA 2001, 98, 4148–4153. [Google Scholar] [CrossRef] [Green Version]
  151. Shen, J.; Yakel, J.L. Functional α7 nicotinic ACh receptors on Astrocytes in rat hippocampal CA1 slices. J. Mol. Neurosci. 2012, 48, 14–21. [Google Scholar] [CrossRef] [Green Version]
  152. Corradi, J.; Bouzat, C. Understanding the bases of function and modulation of a7 nicotinic receptors: Implications for drug discovery. Mol. Pharmacol. 2016, 90, 288–299. [Google Scholar] [CrossRef] [Green Version]
  153. Papouin, T.; Dunphy, J.M.; Tolman, M.; Dineley, K.T.; Haydon, P.G. Septal cholinergic neuromodulation tunes the astrocyte-dependent gating of hippocampal NMDA receptors to wakefulness. Neuron 2017, 94, 840–854.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Aryal, S.; Fu, X.; Sandin, J.N.; Neupane, K.R.; Lakes, J.E.; Grady, M.E.; Richards, C.I. Nicotine induces morphological and functional changes in astrocytes via nicotinic receptor activity. Glia 2021, 69, 2037–2053. [Google Scholar] [CrossRef] [PubMed]
  155. Patel, H.; McIntire, J.; Ryan, S.; Dunah, A.; Loring, R. Anti-inflammatory effects of astroglial α7 nicotinic acetylcholine receptors are mediated by inhibition of the NF-ΚB pathway and activation of the Nrf2 pathway. J. Neuroinflammation 2017, 14, 1–15. [Google Scholar] [CrossRef] [PubMed]
  156. Kirischuk, S.; Parpura, V.; Verkhratsky, A. Sodium dynamics: Another key to astroglial excitability? Trends Neurosci. 2012, 35, 497–506. [Google Scholar] [CrossRef]
  157. Verkhratsky, A.; Noda, M.; Parpura, V.; Kirischuk, S. Sodium fluxes and astroglial function. Adv. Exp. Med. Biol. 2013, 961, 295–305. [Google Scholar] [CrossRef]
  158. Rose, C.R.; Ziemens, D.; Verkhratsky, A. On the special role of NCX in astrocytes: Translating Na+-transients into intracellular Ca2+ signals. Cell Calcium 2020, 86, 102154. [Google Scholar] [CrossRef] [PubMed]
  159. Blaustein, M.P.; Juhaszova, M.; Golovina, V.A.; Church, P.J.; Stanley, E.F. Na/Ca exchager and PMCA localizaiton in neurons and astrocytes: Functional implications. Ann. N. Y. Acad. Sci. 2002, 976, 356–366. [Google Scholar] [CrossRef] [PubMed]
  160. Golovina, V.A.; Song, H.; James, P.F.; Lingrel, J.B.; Blaustein, M.P. Na+ pump alpha2 -subunit expression modulates Ca2+ signaling. Am. J. Physiol. - Cell Physiol. 2003, 284, C475–C486. [Google Scholar] [CrossRef] [PubMed]
  161. Lee, M.Y.; Song, H.; Nakai, J.; Ohkura, M.; Kotlikoff, M.I.; Kinsey, S.P.; Golovina, V.A.; Blaustein, M.P. Local subplasma membrane Ca2+ signals detected by a tethered Ca2+ sensor. Proc. Natl. Acad. Sci. USA 2006, 103, 13232–13237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Takuma, K.; Matsuda, T.; Hashimoto, H.; Kitanaka, J.; Asano, S.; Kishida, Y.; Baba, A. Role of Na(+)-Ca2+ exchanger in agonist-induced Ca2+ signaling in cultured rat astrocytes. J. Neurochem. 1996, 67, 1840–1845. [Google Scholar] [CrossRef] [PubMed]
  163. Goldman, W.; Yarowsky, P.; Juhaszova, M.; Krueger, B.; Blaustein, M. Sodium/calcium exchange in rat cortical astrocytes. J. Neurosci. 1994, 14, 5834–5843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Rojas, H.; Colina, C.; Ramos, M.; Benaim, G.; Jaffe, E.H.; Caputo, C.; DiPolo, R. Na+ entry via glutamate transporter activates the reverse Na+/Ca2+ exchange and triggers -induced Ca2+ release in rat cerebellar Type-1 astrocytes. J. Neurochem. 2007, 100, 1188–1202. [Google Scholar] [CrossRef]
  165. Jackson, J.G.; O’Donnell, J.C.; Takano, H.; Coulter, D.A.; Robinson, M.B. Neuronal activity and glutamate uptake decrease mitochondrial mobility in astrocytes and position mitochondria near glutamate transporters. J. Neurosci. 2014, 34, 1613–1624. [Google Scholar] [CrossRef] [Green Version]
  166. Doengi, M.; Hirnet, D.; Coulon, P.; Pape, H.-C.; Deitmer, J.W.; Lohr, C. GABA uptake-dependent Ca2+ signaling in developing olfactory bulb astrocytes. Proc. Natl. Acad. Sci. USA 2009, 106, 17570. [Google Scholar] [CrossRef] [Green Version]
  167. Reyes, R.C.; Verkhratsky, A.; Parpura, V. Plasmalemmal Na+/Ca2+ exchanger modulates Ca2+-dependent exocytotic release of glutamate from rat cortical astrocytes. ASN Neuro 2012, 4, 33–45. [Google Scholar] [CrossRef] [Green Version]
  168. Paluzzi, S.; Alloisio, S.; Zappettini, S.; Milanese, M.; Raiteri, L.; Nobile, M.; Bonanno, G. Adult astroglia is competent for Na+/Ca2+ exchanger-operated exocytotic glutamate release triggered by mild depolarization. J. Neurochem. 2007, 103, 1196–1207. [Google Scholar] [CrossRef]
  169. Brazhe, A.R.; Verisokin, A.Y.; Verveyko, D.V.; Postnov, D.E. Sodium–calcium exchanger can account for regenerative Ca2+ entry in thin astrocyte processes. Front. Cell. Neurosci. 2018, 12, 1–7. [Google Scholar] [CrossRef]
  170. Héja, L.; Kardos, J. NCX activity generates spontaneous Ca2+ oscillations in the astrocytic leaflet microdomain. Cell Calcium 2020, 86, 102137. [Google Scholar] [CrossRef]
  171. Wade, J.J.; Breslin, K.; Wong-Lin, K.F.; Harkin, J.; Flanagan, B.; Van Zalinge, H.; Hall, S.; Dallas, M.; Bithell, A.; Verkhratsky, A.; et al. Calcium microdomain formation at the perisynaptic cradle due to NCX Reversal: A computational study. Front. Cell. Neurosci. 2019, 13, 1–13. [Google Scholar] [CrossRef] [Green Version]
  172. MacVicar, B.A. Voltage-dependent calcium channels in glial cells. Science 1984, 226, 1345–1347. [Google Scholar] [CrossRef] [PubMed]
  173. Latour, I.; Hamid, J.; Beedle, A.M.; Zamponi, G.W.; MacVicar, B.A. Expression of voltage-gated Ca2+ channel subtypes in cultured astrocytes. Glia 2003, 41, 347–353. [Google Scholar] [CrossRef] [PubMed]
  174. Yaguchi, T.; Nishizaki, T. Extracellular high K+ stimulates vesicular glutamate release from astrocytes by activating voltage-dependent calcium channels. J. Cell. Physiol. 2010, 225, 512–518. [Google Scholar] [CrossRef] [PubMed]
  175. Zhang, Y.; Chen, K.; Sloan, S.A.; Bennett, M.L.; Scholze, A.R.; O’Keeffe, S.; Phatnani, H.P.; Guarnieri, P.; Caneda, C.; Ruderisch, N.; et al. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J. Neurosci. 2014, 34, 11929–11947. [Google Scholar] [CrossRef] [PubMed]
  176. Rungta, R.L.; Bernier, L.; Dissing-olesen, L.; Groten, C.J.; Ledue, J.M.; Ko, R.; Drissler, S.; Macvicar, B.A. Ca2+ transients in astrocyte fine processes occur via Ca2+ influx in the adult mouse hippocampus. Glia 2016, 64, 2093–2103. [Google Scholar] [CrossRef]
Figure 1. Examples of functional roles of astrocyte Ca2+ events. MCEs lead to gliotransmission: (1) ATP/adenosine a. downregulates the excitatory activity by activating presynaptic A1R [60] and b. upregulates inhibitory activity by activating postsynaptic A1R [40]. (2) D-serine enhances LTP via postsynaptic NMDARs [41]. (3) Glutamate released from astrocytes modulates pre- and post-synaptic neuronal glutamate receptors [36,50,56,57,59,61]. (4) In astrocyte endfeet, MCEs cause the production of arachidonic acid (AA) that is metabolized to vasodilative components, such as prostaglandins, and contribute to regulation of cerebral blood flow [12].
Figure 1. Examples of functional roles of astrocyte Ca2+ events. MCEs lead to gliotransmission: (1) ATP/adenosine a. downregulates the excitatory activity by activating presynaptic A1R [60] and b. upregulates inhibitory activity by activating postsynaptic A1R [40]. (2) D-serine enhances LTP via postsynaptic NMDARs [41]. (3) Glutamate released from astrocytes modulates pre- and post-synaptic neuronal glutamate receptors [36,50,56,57,59,61]. (4) In astrocyte endfeet, MCEs cause the production of arachidonic acid (AA) that is metabolized to vasodilative components, such as prostaglandins, and contribute to regulation of cerebral blood flow [12].
Biomolecules 11 01467 g001
Figure 2. Astrocyte Ca2+ pathways activated during synaptic transmission. This diagram highlights the pathways that involve extracellular Ca2+ influx as discussed in this review.
Figure 2. Astrocyte Ca2+ pathways activated during synaptic transmission. This diagram highlights the pathways that involve extracellular Ca2+ influx as discussed in this review.
Biomolecules 11 01467 g002
Figure 3. Functional implications of astrocyte NMDA receptors. The following may occur as a result of Antioxidant protection NMDAR activity, possibly via astrocyte calcium events: (1) Modulation of synaptic activity; ATP gliotransmission is evoked that acts on presynaptic P2XRs and thus downregulates inhibitory activity [130]. (2) Regulation of synaptic strength: reduced astrocyte NMDAR expression decreases the paired-pulse ratio variability [49,139] (3) Protection of neurons against antioxidant stress; NMDAR activation upregulates expression of cdk5/p35 that promotes expression of glutathione precursors through Nrf2 [121]. (4) Regulation of basal astrocyte Ca2+ concentrations, which can define MCEs characteristics such as amplitude and peak frequency [26,27].
Figure 3. Functional implications of astrocyte NMDA receptors. The following may occur as a result of Antioxidant protection NMDAR activity, possibly via astrocyte calcium events: (1) Modulation of synaptic activity; ATP gliotransmission is evoked that acts on presynaptic P2XRs and thus downregulates inhibitory activity [130]. (2) Regulation of synaptic strength: reduced astrocyte NMDAR expression decreases the paired-pulse ratio variability [49,139] (3) Protection of neurons against antioxidant stress; NMDAR activation upregulates expression of cdk5/p35 that promotes expression of glutathione precursors through Nrf2 [121]. (4) Regulation of basal astrocyte Ca2+ concentrations, which can define MCEs characteristics such as amplitude and peak frequency [26,27].
Biomolecules 11 01467 g003
Table 1. Evidence of astrocyte iGluR-mediated Ca2+ activity from Ca2+ imaging in cell culture studies. The concentration of NMDA is noted when over (100 µM) or under (20 µM) the toxic concentration (50 µM). ✓ and ✕ show the presence or absence of function receptors in each study. Agonists: Glutamate (Glu), kainate (KA), quisqualate (QA), Glycine (Gly), N-methyl-D-aspartate (NMDA), α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA).
Table 1. Evidence of astrocyte iGluR-mediated Ca2+ activity from Ca2+ imaging in cell culture studies. The concentration of NMDA is noted when over (100 µM) or under (20 µM) the toxic concentration (50 µM). ✓ and ✕ show the presence or absence of function receptors in each study. Agonists: Glutamate (Glu), kainate (KA), quisqualate (QA), Glycine (Gly), N-methyl-D-aspartate (NMDA), α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA).
Culture PreparationPharmacologyReceptor FunctionalityReference
Rat cortical astrocytes
14–21 days in culture
Agonist: Glu, KA
NMDA (100 µM)
✓Kainate/AMPA receptors
✕ NMDARs
Pearce et al., 1986. [113]
Rat hippocampal astrocytes
1–3 weeks in culture
Agonist: Glu, QA, KA, Gly, NMDA (100 µM)
Blocker: Ca2+-free saline aCSF (EGTA)
✓Kainate/AMPA receptors
✕ NMDARs
Cornell-Bell et al., 1990. [114]
Rat cortical astrocytes
4–9 weeks in culture
Agonist: Glu, KA, QA
NMDA (100 µM)
Blocker: kynurenic acid, Ca2+-free saline (EGTA)
✓Kainate/AMPA receptors
✕ NMDARs
Jensen et al., 1990. [115]
Rat hippocampal astrocytes
2–4 weeks in culture
Agonist: KA, AMPA, Gly, NMDA (100 µM)✕ iGluRs
(no Ca2+ permeable forms)
Cai et al., 1997.
[118]
Rat cerebellar, hippocampal, and cortical astrocytes
10–20-days in culture
Agonist: QA, AMPA
Antagonist: CNQX
✓ AMPARsGlaum et al., 1990. [122]
Rat cortical astrocytes
12–14-days in culture
Agonist: Glu, NMDA (20 µM)
Antagonist: MK801, CNQX
✕ Kainate/AMPA receptors
✓ NMDARs
Jimenez-Blasco et al., 2015. [121]
Rat cerebellar astrocytes
4 weeks in culture
Agonist: Glu/Hypoxia
Antagonist: CNQX
✓ AMPARsKou et al., 2019. [123]
Table 2. Evidence of iGluR-mediated Ca2+ activity from Ca2+ imaging in ex vivo brain slices or acutely isolated astrocytes. Bath or pipette application of drugs are indicated, which affects cell-type specificity. ✓ and ✕ show the presence or absence of function receptors in each study.
Table 2. Evidence of iGluR-mediated Ca2+ activity from Ca2+ imaging in ex vivo brain slices or acutely isolated astrocytes. Bath or pipette application of drugs are indicated, which affects cell-type specificity. ✓ and ✕ show the presence or absence of function receptors in each study.
Astrocyte PreparationiGluR PharmacologyReceptor FunctionalityReference
Hippocampal slices from 10–13-days-old ratsBath-applied✓ iGluRs (type not specified)Porter et al., 1996. [92]
Hippocampal slices from 8-day-old ratsBath-applied✓ NMDARsPasti et al., 1997. [93]
Hippocampal slices of 31–38-days-old ratsBath-applied✓AMPARs
✕ NMDARs
Shelton et al., 1999. [127]
Cortical slice from 1–4-week-old GFAP-EGFP micePatch-applied✓ NMDARsSchipke et al.,
2001. [128]
Hippocampal slice from 10–18-month-old GFAP-EGFP micePatch-applied✓ NMDARsSerrano et al., 2008. [129]
Optic nerve isolated from 15–30-day-old- GFAP-EGFP miceBath-applied✓ AMPARs
✓ NMDARs
Hamilton et al., 2008. [124]
Brain slices and acutely isolated cortical astrocytes from 3-month-old GFAP-EGFP micePatch-applied✓NMDARsPalygin et al., 2010. [109]
Neocortical slice from 1–21-months-old GFAP-EGFP micePatch-applied✓AMPAR
✓NMDAR
Lalo et al., 2011. [132]
Cortical astrocytes isolated from adult GFAP-EGFP micePatch-applied✓ NMDARPalygin et al., 2011. [108]
Cortical astrocytes isolated from adult miceBath-applied✓ NMDARLalo et al., 2014. [130]
Brain slices and acutely isolated cortical astrocytes from 35–59-day-old GFAP-EGFP miceBath-applied✓ NMDARsDzamba et al., 2015. [110]
Olfactory bulb slice from 14–21-day-old Aldh1l1-eGFP miceBath-applied✓ AMPARs
✓ NMDARs
Otsu et al., 2015. [28]
Somatosensory neocortex slice from 21–30-day-old-ratsPatch-applied✓ NMDARsMehina et al., 2017. [131]
Olfactory bulb slice from 8–12-day-old GFAP-EGFP and GLAST-CreERT2-GCaMP6sfl/fl miceBath-applied✓ AMPARsDroste et al., 2017. [134]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmadpour, N.; Kantroo, M.; Stobart, J.L. Extracellular Calcium Influx Pathways in Astrocyte Calcium Microdomain Physiology. Biomolecules 2021, 11, 1467. https://doi.org/10.3390/biom11101467

AMA Style

Ahmadpour N, Kantroo M, Stobart JL. Extracellular Calcium Influx Pathways in Astrocyte Calcium Microdomain Physiology. Biomolecules. 2021; 11(10):1467. https://doi.org/10.3390/biom11101467

Chicago/Turabian Style

Ahmadpour, Noushin, Meher Kantroo, and Jillian L. Stobart. 2021. "Extracellular Calcium Influx Pathways in Astrocyte Calcium Microdomain Physiology" Biomolecules 11, no. 10: 1467. https://doi.org/10.3390/biom11101467

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop