Next Article in Journal
Redefining Autoimmune Disorders’ Pathoetiology: Implications for Mood and Psychotic Disorders’ Association with Neurodegenerative and Classical Autoimmune Disorders
Next Article in Special Issue
Updates on the Physiopathology of Group I Metabotropic Glutamate Receptors (mGluRI)-Dependent Long-Term Depression
Previous Article in Journal
Rickettsia Deregulates Genes Coding for the Neurotoxic Cell Response Pathways in Cerebrocortical Neurons In Vitro
Previous Article in Special Issue
Dysregulated Signaling at Postsynaptic Density: A Systematic Review and Translational Appraisal for the Pathophysiology, Clinics, and Antipsychotics’ Treatment of Schizophrenia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Ryanodine Receptors in Regulating Neuronal Activity and Its Connection to the Development of Alzheimer’s Disease

by
Giuseppe Chiantia
1,†,
Enis Hidisoglu
2,† and
Andrea Marcantoni
2,3,*
1
Department of Neuroscience, University of Turin, 10125 Turin, Italy
2
Department of Drug and Science Technology, University of Torino, Corso Raffaello 30, 10125 Torino, Italy
3
N.I.S. Center, University of Torino, 10125 Turin, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(9), 1236; https://doi.org/10.3390/cells12091236
Submission received: 24 February 2023 / Revised: 19 April 2023 / Accepted: 21 April 2023 / Published: 25 April 2023
(This article belongs to the Special Issue New Advances in Synaptic Dysfunctions and Plasticity)

Abstract

:
Research into the early impacts of Alzheimer’s disease (AD) on synapse function is one of the most promising approaches to finding a treatment. In this context, we have recently demonstrated that the Abeta42 peptide, which builds up in the brain during the processing of the amyloid precursor protein (APP), targets the ryanodine receptors (RyRs) of mouse hippocampal neurons and potentiates calcium (Ca2+) release from the endoplasmic reticulum (ER). The uncontrolled increase in intracellular calcium concentration ([Ca2+]i), leading to the development of Ca2+ dysregulation events and related excitable and synaptic dysfunctions, is a consolidated hallmark of AD onset and possibly other neurodegenerative diseases. Since RyRs contribute to increasing [Ca2+]i and are thought to be a promising target for AD treatment, the goal of this review is to summarize the current level of knowledge regarding the involvement of RyRs in governing neuronal function both in physiological conditions and during the onset of AD.

1. Introduction

Alzheimer’s disease (AD) is characterized by specific hallmarks, including the accumulation of Amyloid Beta (Abeta) oligomers, which are derived from Amyloid Precursor Protein (APP) processing. Their consequent effect on neuronal calcium dysregulation and related synaptic dysfunction has been the subject of numerous investigations in recent years. The uncontrolled increase in [Ca2+]i either results from alterations in several mechanisms of Ca2+ entry through the plasma membrane or Ca2+ release from the endoplasmic reticulum. There is considerable evidence to indicate that Abeta oligomers affect Ca2+ entry through N-methyl-D-aspartate receptors (NMDARs) [1,2,3] and/or voltage gated calcium channels (VGCCs) [4,5,6,7]. The alternative mechanism of [Ca2+]i increase involves the upregulation of ryanodine (RyRs) and inositol triphosphate receptors (IP3Rs) and increasing amounts of evidence suggest that they are impaired during AD onset. These two receptors are not equally distributed along the ER membrane and, while IP3Rs are more concentrated at the somatic region, RyRs prefer the portion of ER membrane that is in the vicinity of the synaptic sites [8], suggesting that they are significantly involved in the control of [Ca2+]i and directly responsible for governing synaptic function. Three different RyR isoforms are known and, while it was previously thought that RyR1 was principally expressed in skeletal muscle, RyR2 in the heart, and RyR3 in neurons [9], it is now widely accepted that all three isoforms are expressed in the brain [10]. Particular importance has been attributed to RyR2 [11] and RyR3 [12] in the hippocampus. RyRs are composed of four identical subunits of approximately 5000 amino acids each forming a square around a central pore and often complexed with other accessory proteins, such as calmodulin, calstabin, phosphodiesterases phosphatases, and kinases that together stabilize RyRs and modulate calcium release [13]. RyRs are the largest known Ca2+ channels and are characterized by high permeability [14]. It has been estimated that the average conductance of RyR1 for calcium is 110 pS and that this can increase to over 700 pS for potassium [14], which is also known to permeate through the pore [15]. Seeing as the ion conductance of most ion channels in the plasma membrane ranges from some single pS to tens of pS [16], the significant contribution of RyRs in governing [Ca2+]i becomes evident. The Ca2+ released from RyRs can be triggered by Ca2+ entry through the Ca2+ channels located in the plasma membrane; a phenomenon called “Ca2+ induced Ca2+ release” (CICR) that is typically observed in cardiac myocytes where RyR2 is the principally expressed isoform. VGCCs [15] and NMDARs [3] also activate RyRs in neurons through CICR. Alternatively, RyRs can be mechanically activated by VGCCs through direct interaction [17]. This latter mechanism is typically observed in skeletal muscle where Ca2+ is released from RyRs upon membrane depolarization [14]. A growing sum of evidence suggests that these receptors play a fundamental role in governing neuronal function. The purpose of this review is to examine and clarify the role of RyRs in regulating neuronal excitability and synaptic properties. Finally, their possible contribution in triggering the onset of neurodegenerative diseases, such as AD, will be discussed and RyR-targeting-based therapeutic strategies will be described.

2. Involvement of RyRs in Synaptic-Function Control

RyRs are clearly involved in long-term synaptic plasticity [18,19,20], but are also engaged in short-term synaptic plasticity with varying degrees of impact, depending on the type of neuron and synapse. Calcium released from RyRs activates the calcium dependent inactivation (CDI) mechanisms of VGCCs in neurons [15,21,22,23]. Although interesting results on thalamocortical neurons [22], have been published and RyRs, but not IP3Rs, have been observed to provide a clear contribution to CDI, no clear data on the contribution of the CDI mechanism, as modulated by calcium released from RyRs, to synaptic function are available. Besides controlling the inactivation mechanisms of presynaptic VGCCs, the calcium released from RyRs can directly contribute to modulating synaptic function. In this regard, significant differences between neurons and synapses can be observed [24,25], together with different mechanisms of neurotransmitter release, represented by synchronous, asynchronous, and spontaneous release [26]. Synaptic release occurs via various vesicle pools, which are all distinguished by different calcium sensors associated with synaptic vesicles [26] and different distances from the calcium source [27]. Calcium sensitivity defines the calcium source responsible for triggering vesicle exocytosis and contributes to further characterizing the physiological properties of the different synapses of central neurons. Spontaneous release is known to be important for neuronal development [28], and for glutamatergic [29] and GABAergic [30] synapse maturation. These quantal events are less reliant, compared to synchronous release, on the input of calcium through voltage-gated calcium channels located at the presynaptic plasma membrane [31], and they can be positively modulated by calcium released from RyRs [32]. Contrasting evidence indicates that, while RyR inhibition in the hippocampus reduces the frequency of mEPSCs (miniature Excitatory Post Synaptic Currents) in CA3 pyramidal neurons both in basal conditions [33] and upon nicotine or caffeine (a RyRs agonist) administration [34], caffeine does not affect the spontaneous release of glutamate in CA1 pyramidal neurons [35]. A different scenario has been depicted when the contribution of calcium released from RyRs in the generation of mIPSCs (miniature Inhibitory Post Synaptic Currents) has been studied, with observation that the administration of nanomolar concentrations of ryanodine in CA3 pyramidal neurons impacted neither the amplitude nor the frequency of mIPSCs [36]. Ryanodine acts as a selective inhibitor of RyRs when administered at micromolar concentrations [18], while at nanomolar concentrations [37], it can trigger RyR opening. In view of the manifold effects on spontaneous release induced by calcium released from RyRs, the presence of specific calcium sensors with various levels of calcium sensitivity on synaptic vesicles has been proposed as a potential explanation [26]. Synchronous release is considered to be mainly triggered by calcium entering through presynaptic VGCCs located at the plasma membrane, as represented by the Cav2.1, 2.2, 2.3 channels [38]. The mechanisms underlying synchronous release entail fast kinetics of calcium-channel activation as well as high local calcium concentrations (up to 100 µM) and close proximity between the calcium source and the calcium sensor [39]. The distance between calcium sensor and calcium source has been estimated to be in the range of 100 nm [26], decreasing up to 10–20 nm in GABAergic synapses [40]. All together, these properties allow the synaptic delay between presynaptic calcium entry and the presynaptic fusion of vesicles containing neurotransmitters within the presynaptic membrane to be reduced. Therefore, the generation of postsynaptic currents that are highly synchronized and characterized by a synaptic delay shorter than 1 ms [40] is guaranteed. RyRs modulate synchronous release and there is growing evidence to indicate that they are distributed at both pre- and postsynaptic sites and that they are involved in the generation of different forms of synaptic plasticity [33,41,42,43,44]. In this regard, a noticeable variation in effects has been observed, with these differences depending on the experimental model and cell type. Though RyRs inhibition significantly affects short-term synaptic plasticity in CA3 pyramidal neurons [33], no effects have been observed in stellate cells [45]. On the other hand, single or sparse APs do not trigger calcium release through RyR activation and, with some exceptions [43], do not significantly contribute to the average amplitude of post synaptic currents [33,45,46]. Similar conclusions were reached in our laboratory, where we observed that RyRs did not contribute significantly to the generation of spontaneous firing (Figure 1a) and single eEPSCs (evoked Excitatory Post Synaptic Currents) amplitude [7] in primary cultured hippocampal neurons (Figure 1e). A possible explanation for this is that the calcium released from RyRs takes part in the glutamate release when neurons are stimulated at a high frequency. Repetitive high frequency stimulation gives rise to increased amounts of asynchronous release, which typically prevails over synchronous release over time. Asynchronous release, along with synchronous release, shares a common pool of readily released vesicles (RRPtot, total Readily Releasable Pool) and is characterized by the presence of calcium sensors with a higher sensitivity to calcium [26]. As a consequence, asynchronous release is not characterized by a strong proximity between calcium source and synaptic vesicles [47], and is triggered by calcium bulk that accumulates at presynaptic sites, especially during prolonged and high-frequency stimulation [48]. Asynchronous release is characterized by a wider range of calcium sources, including RyRs [49]. Both GABAergic [50], and glutamatergic synapses [45,51], are characterized by asynchronous release, whose pivotal role in the control of network excitability [26] and the potentiation of neurotransmitter release [52] has been well-documented. Neuronal networks are characterized by a balance between excitatory and inhibitory synapses, where inhibitory synapses represent 80% of the total synaptic conductance [53,54]. The majority of inhibitory synapses belong to cholecystokinin (CCK) and parvalbumin (PV) positive interneurons, with the former mainly characterized by a higher degree of asynchronous release than the latter [55], and by a calcium-dependent mechanism of vesicle release that is mediated by presynaptic N-type (Cav2.2) calcium channels in CCK neurons and P/Q (Cav2.1) channels in PV neurons. The involvement of calcium released from RyRs in the modulation of the asynchronous release of GABAergic synapses has, so far, only been suggested [27]. In this regard, experiments performed on cerebellar basket cells [43] only revealed that RyR inhibition decreases the synchronous release of GABA, while their involvement in the modulation of asynchronous release is still to be elucidated. Regarding glutamatergic synapses, the precise role that RyRs play in asynchronous release has also been speculated [56], but not fully explained. It has been suggested that the asynchronous (or delayed) release of glutamate is not affected by calcium released from RyRs, at least in a context of presynaptic stimuli characterized by low frequency and short duration [45].

3. Involvement of RyRs in the Control of Neuronal Excitability

RyRs are involved in the control of the firing patterns [57] and frequency inhibition [58] of thalamocortical neurons as they potentiate the amplitude of the afterhyperpolarization current (IAHP) by means of the Ca2+-dependent activation of potassium channels (IKCa), which can be measured at the end of the action potential (AP) waveform. However, other evidence suggests that RyRs are functionally coupled to other potassium channels, such as Kv2.1 [59]. IKCa can be functionally coupled to VGCCs and contribute to shaping the AP and firing patterns. In this regard, small conductance (SK) potassium channels display a high sensitivity to calcium and do not need to be in close proximity to channels, unlike big conductance (BK) potassium channels [60]. This fascinating feature quite closely resembles the different sensitivities to calcium ions shown by synaptic vesicles and once again paves the way for understating calcium’s manifold mechanisms of action in governing neuronal function. IAHP are temporally distinguished in fast (fAHP), medium (mAHP), and slow (sAHP) AHP, while fAHP is attributed to the opening of BK channels, intermediate and small conductance potassium channels (IK, SK) are responsible for the activation of sAHP and mAHP, respectively [12,61,62]. The calcium source required for IKCa activation can include RyRs. It has been shown that the calcium released from RyR3 in hippocampal CA1 pyramidal neurons is responsible for IK activation and governs firing activity [63]. On the other hand, it has been suggested that BK channels are not activated by the calcium released from RyRs in primary cultured hippocampal neurons, and that they do not contribute to modulating the spontaneous firing of the hippocampal network [7] (Figure 1b), although this functional coupling has been observed and assessed in other neurons [64,65]. Finally, there is no proof, at least under physiological circumstances, that RyRs and mAHP have any functional relationship [62].

4. Altered Function of RyR during AD

Moving beyond physiological conditions, increased RyR expression and function have been observed during AD onset, and a relationship between RyR dysfunction and calcium dyshomeostasis has been proposed [66]. Currently available information suggests that the expression of the RyR2 and RyR3 isoforms is increased in hippocampal neurons from APPPS1 mice [67], while the expression of all three RyRs isoforms is increased in a human neuroblastoma cell line that overexpresses Swedish-mutation-harboring APP [68]. By contrast, only the expression of RyR3 is increased in the cortical neurons of TgCRND8 mice [69]. Finally, a functional alteration of RyR2 isoforms has been reported as occurring in the brains of AD patients and in two AD murine models. This alteration leads to a leaky outflow of Ca2+ from the ER and increased [Ca2+]i, which is responsible for Ca2+ dyshomeostasis and synaptic dysfunction [70]. In spite of these findings, a precise description of how AD affects RyR function is still lacking, probably because of the multiple causes that contribute to generating the disease and the use of different experimental models to study the onset of this pathology. In light of this, we draw the conclusion that, even though the available data indicate a potentiation of the mechanism by which Ca2+ is released from RyRs, the presence of a RyRs isoform that is more crucial to initiating this effect is still unknown.

4.1. Consequences of RyR Dysfunction on Synaptic Activity during AD

Alzheimer’s disease has been associated with synaptic disruption [71,72], leading to network hyperactivity [73], followed by decreased neuronal activity [74,75]. Nevertheless, the precise changes induced by AD on synaptic properties remain unclear. The conflicting observations depend on several factors, including models, neuronal type, and also the contrasting effects induced by the amyloid beta peptides administered at different concentrations and at different stages of aggregation. Recent data indicate that the glutamate release probability and amplitude of eEPSCs are reduced in 6/7-month old APP/PS1 mice, without changes in the size of the readily releasable pool (RRP) [76]. Similar results have been obtained by treating hippocampal neurons with 400 nM of Abeta42 oligomers [76]. Accordingly, our recent observations suggest that Abeta42 oligomers administered at concentrations ranging between 200 nM and 1 µM decreased spontaneous network firing via the inhibition of glutamatergic synapses. Therefore, it has been suggested that, when glutamatergic synapses are stimulated at high frequencies (above 50 Hz), Abeta42 oligomers increase the amount of calcium released from RyRs. This increased release is, in turn, responsible for the activation of presynaptic BK channels [7], causing presynaptic membrane hyperpolarization, reducing the quantity of calcium entering through presynaptic VGCCs and leads to synaptic and firing inhibition that can be partly restored by dantrolene and paxilline, which are RyR- and BK-channel inhibitors, respectively (Figure 1c,d). The impact of Abeta42 oligomers on glutamatergic synapses has recently been detailed [3,77], revealing that, although the above mechanism is observed at high neuronal firing frequencies, when considering single eEPSCs that are generated via activation of the NMDA receptor, Abeta42 increases their amplitude by increasing the glutamate release probability, the number of release sites, and the size of the readily releasable pool of vesicles available for synchronous release (RRPsyn) (Figure 2).
Similar results have been observed in a mouse model [78] that is characterized by neuronal hyperexcitability and dependent on an aberrant amount of calcium being released from presynaptic RyRs, which in turn, increases the release probability of glutamatergic synapses. The fact that Abeta42 induces this range of diverse effects may depend not only on the degree of synaptic stimulation in terms of frequencies but also on the progression of the pathology. Therefore, it has been suggested that stimulating effects characterize the early stages of AD, when Abeta amyloid accumulation is lower, while inhibitory effects are more frequently observed during the late stages of the disease [79]. Furthermore, we have recently shown that Abeta42 accumulation at excitatory synapses targets the postsynaptic membrane and reduces the number of NMDA receptors [3] (Figure 2). NMDARs are involved in long-term synaptic plasticity and are responsible for the activation of the CICR mechanism following RyR activation [3]. Their decrease following Abeta42 accumulation may contribute to the impairment of long-term synaptic plasticity and related memory formation during AD onset [80]. The effect of AD onset on GABAergic synapses has also been studied, and an increased number of GABAergic projections towards the CA1 hippocampal region [81] has been observed in APPPS1 mice, together with a potentiation of GABAergic synaptic activity observed in transgenic mice overexpressing hAPP carrying the Swedish and Indiana FAD mutations [82,83]. These results contradict those of other studies, in which AD has been shown to have an inhibitory impact on GABAergic synapses observed in double transgenic APP23xPS45 mice that overexpress both the amyloid precursor protein (APPSwe) and mutant presenilin 1 [84]. We have recently focused on the effects of Abeta42 oligomers on GABAergic synapses, and observed an increase in eIPSC (evoked Inhibitory Post Synaptic Currents) amplitude together with an increased number of release sites and RRPsyn size (dependent on a higher amount of calcium release from RyRs [85]), although the probability of release did not change (Figure 3). These results suggest that the Abeta42 mechanisms of action may be different depending on the synapses involved. We also focused on the effects of Abeta42 on the asynchronous release of GABA and observed that Abeta42 had an inhibitory effect, contradicting the observations made when synchronous release was measured [85]. This result explains our results, which indicated that RRPtot was of comparable size in the controls and the presence of Abeta42 [85] (Figure 3). Therefore, we concluded that the overactivation of RyRs is an example of a recurring feature, although the effects of AD and, in particular, of Abeta42 on synaptic function are multifaced and depend on the type of synapse and stimulation mode (high vs. low frequency stimulation, long vs. short term stimulation).

4.2. Consequences of RyR Dysfunction on Neuron Excitability during AD

General hyperexcitability is often observed during AD, although many exceptions have been described [7,62,86]. Early studies performed in 5xFAD mice focused on IAHP [61], and observed that the medium (mAHP) and slow (sAHP) components were reduced and responsible for increased intrinsic excitability. This latter is known to be modulated by calcium released from RyRs [63], but the mechanism behind the recruitment of IKCa during AD, along with the rise in intrinsic hyperexcitability, is still unclear, and a reduction in firing activity seems to be a more logical consequence of the increased calcium released from RyRs. Another hypothesis suggests that the amount of A-type K+ current is decreased in 5xFAD mice [87], and following the administration of Abeta42 oligomers [88], that hyperexcitability is consequently observed. In this regard, it has been suggested that A-type K+-channel expression depends on the Kv4.2 subunit, which in turn, is found to be inhibited by the calcium released from RyRs [87], explaining why an increase in the amount of calcium released from RyRs, caused by the reduced expression of A-type K+ channels, is responsible for neuronal hyperexcitability.

4.3. Proposed RyR-Inhibition-Based Therapies for AD Treatment

In AD mouse models, it has recently been shown that the accumulation of Abeta oligomers prevents the proper interaction between RyRs and the intracellular stabilizer proteins calstabins, preventing the complete closure of RyRs, which assume a leaky conformational state and generate continuous slow release and accumulation of calcium in the cytoplasm [70]. It has been suggested that this process is associated with the PKA-dependent phosphorylation of RyRs, a pathway that is over activated by amyloid beta peptides via the activation of beta2 adrenoreceptors [89]. Two different isoforms of calstabins are known, and while calstabin1 is preferentially associated with RyR1 and RyR3, calstabin2 has a higher affinity for RyR2 [90]. In our laboratory, we have recently shown that RyR inhibition, via the administration of dantrolene, counteracts the increases in [Ca2+]i that are induced by Abeta42 and improves the excitable and synaptic properties of hippocampal neurons [3,7] (Figure 1c and Figure 2). Dantrolene is actually the only RyRs inhibitor (RyRI) used for the clinical treatment of malignant hyperthermia [91], but its use for AD treatment is not conceivable due to problems related to stability, selectivity, and permeability through the blood–brain barrier (BBB) [92]. New classes of RyRIs are being studied, and among those of particular interest, we can find the class of RyCal compounds (RyCals) [70,85,89], which can promote the binding between calstabins and RyRs. The only commercially available RyCal molecule is S107, but data on its effect on neuronal function are scarce. In this context, we have recently shown that S107 counteracts the effects induced by Abeta42 by reducing eIPSC amplitude to values comparable to those observed under control conditions (Figure 3). Moreover, the unbalance between synchronous and asynchronous release is also recovered by s107 administration [85] (Figure 3). Interesting results have been obtained through the administration of carvedilol [87,93]. This compound is a racemic mixture of s and r carvedilol, and while the S-enantiomer is an adrenoreceptor inhibitor, the R-enantiomer inhibits RyRs activity without any affinity for beta adrenergic receptors [93]. In view of the cardiovascular effects caused by the inhibition of adrenergic receptors, there has been a preference for using R-carvedilol in the treatment of impairments induced by AD onset. Interestingly, a significant rescue of neuronal function following R-carvedilol administration has been observed in 5xFAD mice, with this effect not being detectable in the presence of the racemic mixture [87]. On the other hand, promising results have also been achieved through the inhibition of beta2 adrenoreceptors [89], indicating that perhaps the racemic mixture of carvedilol should be used. In conclusion, there is no clear evidence as to whether the racemic mixture or single enantiomer of carvedilol are preferable for the targeting of RyRs and related synaptic dysfunction. Moreover, the question of whether the inhibition of beta2 adrenoreceptors may represent a promising strategy for AD treatment is still yet to be satisfactorily answered.

5. Conclusions

Studies performed in recent years indicate that RyRs may play a role in the onset of AD and other neurodegenerative diseases [94]. The development of pharmacological therapy is far from being established because different RyR isoforms, of which the exact function is partly unknown due to a lack of selective RyRs-isoform inhibitors, are present in the brain. Additionally, it is necessary to consider their principal roles in other organs, including the heart and arteries. While there is considerable evidence to suggest that the inhibition of RyRs might protect neuronal function during AD, much remains to be done, especially in the understanding of the mechanisms induced by RyR dysfunctions that lead to alterations in synaptic properties and the intrinsic excitability of central neurons during AD onset.

Author Contributions

Conceptualization, G.C., E.H. and A.M.; Methodology, G.C., E.H. and A.M.; Writing—original draft preparation, A.M.; writing—review and editing, G.C., E.H. and A.M.; supervision, A.M.; funding acquisition, A.M. and E.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The University of Turin, local funds, EMBO Short-term fellowship (EMBO-7835) to E.H., F.F.A.B.R. funds (Fondo Finanziamento delle Attività Base di Ricerca) from Italian M.I.U.R. to A.M.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.; Li, P.; Feng, J.; Wu, M. Dysfunction of NMDA receptors in Alzheimer’s disease. Neurol. Sci. Off. J. Ital. Neurol. Soc. Ital. Soc. Clin. Neurophysiol. 2016, 37, 1039–1047. [Google Scholar] [CrossRef] [PubMed]
  2. Snyder, E.M.; Nong, Y.; Almeida, C.G.; Paul, S.; Moran, T.; Choi, E.Y.; Nairn, A.C.; Salter, M.W.; Lombroso, P.J.; Gouras, G.K.; et al. Regulation of NMDA receptor trafficking by amyloid-beta. Nat. Neurosci. 2005, 8, 1051–1058. [Google Scholar] [CrossRef] [PubMed]
  3. Marcantoni, A.; Cerullo, M.S.; Buxeda, P.; Tomagra, G.; Giustetto, M.; Chiantia, G.; Carabelli, V.; Carbone, E. Amyloid Beta42 oligomers up-regulate the excitatory synapses by potentiating presynaptic release while impairing postsynaptic NMDA receptors. J. Physiol. 2020, 598, 2183–2197. [Google Scholar] [CrossRef] [PubMed]
  4. Nimmrich, V.; Grimm, C.; Draguhn, A.; Barghorn, S.; Lehmann, A.; Schoemaker, H.; Hillen, H.; Gross, G.; Ebert, U.; Bruehl, C. Amyloid β oligomers (Aβ1–42 globulomer) suppress spontaneous synaptic activity by inhibition of P/Q-type calcium currents. J. Neurosci. Off. J. Soc. Neurosci. 2008, 28, 788–797. [Google Scholar] [CrossRef]
  5. Wang, Y.; Mattson, M.P. L-type Ca2+ currents at CA1 synapses, but not CA3 or dentate granule neuron synapses, are increased in 3xTgAD mice in an age-dependent manner. Neurobiol. Aging 2013, 35, 88–95. [Google Scholar] [CrossRef] [PubMed]
  6. Thibault, O.; Pancani, T.; Landfield, P.W.; Norris, C.M. Reduction in neuronal L-type calcium channel activity in a double knock-in mouse model of Alzheimer’s disease. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2012, 1822, 546–549. [Google Scholar] [CrossRef]
  7. Gavello, D.; Calorio, C.; Franchino, C.; Cesano, F.; Carabelli, V.; Carbone, E.; Marcantoni, A. Early Alterations of Hippocampal Neuronal Firing Induced by Abeta42. Cereb. Cortex 2018, 28, 433–446. [Google Scholar] [CrossRef]
  8. Chakroborty, S.; Briggs, C.; Miller, M.B.; Goussakov, I.; Schneider, C.; Kim, J.; Wicks, J.; Richardson, J.C.; Conklin, V.; Cameransi, B.G.; et al. Stabilizing ER Ca2+ channel function as an early preventative strategy for Alzheimer’s disease. PLoS ONE 2013, 7, e52056. [Google Scholar] [CrossRef]
  9. Lanner, J.T.; Georgiou, D.K.; Joshi, A.D.; Hamilton, S.L. Ryanodine Receptors: Structure, Expression, Molecular Details, and Function in Calcium Release. Cold Spring Harb. Perspect. Biol. 2010, 2, a003996. [Google Scholar] [CrossRef]
  10. Giannini, G.; Conti, A.; Mammarella, S.; Scrobogna, M.; Sorrentino, V. The ryanodine receptor/calcium channel genes are widely and differentially expressed in murine brain and peripheral tissues. J. Cell Biol. 1995, 128, 893–904. [Google Scholar] [CrossRef]
  11. Hiess, F.; Yao, J.; Song, Z.; Sun, B.; Zhang, Z.; Huang, J.; Chen, L.; Institoris, A.; Estillore, J.P.; Wang, R.; et al. Subcellular localization of hippocampal ryanodine receptor 2 and its role in neuronal excitability and memory. Commun. Biol. 2022, 5, 183. [Google Scholar] [CrossRef] [PubMed]
  12. Tedoldi, A.; Ludwig, P.; Fulgenzi, G.; Takeshima, H.; Pedarzani, P.; Stocker, M. Calcium-induced calcium release and type 3 ryanodine receptors modulate the slow afterhyperpolarising current, sIAHP, and its potentiation in hippocampal pyramidal neurons. PLoS ONE 2020, 15, e0230465. [Google Scholar] [CrossRef] [PubMed]
  13. Bers, D.M. Macromolecular complexes regulating cardiac ryanodine receptor function. J. Mol. Cell. Cardiol. 2004, 37, 417–429. [Google Scholar] [CrossRef]
  14. Zalk, R.; Clarke, O.B.; des Georges, A.; Grassucci, R.A.; Reiken, S.; Mancia, F.; Hendrickson, W.A.; Frank, J.; Marks, A.R. Structure of a mammalian ryanodine receptor. Nature 2015, 517, 44–49. [Google Scholar] [CrossRef]
  15. Verkhratsky, A. Physiology and pathophysiology of the calcium store in the endoplasmic reticulum of neurons. Physiol. Rev. 2004, 85, 201–279. [Google Scholar] [CrossRef] [PubMed]
  16. Hille, B. Ionic channels: Molecular pores of excitable membranes. Harvey Lect. 1986, 82, 47–69. [Google Scholar]
  17. Mouton, J.; Marty, I.; Villaz, M.; Feltz, A.; Maulet, Y. Molecular interaction of dihydropyridine receptors with type-1 ryanodine receptors in rat brain. Biochem. J. 2001, 354, 597–603. [Google Scholar] [CrossRef]
  18. Wang, Y.; Wu, J.; Rowan, M.J.; Anwyl, R. Ryanodine produces a low frequency stimulation-induced NMDA receptor-independent long-term potentiation in the rat dentate gyrus in vitro. J. Physiol. 1996, 495 Pt 3, 755–767. [Google Scholar] [CrossRef]
  19. Arias-Cavieres, A.; Barrientos, G.C.; Sánchez, G.; Elgueta, C.; Muñoz, P.; Hidalgo, C. Ryanodine Receptor-Mediated Calcium Release Has a Key Role in Hippocampal LTD Induction. Front. Cell. Neurosci. 2018, 12, 403. [Google Scholar] [CrossRef]
  20. Rose, C.R.; Konnerth, A. Stores not just for storage. intracellular calcium release and synaptic plasticity. Neuron 2001, 31, 519–522. [Google Scholar] [CrossRef]
  21. Budde, T.; Meuth, S.; Pape, H.C. Calcium-dependent inactivation of neuronal calcium channels. Nat. Rev. Neurosci. 2002, 3, 873–883. [Google Scholar] [CrossRef]
  22. Rankovic, V.; Ehling, P.; Coulon, P.; Landgraf, P.; Kreutz, M.R.; Munsch, T.; Budde, T. Intracellular Ca2+ release-dependent inactivation of Ca2+ currents in thalamocortical relay neurons. Eur. J. Neurosci. 2010, 31, 439–449. [Google Scholar] [CrossRef]
  23. Schwartz, A.D.; Whitacre, C.L.; Wilson, D.F. Do ryanodine receptors regulate transmitter release at the neuromuscular junction of rat? Neurosci. Lett. 1999, 274, 163–166. [Google Scholar] [CrossRef]
  24. Chanaday, N.L.; Kavalali, E.T. Role of the endoplasmic reticulum in synaptic transmission. Curr. Opin. Neurobiol. 2022, 73, 102538. [Google Scholar] [CrossRef]
  25. Courtney, N.A.; Briguglio, J.S.; Bradberry, M.M.; Greer, C.; Chapman, E.R. Excitatory and Inhibitory Neurons Utilize Different Ca2+ Sensors and Sources to Regulate Spontaneous Release. Neuron 2018, 98, 977–991.e975. [Google Scholar] [CrossRef]
  26. Kaeser, P.S.; Regehr, W.G. Molecular mechanisms for synchronous, asynchronous, and spontaneous neurotransmitter release. Annu. Rev. Physiol. 2014, 76, 333–363. [Google Scholar] [CrossRef]
  27. Valiullina-Rakhmatullina, F.F.; Bolshakov, A.P.; Rozov, A.V. Three Modalities of Synaptic Neurotransmitter Release: Rapid Synchronized, Multivesicular, and Asynchronous. Similarities and Differences in Mechanisms. Neurosci. Behav. Physiol. 2020, 50, 102–108. [Google Scholar] [CrossRef]
  28. Andreae, L.C.; Fredj, N.B.; Burrone, J. Independent vesicle pools underlie different modes of release during neuronal development. J. Neurosci. Off. J. Soc. Neurosci. 2012, 32, 1867–1874. [Google Scholar] [CrossRef]
  29. McKinney, R.A.; Capogna, M.; Dürr, R.; Gähwiler, B.H.; Thompson, S.M. Miniature synaptic events maintain dendritic spines via AMPA receptor activation. Nat. Neurosci. 1999, 2, 44–49. [Google Scholar] [CrossRef]
  30. Ganguly, K.; Schinder, A.F.; Wong, S.T.; Poo, M. GABA itself promotes the developmental switch of neuronal GABAergic responses from excitation to inhibition. Cell 2001, 105, 521–532. [Google Scholar] [CrossRef]
  31. Baldelli, P.; Novara, M.; Carabelli, V.; Hernandez-Guijo, J.M.; Carbone, E. BDNF up-regulates evoked GABAergic transmission in developing hippocampus by potentiating presynaptic N- and P/Q-type Ca2+ channels signalling. Eur. J. Neurosci. 2002, 16, 2297–2310. [Google Scholar] [CrossRef] [PubMed]
  32. Llano, I.; González, J.; Caputo, C.; Lai, F.A.; Blayney, L.M.; Tan, Y.P.; Marty, A. Presynaptic calcium stores underlie large-amplitude miniature IPSCs and spontaneous calcium transients. Nat. Neurosci. 2000, 3, 1256–1265. [Google Scholar] [CrossRef] [PubMed]
  33. Emptage, N.J.; Reid, C.A.; Fine, A. Calcium stores in hippocampal synaptic boutons mediate short-term plasticity, store-operated Ca2+ entry, and spontaneous transmitter release. Neuron 2001, 29, 197–208. [Google Scholar] [CrossRef] [PubMed]
  34. Sharma, G.; Vijayaraghavan, S. Modulation of presynaptic store calcium induces release of glutamate and postsynaptic firing. Neuron 2003, 38, 929–939. [Google Scholar] [CrossRef] [PubMed]
  35. Chakroborty, S.; Goussakov, I.; Miller, M.B.; Stutzmann, G.E. Deviant ryanodine receptor-mediated calcium release resets synaptic homeostasis in presymptomatic 3xTg-AD mice. J. Neurosci. Off. J. Soc. Neurosci. 2009, 29, 9458–9470. [Google Scholar] [CrossRef]
  36. Savić, N.; Sciancalepore, M. Intracellular calcium stores modulate miniature GABA-mediated synaptic currents in neonatal rat hippocampal neurons. Eur. J. Neurosci. 1998, 10, 3379–3386. [Google Scholar] [CrossRef]
  37. Adasme, T.; Paula-Lima, A.; Hidalgo, C. Inhibitory ryanodine prevents ryanodine receptor-mediated Ca2+ release without affecting endoplasmic reticulum Ca2+ content in primary hippocampal neurons. Biochem. Biophys. Res. Commun. 2015, 458, 57–62. [Google Scholar] [CrossRef]
  38. Catterall, W.A. Structure and regulation of voltage-gated Ca2+ channels. Annu. Rev. Cell Dev. Biol. 2000, 16, 521–555. [Google Scholar] [CrossRef]
  39. Wen, H.; Hubbard, J.M.; Rakela, B.; Linhoff, M.W.; Mandel, G.; Brehm, P. Synchronous and asynchronous modes of synaptic transmission utilize different calcium sources. Elife 2013, 2, e01206. [Google Scholar] [CrossRef]
  40. Bucurenciu, I.; Kulik, A.; Schwaller, B.; Frotscher, M.; Jonas, P. Nanodomain coupling between Ca2+ channels and Ca2+ sensors promotes fast and efficient transmitter release at a cortical GABAergic synapse. Neuron 2008, 57, 536–545. [Google Scholar] [CrossRef]
  41. Padamsey, Z.; Foster, W.J.; Emptage, N.J. Intracellular Ca2+ Release and Synaptic Plasticity: A Tale of Many Stores. Neurosci. A Rev. J. Bringing Neurobiol. Neurol. Psychiatry 2019, 25, 208–226. [Google Scholar] [CrossRef]
  42. Johenning, F.W.; Theis, A.K.; Pannasch, U.; Rückl, M.; Rüdiger, S.; Schmitz, D. Ryanodine Receptor Activation Induces Long-Term Plasticity of Spine Calcium Dynamics. PLoS Biol. 2015, 13, e1002181. [Google Scholar] [CrossRef]
  43. Galante, M.; Marty, A. Presynaptic ryanodine-sensitive calcium stores contribute to evoked neurotransmitter release at the basket cell-Purkinje cell synapse. J. Neurosci. Off. J. Soc. Neurosci. 2003, 23, 11229–11234. [Google Scholar] [CrossRef]
  44. Narita, K.; Akita, T.; Hachisuka, J.; Huang, S.; Ochi, K.; Kuba, K. Functional coupling of Ca2+ channels to ryanodine receptors at presynaptic terminals. Amplification of exocytosis and plasticity. J. Gen. Physiol. 2000, 115, 519–532. [Google Scholar] [CrossRef]
  45. Carter, A.G.; Vogt, K.E.; Foster, K.A.; Regehr, W.G. Assessing the role of calcium-induced calcium release in short-term presynaptic plasticity at excitatory central synapses. J. Neurosci. Off. J. Soc. Neurosci. 2002, 22, 21–28. [Google Scholar] [CrossRef]
  46. Smith, A.B.; Cunnane, T.C. Ryanodine-sensitive calcium stores involved in neurotransmitter release from sympathetic nerve terminals of the guinea-pig. J. Physiol. 1996, 497 Pt 3, 657–664. [Google Scholar] [CrossRef]
  47. Jiang, M.; Zhu, J.; Liu, Y.; Yang, M.; Tian, C.; Jiang, S.; Wang, Y.; Guo, H.; Wang, K.; Shu, Y. Enhancement of asynchronous release from fast-spiking interneuron in human and rat epileptic neocortex. PLoS Biol. 2012, 10, e1001324. [Google Scholar] [CrossRef]
  48. Lu, T.; Trussell, L.O. Inhibitory transmission mediated by asynchronous transmitter release. Neuron 2000, 26, 683–694. [Google Scholar] [CrossRef]
  49. Narita, K.; Akita, T.; Osanai, M.; Shirasaki, T.; Kijima, H.; Kuba, K. A Ca2+-induced Ca2+ release mechanism involved in asynchronous exocytosis at frog motor nerve terminals. J. Gen. Physiol. 1998, 112, 593–609. [Google Scholar] [CrossRef]
  50. Medrihan, L.; Cesca, F.; Raimondi, A.; Lignani, G.; Baldelli, P.; Benfenati, F. Synapsin II desynchronizes neurotransmitter release at inhibitory synapses by interacting with presynaptic calcium channels. Nat. Commun. 2013, 4, 1512. [Google Scholar] [CrossRef]
  51. Iremonger, K.J.; Bains, J.S. Integration of asynchronously released quanta prolongs the postsynaptic spike window. J. Neurosci. Off. J. Soc. Neurosci. 2007, 27, 6684–6691. [Google Scholar] [CrossRef]
  52. Atluri, P.P.; Regehr, W.G. Delayed release of neurotransmitter from cerebellar granule cells. J. Neurosci. Off. J. Soc. Neurosci. 1998, 18, 8214–8227. [Google Scholar] [CrossRef] [PubMed]
  53. den Boon, F.S.; Werkman, T.R.; Schaafsma-Zhao, Q.; Houthuijs, K.; Vitalis, T.; Kruse, C.G.; Wadman, W.J.; Chameau, P. Activation of type-1 cannabinoid receptor shifts the balance between excitation and inhibition towards excitation in layer II/III pyramidal neurons of the rat prelimbic cortex. Pflug. Arch. Eur. J. Physiol. 2014, 467, 1551–1564. [Google Scholar] [CrossRef]
  54. Valero, M.; Navas-Olive, A.; de la Prida, L.M.; Buzsáki, G. Inhibitory conductance controls place field dynamics in the hippocampus. Cell Rep. 2022, 40, 111232. [Google Scholar] [CrossRef] [PubMed]
  55. Hefft, S.; Jonas, P. Asynchronous GABA release generates long-lasting inhibition at a hippocampal interneuron-principal neuron synapse. Nat. Neurosci. 2005, 8, 1319–1328. [Google Scholar] [CrossRef] [PubMed]
  56. Rozov, A.; Bolshakov, A.P.; Valiullina-Rakhmatullina, F. The Ever-Growing Puzzle of Asynchronous Release. Front. Cell. Neurosci. 2019, 13, 28. [Google Scholar] [CrossRef]
  57. Budde, T.; Sieg, F.; Braunewell, K.H.; Gundelfinger, E.D.; Pape, H.C. Ca2+-induced Ca2+ release supports the relay mode of activity in thalamocortical cells. Neuron 2000, 26, 483–492. [Google Scholar] [CrossRef]
  58. Cheong, E.; Kim, C.; Choi, B.J.; Sun, M.; Shin, H.S. Thalamic ryanodine receptors are involved in controlling the tonic firing of thalamocortical neurons and inflammatory pain signal processing. J. Neurosci. Off. J. Soc. Neurosci. 2011, 31, 1213–1218. [Google Scholar] [CrossRef]
  59. Mandikian, D.; Bocksteins, E.; Parajuli, L.K.; Bishop, H.I.; Cerda, O.; Shigemoto, R.; Trimmer, J.S. Cell type-specific spatial and functional coupling between mammalian brain Kv2.1 K+ channels and ryanodine receptors. J. Comp. Neurol. 2014, 522, 3555–3574. [Google Scholar] [CrossRef] [PubMed]
  60. Fakler, B.; Adelman, J.P. Control of KCa channels by calcium nano/microdomains. Neuron 2008, 59, 873–881. [Google Scholar] [CrossRef]
  61. Sahu, G.; Turner, R.W. The Molecular Basis for the Calcium-Dependent Slow Afterhyperpolarization in CA1 Hippocampal Pyramidal Neurons. Front. Physiol. 2021, 12, 759707. [Google Scholar] [CrossRef] [PubMed]
  62. Chakroborty, S.; Kim, J.; Schneider, C.; Jacobson, C.; Molgo, J.; Stutzmann, G.E. Early presynaptic and postsynaptic calcium signaling abnormalities mask underlying synaptic depression in presymptomatic Alzheimer’s disease mice. J. Neurosci. Off. J. Soc. Neurosci. 2012, 32, 8341–8353. [Google Scholar] [CrossRef]
  63. van de Vrede, Y.; Fossier, P.; Baux, G.; Joels, M.; Chameau, P. Control of IsAHP in mouse hippocampus CA1 pyramidal neurons by RyR3-mediated calcium-induced calcium release. Pflug. Arch. Eur. J. Physiol. 2007, 455, 297–308. [Google Scholar] [CrossRef] [PubMed]
  64. Chavis, P.; Ango, F.; Michel, J.M.; Bockaert, J.; Fagni, L. Modulation of big K+ channel activity by ryanodine receptors and L-type Ca2+ channels in neurons. Eur. J. Neurosci. 1998, 10, 2322–2327. [Google Scholar] [CrossRef]
  65. Beurg, M.; Hafidi, A.; Skinner, L.J.; Ruel, J.; Nouvian, R.; Henaff, M.; Puel, J.L.; Aran, J.M.; Dulon, D. Ryanodine receptors and BK channels act as a presynaptic depressor of neurotransmission in cochlear inner hair cells. Eur. J. Neurosci. 2005, 22, 1109–1119. [Google Scholar] [CrossRef] [PubMed]
  66. Yu, J.T.; Chang, R.C.; Tan, L. Calcium dysregulation in Alzheimer’s disease: From mechanisms to therapeutic opportunities. Prog. Neurobiol. 2009, 89, 240–255. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, J.; Supnet, C.; Sun, S.; Zhang, H.; Good, L.; Popugaeva, E.; Bezprozvanny, I. The role of ryanodine receptor type 3 in a mouse model of Alzheimer disease. Channels 2014, 8, 230–242. [Google Scholar] [CrossRef]
  68. Oules, B.; Del Prete, D.; Greco, B.; Zhang, X.; Lauritzen, I.; Sevalle, J.; Moreno, S.; Paterlini-Brechot, P.; Trebak, M.; Checler, F.; et al. Ryanodine receptor blockade reduces amyloid-beta load and memory impairments in Tg2576 mouse model of Alzheimer disease. J. Neurosci. Off. J. Soc. Neurosci. 2012, 32, 11820–11834. [Google Scholar] [CrossRef]
  69. Supnet, C.; Grant, J.; Kong, H.; Westaway, D.; Mayne, M. Amyloid-β-(1-42) increases ryanodine receptor-3 expression and function in neurons of TgCRND8 mice. J. Biol. Chem. 2006, 281, 38440–38447. [Google Scholar] [CrossRef]
  70. Lacampagne, A.; Liu, X.; Reiken, S.; Bussiere, R.; Meli, A.C.; Lauritzen, I.; Teich, A.F.; Zalk, R.; Saint, N.; Arancio, O.; et al. Post-translational remodeling of ryanodine receptor induces calcium leak leading to Alzheimer’s disease-like pathologies and cognitive deficits. Acta Neuropathol. 2017, 134, 749–767. [Google Scholar] [CrossRef]
  71. Walsh, D.M.; Selkoe, D.J. Amyloid β-protein and beyond: The path forward in Alzheimer’s disease. Curr. Opin. Neurobiol. 2020, 61, 116–124. [Google Scholar] [CrossRef] [PubMed]
  72. Selkoe, D.J. Alzheimer’s disease is a synaptic failure. Science 2002, 298, 789–791. [Google Scholar] [CrossRef] [PubMed]
  73. Mucke, L.; Selkoe, D.J. Neurotoxicity of amyloid beta-protein: Synaptic and network dysfunction. Cold Spring Harb. Perspect. Med. 2012, 2, a006338. [Google Scholar] [CrossRef] [PubMed]
  74. Stargardt, A.; Swaab, D.F.; Bossers, K. Storm before the quiet: Neuronal hyperactivity and Aβ in the presymptomatic stages of Alzheimer’s disease. Neurobiol. Aging 2015, 36, 1–11. [Google Scholar] [CrossRef] [PubMed]
  75. Machulda, M.M.; Ward, H.A.; Borowski, B.; Gunter, J.L.; Cha, R.H.; O’Brien, P.C.; Petersen, R.C.; Boeve, B.F.; Knopman, D.; Tang-Wai, D.F.; et al. Comparison of memory fMRI response among normal, MCI, and Alzheimer’s patients. Neurology 2003, 61, 500–506. [Google Scholar] [CrossRef] [PubMed]
  76. He, Y.; Wei, M.; Wu, Y.; Qin, H.; Li, W.; Ma, X.; Cheng, J.; Ren, J.; Shen, Y.; Chen, Z.; et al. Amyloid β oligomers suppress excitatory transmitter release via presynaptic depletion of phosphatidylinositol-4,5-bisphosphate. Nat. Commun. 2019, 10, 1193. [Google Scholar] [CrossRef] [PubMed]
  77. Finch, M.S.; Bagit, A.; Marko, D.M. Amyloid beta 42 oligomers induce neuronal and synaptic receptor dysfunctions. J. Physiol. 2020, 598, 3545–3546. [Google Scholar] [CrossRef]
  78. Lerdkrai, C.; Asavapanumas, N.; Brawek, B.; Kovalchuk, Y.; Mojtahedi, N.; Olmedillas Del Moral, M.; Garaschuk, O. Intracellular Ca2+ stores control in vivo neuronal hyperactivity in a mouse model of Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2018, 115, E1279–E1288. [Google Scholar] [CrossRef]
  79. Puzzo, D.; Privitera, L.; Leznik, E.; Fa, M.; Staniszewski, A.; Palmeri, A.; Arancio, O. Picomolar amyloid-beta positively modulates synaptic plasticity and memory in hippocampus. J. Neurosci. Off. J. Soc. Neurosci. 2009, 28, 14537–14545. [Google Scholar] [CrossRef]
  80. Jacobsen, J.S.; Wu, C.C.; Redwine, J.M.; Comery, T.A.; Arias, R.; Bowlby, M.; Martone, R.; Morrison, J.H.; Pangalos, M.N.; Reinhart, P.H.; et al. Early-onset behavioral and synaptic deficits in a mouse model of Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2006, 103, 5161–5166. [Google Scholar] [CrossRef]
  81. Hollnagel, J.O.; Elzoheiry, S.; Gorgas, K.; Kins, S.; Beretta, C.A.; Kirsch, J.; Kuhse, J.; Kann, O.; Kiss, E. Early alterations in hippocampal perisomatic GABAergic synapses and network oscillations in a mouse model of Alzheimer’s disease amyloidosis. PLoS ONE 2019, 14, e0209228. [Google Scholar] [CrossRef] [PubMed]
  82. Hidisoglu, E.; Kantar-Gok, D.; Er, H.; Acun, A.D.; Yargicoglu, P. Alterations in spontaneous delta and gamma activity might provide clues to detect changes induced by amyloid-β administration. Eur. J. Neurosci. 2018, 47, 1013–1023. [Google Scholar] [CrossRef] [PubMed]
  83. Palop, J.J.; Chin, J.; Roberson, E.D.; Wang, J.; Thwin, M.T.; Bien-Ly, N.; Yoo, J.; Ho, K.O.; Yu, G.Q.; Kreitzer, A.; et al. Aberrant excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal circuits in mouse models of Alzheimer’s disease. Neuron 2007, 55, 697–711. [Google Scholar] [CrossRef] [PubMed]
  84. Busche, M.A.; Eichhoff, G.; Adelsberger, H.; Abramowski, D.; Wiederhold, K.H.; Haass, C.; Staufenbiel, M.; Konnerth, A.; Garaschuk, O. Clusters of hyperactive neurons near amyloid plaques in a mouse model of Alzheimer’s disease. Science 2008, 321, 1686–1689. [Google Scholar] [CrossRef] [PubMed]
  85. Hidisoglu, E.; Chiantia, G.; Franchino, C.; Tomagra, G.; Giustetto, M.; Carbone, E.; Carabelli, V.; Marcantoni, A. The RyR-calstabin interaction stabilizer S107 protects hippocampal neurons from GABAergic synaptic alterations induced by Abeta42 oligomers. J. Physiol. 2022, 600, 5295–5309. [Google Scholar] [CrossRef]
  86. Marcantoni, A.; Raymond, E.F.; Carbone, E.; Marie, H. Firing properties of entorhinal cortex neurons and early alterations in an Alzheimer’s disease transgenic model. Pflug. Arch. Eur. J. Physiol. 2014, 466, 1437–1450. [Google Scholar] [CrossRef] [PubMed]
  87. Yao, J.; Sun, B.; Institoris, A.; Zhan, X.; Guo, W.; Song, Z.; Liu, Y.; Hiess, F.; Boyce, A.K.J.; Ni, M.; et al. Limiting RyR2 Open Time Prevents Alzheimer’s Disease-Related Neuronal Hyperactivity and Memory Loss but Not β-Amyloid Accumulation. Cell. Rep. 2020, 32, 108169. [Google Scholar] [CrossRef]
  88. Scala, F.; Fusco, S.; Ripoli, C.; Piacentini, R.; Li Puma, D.D.; Spinelli, M.; Laezza, F.; Grassi, C.; D’Ascenzo, M. Intraneuronal Aβ accumulation induces hippocampal neuron hyperexcitability through A-type K+ current inhibition mediated by activation of caspases and GSK-3. Neurobiol. Aging 2015, 36, 886–900. [Google Scholar] [CrossRef]
  89. Bussiere, R.; Lacampagne, A.; Reiken, S.; Liu, X.; Scheuerman, V.; Zalk, R.; Martin, C.; Checler, F.; Marks, A.R.; Chami, M. Amyloid β production is regulated by β2-adrenergic signaling-mediated post-translational modifications of the ryanodine receptor. J. Biol. Chem. 2017, 292, 10153–10168. [Google Scholar] [CrossRef]
  90. Wehrens, X.H.; Lehnart, S.E.; Marks, A.R. Intracellular calcium release and cardiac disease. Annu. Rev. Physiol. 2005, 67, 69–98. [Google Scholar] [CrossRef]
  91. Liang, L.; Wei, H. Dantrolene, a treatment for Alzheimer disease? Alzheimer Dis. Assoc. Disord. 2015, 29, 1–5. [Google Scholar] [CrossRef] [PubMed]
  92. Muehlschlegel, S.; Sims, J.R. Dantrolene: Mechanisms of neuroprotection and possible clinical applications in the neurointensive care unit. Neurocrit. Care 2009, 10, 103–115. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, J.; Zhou, Q.; Smith, C.D.; Chen, H.; Tan, Z.; Chen, B.; Nani, A.; Wu, G.; Song, L.S.; Fill, M.; et al. Non-β-blocking R-carvedilol enantiomer suppresses Ca2+ waves and stress-induced ventricular tachyarrhythmia without lowering heart rate or blood pressure. Biochem. J. 2015, 470, 233–242. [Google Scholar] [CrossRef] [PubMed]
  94. Huang, L.; Xue, Y.; Feng, D.; Yang, R.; Nie, T.; Zhu, G.; Tao, K.; Gao, G.; Yang, Q. Blockade of RyRs in the ER Attenuates 6-OHDA-Induced Calcium Overload, Cellular Hypo-Excitability and Apoptosis in Dopaminergic Neurons. Front. Cell. Neurosci. 2017, 11, 52. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Abeta42 inhibits hippocampal network excitability by targeting RyRs. (a)Spontaneous electrical activity of primary cultured hippocampal network, as recorded by MEA (one representative trace), showing that, in the absence of Abeta42 (ctr), RyR inhibition through dantrolene does not affect the spontaneous frequency of extracellularly recorded action potentials. (b) Similar results have been obtained by blocking BK (big conductance potassium) channels using paxilline, suggesting that, under control conditions, neither the RyR nor BK channels are involved in the modulation of hippocampal network excitability. (c,d) When neurons are incubated with Abeta42 oligomers, it is observed that decreased firing activity is partly rescued by dantrolene or paxilline administration. (e) eEPSC amplitude is increased following paxilline or dantrolene administration in hippocampal neurons previously treated with Abeta42, while no effects were observed in the absence of Abeta42 (untreated neurons) adapted from [7].
Figure 1. Abeta42 inhibits hippocampal network excitability by targeting RyRs. (a)Spontaneous electrical activity of primary cultured hippocampal network, as recorded by MEA (one representative trace), showing that, in the absence of Abeta42 (ctr), RyR inhibition through dantrolene does not affect the spontaneous frequency of extracellularly recorded action potentials. (b) Similar results have been obtained by blocking BK (big conductance potassium) channels using paxilline, suggesting that, under control conditions, neither the RyR nor BK channels are involved in the modulation of hippocampal network excitability. (c,d) When neurons are incubated with Abeta42 oligomers, it is observed that decreased firing activity is partly rescued by dantrolene or paxilline administration. (e) eEPSC amplitude is increased following paxilline or dantrolene administration in hippocampal neurons previously treated with Abeta42, while no effects were observed in the absence of Abeta42 (untreated neurons) adapted from [7].
Cells 12 01236 g001
Figure 2. Glutamatergic eEPSCs that are dependent on the activation of the NMDA receptor are potentiated by Abeta42. (a,b) Abeta42 increases the amount of calcium released from RyRs by potentiating the release probability associated with glutamate release, increasing the size of the readily releasable pool associated with synchronous release (RRPsyn), and the number of release sites, but postsynaptically decreases the number of synaptic NMDA receptors. (c) Dantrolene, an allosteric inhibitor of RyRs, restores the synaptic properties observed under control conditions (a).
Figure 2. Glutamatergic eEPSCs that are dependent on the activation of the NMDA receptor are potentiated by Abeta42. (a,b) Abeta42 increases the amount of calcium released from RyRs by potentiating the release probability associated with glutamate release, increasing the size of the readily releasable pool associated with synchronous release (RRPsyn), and the number of release sites, but postsynaptically decreases the number of synaptic NMDA receptors. (c) Dantrolene, an allosteric inhibitor of RyRs, restores the synaptic properties observed under control conditions (a).
Cells 12 01236 g002
Figure 3. GABAergic eIPSCs are potentiated by Abeta42. (a,b) Abeta42 increases the amount of calcium released from RyRs and the number of release sites. The total readily releasable pool (RRPtot) size associated with synchronous and asynchronous release (RRPsyn and RRPasyn) and the release probability associated with GABA release are not affected by Abeta42, while increased RRPsyn size is observed together with decreased RRPasyn size. At the postsynaptic site, Abeta42 increases the number of GABAA receptors. (c) The administration of the RyRs-calstabin interaction stabilizer S107 restores the synaptic properties observed under control conditions (a).
Figure 3. GABAergic eIPSCs are potentiated by Abeta42. (a,b) Abeta42 increases the amount of calcium released from RyRs and the number of release sites. The total readily releasable pool (RRPtot) size associated with synchronous and asynchronous release (RRPsyn and RRPasyn) and the release probability associated with GABA release are not affected by Abeta42, while increased RRPsyn size is observed together with decreased RRPasyn size. At the postsynaptic site, Abeta42 increases the number of GABAA receptors. (c) The administration of the RyRs-calstabin interaction stabilizer S107 restores the synaptic properties observed under control conditions (a).
Cells 12 01236 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chiantia, G.; Hidisoglu, E.; Marcantoni, A. The Role of Ryanodine Receptors in Regulating Neuronal Activity and Its Connection to the Development of Alzheimer’s Disease. Cells 2023, 12, 1236. https://doi.org/10.3390/cells12091236

AMA Style

Chiantia G, Hidisoglu E, Marcantoni A. The Role of Ryanodine Receptors in Regulating Neuronal Activity and Its Connection to the Development of Alzheimer’s Disease. Cells. 2023; 12(9):1236. https://doi.org/10.3390/cells12091236

Chicago/Turabian Style

Chiantia, Giuseppe, Enis Hidisoglu, and Andrea Marcantoni. 2023. "The Role of Ryanodine Receptors in Regulating Neuronal Activity and Its Connection to the Development of Alzheimer’s Disease" Cells 12, no. 9: 1236. https://doi.org/10.3390/cells12091236

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