Next Article in Journal
Characterization of MicroRNAs Associated with Reproduction in the Brown Planthopper, Nilaparvata lugens
Previous Article in Journal
Ceramide and Sphingosine-1-Phosphate in Neurodegenerative Disorders and Their Potential Involvement in Therapy
Previous Article in Special Issue
Group I Metabotropic Glutamate Receptors and Interacting Partners: An Update
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Opinion

Role of Group I Metabotropic Glutamate Receptors in Spike Timing-Dependent Plasticity

by
Irene Martínez-Gallego
,
Antonio Rodríguez-Moreno
* and
Yuniesky Andrade-Talavera
*
Laboratory of Cellular Neuroscience and Plasticity, Department of Physiology, Anatomy and Cell Biology, Universidad Pablo de Olavide, ES-41013 Seville, Spain
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(14), 7807; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23147807
Submission received: 22 June 2022 / Revised: 11 July 2022 / Accepted: 14 July 2022 / Published: 15 July 2022

Abstract

:
Metabotropic glutamate receptors (mGluRs) are G-protein-coupled receptors that exhibit enormous diversity in their expression patterns, sequence homology, pharmacology, biophysical properties and signaling pathways in the brain. In general, mGluRs modulate different traits of neuronal physiology, including excitability and plasticity processes. Particularly, group I mGluRs located at the pre- or postsynaptic compartments are involved in spike timing-dependent plasticity (STDP) at hippocampal and neocortical synapses. Their roles of participating in the underlying mechanisms for detection of activity coincidence in STDP induction are debated, and diverse findings support models involving mGluRs in STDP forms in which NMDARs do not operate as classical postsynaptic coincidence detectors. Here, we briefly review the involvement of group I mGluRs in STDP and their possible role as coincidence detectors.

1. Introduction

Glutamate is the major excitatory neurotransmitter of the central nervous system, and its actions are mediated by the activation of a diverse family of receptors that can be divided into two large sets comprising ionotropic glutamate receptors (the α-amino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA)-, N-methyl-D-aspartate (NMDA)- and kainate (KA)-type receptors) and metabotropic glutamate receptors (mGluRs). Although the fast glutamate excitatory synaptic transmission is typically mediated by the ionotropic ligand-gated glutamate receptors, the slower and long-lasting effects of glutamate are generally mediated by mGluRs [1,2]. mGluRs are G-protein-coupled receptors (GPCRs) that share common topology and exhibit enormous diversity in their expression patterns, sequence homology, pharmacology, biophysical properties and signaling pathways among the different receptor’s subtypes [2,3].
Based on these properties, mGluRs can be divided into three groups: group I, II and III (mGluR I, mGluR II and mGluR III, respectively). mGluR I includes mGluR1 and mGluR5 receptors that are positively coupled to phospholipase C (PLC), whereas mGluR II (comprising mGluR2 and mGluR3) and mGluR III (including mGluR4 and mGluR6-8) are negatively coupled to the formation of adenylate cyclase-mediated cAMP [4]. The metabotropic nature of mGluR signaling was first discovered by the demonstration that glutamate could stimulate the formation of inositol trisphosphate (IP3) production, thus showing that glutamate, similar to other neurotransmitters such as acetylcholine, could also trigger intracellular pathways by activating G protein-coupled receptors [5].
mGluRs modulate different aspects of neuronal physiology, particularly, excitability and plasticity [1,4,6]. They are widely expressed throughout the brain and are found at both pre- and postsynaptic sites of excitatory glutamatergic synapses. Their location makes mGluRs ideally suited to modulate diverse processes and mechanisms of synaptic transmission and plasticity (e.g., glutamate release and intracellular pathways underlying postsynaptic forms of plasticity). Synaptic plasticity has been widely accepted as a possible functional substrate for memory encoding, information processing and neuronal circuit refinement during development [7]. In a classical view, repetitive activation of synapses drives synaptic changes that entail long-term potentiation (LTP) or long-term depression (LTD) of the synaptic transmission. These forms of long-term changes of synaptic strength studied in ex vivo and in vivo preparations depend on the pattern of stimulation used [8,9,10,11,12,13]. For instance, high-frequency stimulation-induced LTP (HFS-LTP) has been intensively studied and it is known to be NMDAR-dependent [5,14].
Under certain conditions, mGluRs can serve as co-triggers for the induction of NMDAR-dependent LTP [15,16], possibly by facilitating the activation of NMDARs [6,17]. mGluR II is well known to be involved in LTD in the hippocampus [18,19,20] and mGluR III has also been involved in plasticity [21,22]. Depending on the brain region, postsynaptic cell type, and specific intracellular pathways, mGluR I is particularly known for inducing LTD, which can be mediated by either mGluR1 or mGluR5 [23,24]. In addition to their role in LTD, the activation of mGluR I potentiates NMDA-receptor-mediated currents [25,26], and it can also depolarize several types of neurons through activation of a Ca2+-dependent cation conductance [27,28].
Some conceptual controversies exist regarding the role of mGluR I and the type of synaptic plasticity it involves (see Jones, 2017; [29] for review). Notably, this group of metabotropic receptors was proposed as a crucial player in the mechanisms underlying the detection of activity coincidence in the form of synaptic plasticity that depends on the precise temporal coincidence and order of pre- and postsynaptic activities: the spike timing-dependent plasticity (STDP) [30,31,32]. STDP has been found in all of the species in which it has been studied (from insects to humans) [33]. Different forms of STDP have been described depending on the specific cell and synapse type [34,35,36,37,38,39,40], the state of the network [41], neuromodulatory agents [42,43] and the developmental stage [44,45] in which the study was carried out, and it is believed that STDP endows a supporting role to memory formation and maintenance [9,30]. Here, we review and discuss the involvement of mGluR I in STDP and its role as a coincidence detector.

2. The STDP Phenomena and mGluRs Involvement

STDP is a form of synaptic plasticity in which the coincidence of pre- and postsynaptic activities (spiking) within a few milliseconds dictates a long-lasting potentiation (t-LTP) or depression (t-LTD) of synaptic transmission depending on the order of the spiking occurrence. In excitatory synapses, when the presynaptic firing precedes the postsynaptic spiking (“pre-post”), t-LTP is induced, whereas when the order is the inverse (“post-pre”), t-LTD is induced [30,33] (Figure 1). However, there are exceptions according to the recent advances in the field (i.e., post-pre protocol induces t-LTP in the hippocampal area CA1 of mice at postnatal days 35–42 (P35-P42) [44], Figure 1B) (see also Feldman, 2012; [30] for review).
Different forms of STDP were found in diverse brain areas, including the hippocampus [34,44,46], diverse cortical synapses [36,45,47,48,49], the cerebellum [38], the spinal cord [50], the striatum [37] and the amygdala [51] among others, and mGluR I is involved in t-LTP and t-LTD at different synapses in some of the these regions, covering different functions. Thus, in the somatosensory cortex, postsynaptic mGluR I participates in the production and release of eCB, which acts as a retrograde signaling molecule [47]; in neocortical synapses onto interneurons, where it participates in t-LTD [52]; in the substantia gelatinosa, it regulates the polarity of STDP [50] and in corticostriatal synapses, it acts as a coincidence detector for t-LTD [37]. In the cerebellum, mGluR I participates in the induction of t-LTP but not t-LTD [38]. Recently, it has been observed that eyeblink conditioning, a form of Pavlovian learning that engages discrete areas of the cerebellar cortex and deep cerebellar nuclei, is impaired in mGluR1 knockout mice. Moreover, administration of the mGluR1/5 agonist DHPG into the lobulus simplex region of the cerebellar cortex promotes eyeblink conditioning in rats, which indicates that cerebellar mGluR1 plays a role in cerebellar-dependent associative learning [53]. mGluRs have been also found to be involved in STDP in the hippocampus, where they seem to gate NMDAR-mediated t-LTP [54], participate postsynaptically in the induction of a presynaptic form of t-LTD by promoting the production and release of eCB that acts as a retrograde signaling molecule [34,44] or presynaptically participate in a newly discovered form of presynaptic t-LTD [44]. Whether t-LTD and t-LTP occur at the same time on the same synapse, the underlying pathways involved in their expression remain to be fully determined. Additionally, whether these forms of STDP share the same or distinct pools of Ca2+ in the same dendritic spine is still puzzling and deserves further research.

3. Group I mGluR Involvement in the Coincidence Detection for STDP

A mechanism for detection of the pre- and postsynaptic activities and their correlated timings appears crucial for codifying the spiking coincidence that drives STDP [30,33]. In a classical view, postsynaptic NMDARs are the receptors acting as coincidence detectors [5,6,30] (Figure 1A) and their contribution to STDP has been largely documented from either pre- or postsynaptic synaptic compartments [31,33,34,39,47,48,49]. However, presynaptically located NMDARs do not likely hold the role of coincidence detectors in NMDAR-dependent forms of STDP [55]. This evidence suggests the existence of different mechanisms of STDP that are mediated by different coincidence detectors [34,36,39,40,47,56,57,58,59,60] (Figure 1). Therefore, two different heterogeneous classes of STDP could be proposed: one class that involves NMDARs as the primary coincidence detectors [34,36,39,40,47,56,57,58,59] (Figure 1A) and another class that does not require NMDARs, or they do not codify the coincidence [34,39,40,44,46,47,60,61,62,63], and includes pre- or postsynaptic mGluRs (Figure 1B,C). Accordingly, irrespective of NMDAR involvement, as previously mentioned, it is known that STDP requires other players including mGluRs and retrograde messengers such as endocannabinoids (eCBs) or NO and involve astrocytes [34,39,40,44,46,47,60,61,62,63] (Figure 1B,C). Moreover, mGluRs could be proposed as relevant players in the coincidence detection mechanisms for STDP where NMDARs are not involved in such mechanisms.
In this regard, unlike ionotropic glutamate receptors, mGluRs are thought to operate within a larger timescale [6]. Thus, it turns out to be more intriguing how mGluRs can efficiently contribute as detectors of coincident activity that takes place in a time window of a few milliseconds (i.e., 5–10 ms) during few repetitive pairings (i.e., <100 times) as it happens for STDP [30,33]. However, even though GPCRs such as mGluRs are supposed to signal over a timescale of seconds to minutes, mGluR signaling has been suggested to also occur faster and with a timescale similar to that observed for ionotropic glutamate receptors [6]. This notion is supported by conformational studies showing large-scale, activation-associated mGluR1 inter-subunit changes on a millisecond timescale [64].
Consequently, most of the coincidence detector models propose mGluR-voltage-gated Ca2+ channel (VGCC)-IP3R signaling as a principal postsynaptic coincidence detector. In addition, phospholipase C (PLC) and IP3Rs have been proposed for serving well as molecular coincidence detectors. This pathway entails a strong Ca2+ dependence that can synergistically contribute to eCB signaling, thus driving some forms of t-LTD [30,47,65]. In fact, these models find support in several forms of STDP that depend on changes in the cytosolic Ca2+ dynamics involving Ca2+ influx through VGCCs and/or Ca2+ mobilization from internal stores [34,37,38,44,47,50,52,66,67,68,69] (Figure 1). Moreover, there is evidence that in the spine machinery, the release of Ca2+ from internal stores and Ca2+ transients through VGCCs are likely to provide highly localized and input-specific Ca2+ signals to induce synaptic plasticity [65,67].
Particularly, Ca2+ from VGCCs considerably facilitates mGluR-dependent PLC activation, acting independently as a co-agonist of IP3Rs to promote IP3-dependent Ca2+ release [70]. Thus, in this model, a Ca2+ influx through VGCCs during each postsynaptic action potential (spike) could transiently trigger mGluR-IP3 signaling [67,71,72]. This could provide an adequate timing for the presynaptic spiking to drive mGluR signaling and release enough Ca2+ to trigger Ca2+-dependent eCB synthesis and release that ultimately leads to LTD. A similar mechanism has been proposed for short-term synaptic depression with the involvement of VGCCs and mGluR I activation that synergistically drives eCB release [70,73]. This model appears essentially the same as the two-coincidence detector model for STDP that was previously proposed by Karmarkar and Buonomano (2002). Additionally, small increases of IP3, which are not sufficient to stimulate release directly, can enhance the Ca2+ sensitivity of the IP3Rs, thereby transforming the cytoplasm into an excitable medium able to produce Ca2+ waves [65,74]. For example, inhibiting the hydrolysis of IP3 greatly enhances the sensitivity of neurons to synaptic stimulation [65].
As indicated above, mGluR I also mediates a form of cortical t-LTD that is independent of postsynaptic NMDARs and is presynaptically expressed [47]. This form of t-LTD, induced at layer 4 to layer 2/3 synapses in the primary somatosensory cortex, involves postsynaptic mGluRs and retrograde eCB signaling, suggesting a common signaling motif that is in line with the proposed roles of mGluR I in the models of coincidence detection. More recently, in the hippocampus, a presynaptic form of hippocampal t-LTD shows similar properties to that previously described in the neocortical synapses during development (12-18 postnatal days). This form of t-LTD is presynaptic, requires postsynaptic mGluR5, eCB type 1 receptors (CB1R), postsynaptic Ca2+, astrocytic signaling (D-serine release) and non-postsynaptic NMDA receptors at Schaffer collateral-CA1 synapses [34]. Hence, the described mechanism matches with the above-mentioned class of STDP (Figure 1A).
Notably, this form of hippocampal t-LTD switches from depression to potentiation (t-LTP) across a wide range of spike timings as young mice mature towards the fifth postnatal week [44]. Interestingly, this form of t-LTP is also expressed presynaptically and requires the activation of mGluR5, but not NMDARs. In addition, the required activation of mGluR5 appears to reside presynaptically. At these synapses, presynaptic mGluRs have been described to bidirectionally modulate glutamate release [75]. Moreover, glial cells are also thought to express mGluRs that probably contribute to the influence of synaptic plasticity [76,77]. In turn, mGluRs located in the astrocytes appeared not to be involved in this particular form of t-LTP.

4. Concluding Remarks

Cooperative and correlated activity within neuronal circuits underlie information processing through the formation of neuronal network ensembles and short- and long-lasting plastic changes that are associated with and/or underlie higher cognitive processes such as memory formation and recall [13,78,79]. In the case of STDP, cumulative evidence suggests that either pre- or postsynaptic mGluR I may drive t-LTP and t-LTD, acting as a key component of non-classical postsynaptic NMDAR-based detection of activity coincidence. Thus, group I mGluRs have emerged as potent modulators/drivers of significant aspects of neuronal circuit functioning including STDP. Additionally, growing evidence posits mGluRs as suitable drug targets to treat neurological disorders such as anxiety, Parkinson’s disease, autism spectrum disorders, Alzheimer’s disease, fragile X syndrome and drug abuse [4,29,80,81] where STDP has been or could be found impaired. However, despite the great advances in studying the functional role of mGluRs in neuronal circuits and, in particular, in STDP, more research is needed to cover missing mechanistic insights that probably depend on the current experimental and technical limitations. In this regard, a direct demonstration of the functional presence of mGluRs at the presynaptic compartment could be achieved by paired recordings of connected pyramidal neurons while blocking the metabotropic signaling just in the presynaptic neuron, as it was previously performed to demonstrate the functional presence of NMDARs driving t-LTD in layer 4–layer 2/3 of the mouse barrel cortex [39,56]. As well, the occurrence of t-LTD and t-LTP at the same synapse involving presynaptic NMDARs, mGluRs and mGluRs located in the astrocytes could be unveiled by performing the above-mentioned arduous approach. Development of caged mGluR I antagonists that could be uncaged locally would allow for specific determinations of mGluR involvement in STDP at cellular and subcellular compartments.

Author Contributions

Conceptualization, A.R.-M. and Y.A.-T.; writing—original draft preparation, I.M.-G., A.R.-M. and Y.A.-T.; writing—review and editing, A.R.-M. and Y.A.-T.; supervision, A.R.-M. and Y.A.-T.; project administration, A.R.-M.; funding acquisition, A.R.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Regional Government of Andalusia and FEDER, grant number P2000881 and Agencia Estatal de Investigación/FEDER, grant number PID2019-107677GB-I00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bodzęta, A.; Scheefhals, N.; MacGillavry, H.D. Membrane trafficking and positioning of mGluRs at presynaptic and postsynaptic sites of excitatory synapses. Neuropharmacology 2021, 200, 108799. [Google Scholar] [CrossRef] [PubMed]
  2. Nakanishi, S. Molecular Diversity of Glutamate Receptors and Implications for Brain Function. Science 1992, 258, 597–603. [Google Scholar] [CrossRef] [PubMed]
  3. Kroon, T.; Dawitz, J.; Kramvis, I.; Anink, J.; Obermayer, J.; Verhoog, M.B.; Wilbers, R.; Goriounova, N.A.; Idema, S.; Baayen, J.C.; et al. Group I mGluR-mediated activation of martinotti cells inhibits local cortical circuitry in human cortex. Front. Cell. Neurosci. 2019, 13, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Niswender, C.M.; Conn, P.J. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annu. Rev. Pharmacol. Toxicol. 2010, 50, 295–322. [Google Scholar] [CrossRef] [Green Version]
  5. Collingridge, G.L.; Abraham, W.C. Glutamate receptors and synaptic plasticity: The impact of Evans and Watkins. Neuropharmacology 2021, 206, 108922. [Google Scholar] [CrossRef]
  6. Reiner, A.; Levitz, J. Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert. Neuron 2018, 98, 1080–1098. [Google Scholar] [CrossRef] [Green Version]
  7. Mateos-Aparicio, P.; Rodríguez-Moreno, A. The Impact of Studying Brain Plasticity. Front. Cell. Neurosci. 2019, 13, 66. [Google Scholar] [CrossRef] [Green Version]
  8. Dolphin, A.C.; Errington, M.L.; Bliss, T. V Long-term potentiation of the perforant path in vivo is associated with increased glutamate release. Nature 1982, 297, 496–498. [Google Scholar] [CrossRef]
  9. Bliss, T.V.P.; Collingridge, G.L. A synaptic model of memory: long-term potentiation in the hippocampus. Nature 1993, 361, 31–39. [Google Scholar] [CrossRef]
  10. Bi, G.; Poo, M. Synaptic Modifications in Cultured Hippocampal Neurons: Dependence on Spike Timing, Synaptic Strength, and Postsynaptic Cell Type. J. Neurosci. 1998, 18, 10464–10472. [Google Scholar] [CrossRef]
  11. Malenka, R.C.; Bear, M.F. LTP and LTD: an embarrassment of riches. Neuron 2004, 44, 5–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Neves, G.; Cooke, S.F.; Bliss, T.V.P. Synaptic plasticity, memory and the hippocampus: a neural network approach to causality. Nat. Rev. Neurosci. 2008, 9, 65–75. [Google Scholar] [CrossRef] [PubMed]
  13. Andrade-Talavera, Y.; Rodríguez-Moreno, A. Synaptic Plasticity and Oscillations in Alzheimer’s Disease: A Complex Picture of a Multifaceted Disease. Front. Mol. Neurosci. 2021, 14, 696476. [Google Scholar] [CrossRef] [PubMed]
  14. Park, P.; Kang, H.; Sanderson, T.M.; Bortolotto, Z.A.; Georgiou, J.; Zhuo, M.; Kaang, B.K.; Collingridge, G.L. The Role of Calcium-Permeable AMPARs in Long-Term Potentiation at Principal Neurons in the Rodent Hippocampus. Front. Synaptic Neurosci. 2018, 10, 1–11. [Google Scholar] [CrossRef] [Green Version]
  15. Bashir, Z.I.; Bortolotto, Z.A.; Davies, C.H.; Berretta, N.; Irving, A.J.; Seal, A.J.; Henley, J.M.; Jane, D.E.; Watkins, J.C.; Collingridge, G.L. Induction of LTP in the hippocampus needs synaptic activation of glutamate metabotropic receptors. Nature 1993, 363, 347–350. [Google Scholar] [CrossRef]
  16. Bortolotto, Z.A.; Bashir, Z.I.; Davies, C.H.; Collingridge, G.L. A molecular switch activated by metabotropic glutamate receptors regulates induction of long-term potentiation. Nature 1994, 368, 740–743. [Google Scholar] [CrossRef]
  17. Tigaret, C.M.; Chamberlain, S.E.L.; Sadowski, J.H.L.P.; Hall, J.; Ashby, M.C.; Mellor, J.R. Convergent Metabotropic Signaling Pathways Inhibit SK Channels to Promote Synaptic Plasticity in the Hippocampus. J. Neurosci. 2018, 38, 9252–9262. [Google Scholar] [CrossRef] [Green Version]
  18. Kobayashi, K.; Manabe, T.; Takahashi, T. Presynaptic Long-Term Depression at the Hippocampal Mossy Fiber—CA3 Synapse. Science 1996, 273, 648–650. [Google Scholar] [CrossRef]
  19. Negrete-Díaz, J.V.; Sihra, T.S.; Delgado-García, J.M.; Rodríguez-Moreno, A. Kainate receptor-mediated presynaptic inhibition converges with presynaptic inhibition mediated by Group II mGluRs and long-term depression at the hippocampal mossy fiber-CA3 synapse. J. Neural Transm. 2007, 114, 1425–1431. [Google Scholar] [CrossRef]
  20. Lyon, L.; Borel, M.; Carrión, M.; Kew, J.N.C.; Corti, C.; Harrison, P.J.; Burnet, P.W.J.; Paulsen, O.; Rodríguez-Moreno, A. Hippocampal mossy fiber long-term depression in Grm2/3 double knockout mice. Synapse 2011, 65, 945–954. [Google Scholar] [CrossRef]
  21. Mercier, M.S.; Lodge, D. Group III metabotropic glutamate receptors: pharmacology, physiology and therapeutic potential. Neurochem. Res. 2014, 39, 1876–1894. [Google Scholar] [CrossRef] [PubMed]
  22. Dasgupta, A.; Lim, Y.J.; Kumar, K.; Baby, N.; Pang, K.L.K.; Benoy, A.; Behnisch, T.; Sajikumar, S. Group III metabotropic glutamate receptors gate long-term potentiation and synaptic tagging/capture in rat hippocampal area CA2. Elife 2020, 9, e55344. [Google Scholar] [CrossRef]
  23. Lüscher, C.; Huber, K.M. Group 1 mGluR-Dependent Synaptic Long-Term Depression: Mechanisms and Implications for Circuitry and Disease. Neuron 2010, 65, 445–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sherman, S.M. The function of metabotropic glutamate receptors in thalamus and cortex. Neuroscientist 2014, 20, 136–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wang, X.F.; Daw, N.W. Metabotropic glutamate receptors potentiate responses to NMDA and AMPA from layer V cells in rat visual cortex. J. Neurophysiol. 1996, 76, 808–815. [Google Scholar] [CrossRef]
  26. Mannaioni, G.; Marino, M.J.; Valenti, O.; Traynelis, S.F.; Conn, P.J. Metabotropic glutamate receptors 1 and 5 differentially regulate CA1 pyramidal cell function. J. Neurosci. 2001, 21, 5925–5934. [Google Scholar] [CrossRef] [Green Version]
  27. Guérineau, N.C.; Bossu, J.L.; Gähwiler, B.H.; Gerber, U. Activation of a nonselective cationic conductance by metabotropic glutamatergic and muscarinic agonists in CA3 pyramidal neurons of the rat hippocampus. J. Neurosci. 1995, 15, 4395–4407. [Google Scholar] [CrossRef]
  28. Guérineau, N.C.; Gähwiler, B.H.; Gerber, U. Reduction of resting K+ current by metabotropic glutamate and muscarinic receptors in rat CA3 cells: mediation by G-proteins. J. Physiol. 1994, 474, 27–33. [Google Scholar] [CrossRef] [Green Version]
  29. Jones, O.D. Do group I metabotropic glutamate receptors mediate LTD? Neurobiol. Learn. Mem. 2017, 138, 85–97. [Google Scholar] [CrossRef]
  30. Feldman, D.E. The Spike-Timing Dependence of Plasticity. Neuron 2012, 75, 556–571. [Google Scholar] [CrossRef] [Green Version]
  31. Corlew, R.; Brasier, D.J.; Feldman, D.E.; Philpot, B.D. Presynaptic NMDA receptors: newly appreciated roles in cortical synaptic function and plasticity. Neuroscientist 2008, 14, 609–625. [Google Scholar] [CrossRef] [Green Version]
  32. Rodriguez-Moreno, A.; Banerjee, A.; Paulsen, O. Presynaptic NMDA Receptors and Spike Timing-Dependent Depression at Cortical Synapses. Front. Synaptic Neurosci. 2010, 2, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Markram, H.; Gerstner, W.; Sjöström, P.J. Spike-Timing-Dependent Plasticity: A Comprehensive Overview. Front. Synaptic Neurosci. 2012, 4, 2010–2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Andrade-Talavera, Y.; Duque-Feria, P.; Paulsen, O.; Rodríguez-Moreno, A. Presynaptic Spike Timing-Dependent Long-Term Depression in the Mouse Hippocampus. Cereb. Cortex 2016, 26, 3637–3654. [Google Scholar] [CrossRef] [Green Version]
  35. Min, R.; Nevian, T. Astrocyte signaling controls spike timing-dependent depression at neocortical synapses. Nat. Neurosci. 2012, 15, 746–753. [Google Scholar] [CrossRef] [PubMed]
  36. Banerjee, A.; González-Rueda, A.; Sampaio-Baptista, C.; Paulsen, O.; Rodríguez-Moreno, A. Distinct mechanisms of spike timing-dependent LTD at vertical and horizontal inputs onto L2/3 pyramidal neurons in mouse barrel cortex. Physiol. Rep. 2014, 2, e00271. [Google Scholar] [CrossRef]
  37. Fino, E.; Paille, V.; Cui, Y.; Morera-Herreras, T.; Deniau, J.-M.; Venance, L. Distinct coincidence detectors govern the corticostriatal spike timing-dependent plasticity. J. Physiol. 2010, 588, 3045–3062. [Google Scholar] [CrossRef] [PubMed]
  38. Sgritta, M.; Locatelli, F.; Soda, T.; Prestori, F.; D’Angelo, E.U. Hebbian Spike-Timing Dependent Plasticity at the Cerebellar Input Stage. J. Neurosci. 2017, 37, 2809–2823. [Google Scholar] [CrossRef]
  39. Banerjee, A.; Meredith, R.M.; Rodríguez-Moreno, A.; Mierau, S.B.; Auberson, Y.P.; Paulsen, O. Double dissociation of spike timing-dependent potentiation and depression by subunit-preferring NMDA receptor antagonists in mouse barrel cortex. Cereb. Cortex 2009, 19, 2959–2969. [Google Scholar] [CrossRef]
  40. Duguid, I.; Sjöström, P.J. Novel presynaptic mechanisms for coincidence detection in synaptic plasticity. Curr. Opin. Neurobiol. 2006, 16, 312–322. [Google Scholar] [CrossRef]
  41. Kwag, J.; Paulsen, O. The timing of external input controls the sign of plasticity at local synapses. Nat. Neurosci. 2009, 12, 1219–1221. [Google Scholar] [CrossRef] [PubMed]
  42. Brzosko, Z.; Schultz, W.; Paulsen, O. Retroactive modulation of spike timing-dependent plasticity by dopamine. Elife 2015, 4, e09685. [Google Scholar] [CrossRef]
  43. Brzosko, Z.; Mierau, S.B.; Paulsen, O. Neuromodulation of Spike-Timing-Dependent Plasticity: Past, Present, and Future. Neuron 2019, 103, 563–581. [Google Scholar] [CrossRef] [PubMed]
  44. Falcón-Moya, R.; Pérez-Rodríguez, M.; Prius-Mengual, J.; Andrade-Talavera, Y.; Arroyo-García, L.E.; Pérez-Artés, R.; Mateos-Aparicio, P.; Guerra-Gomes, S.; Oliveira, J.F.; Flores, G.; et al. Astrocyte-mediated switch in spike timing-dependent plasticity during hippocampal development. Nat. Commun. 2020, 11, 4388. [Google Scholar] [CrossRef] [PubMed]
  45. Martínez-Gallego, I.; Pérez-Rodríguez, M.; Coatl-Cuaya, H.; Flores, G.; Rodríguez-Moreno, A. Adenosine and astrocytes determine the developmental dynamics of spike timing-dependent plasticity in the somatosensory cortex. J. Neurosci. 2022. ahead of print. [Google Scholar] [CrossRef]
  46. Pérez-Rodríguez, M.; Arroyo-García, L.E.; Prius-Mengual, J.; Andrade-Talavera, Y.; Armengol, J.A.; Pérez-Villegas, E.M.; Duque-Feria, P.; Flores, G.; Rodríguez-Moreno, A. Adenosine Receptor-Mediated Developmental Loss of Spike Timing-Dependent Depression in the Hippocampus. Cereb. Cortex 2019, 29, 3266–3281. [Google Scholar] [CrossRef] [Green Version]
  47. Bender, V.A.; Bender, K.J.; Brasier, D.J.; Feldman, D.E. Two coincidence detectors for spike timing-dependent plasticity in somatosensory cortex. J. Neurosci. 2006, 26, 4166–4177. [Google Scholar] [CrossRef]
  48. Brasier, D.J.; Feldman, D.E. Synapse-Specific Expression of Functional Presynaptic NMDA Receptors in Rat Somatosensory Cortex. J. Neurosci. 2008, 28, 2199–2211. [Google Scholar] [CrossRef]
  49. Rodríguez-Moreno, A.; Paulsen, O. Spike timing-dependent long-term depression requires presynaptic NMDA receptors. Nat. Neurosci. 2008, 11, 744–745. [Google Scholar] [CrossRef]
  50. Jung, S.J.; Kim, S.J.; Park, Y.K.; Oh, S.B.; Cho, K.; Kim, J. Group I mGluR regulates the polarity of spike-timing dependent plasticity in substantia gelatinosa neurons. Biochem. Biophys. Res. Commun. 2006, 347, 509–516. [Google Scholar] [CrossRef]
  51. Jung, S.-Y.; Kim, J.; Kwon, O. Bin; Jung, J.H.; An, K.; Jeong, A.Y.; Lee, C.J.; Choi, Y.-B.; Bailey, C.H.; Kandel, E.R.; et al. Input-specific synaptic plasticity in the amygdala is regulated by neuroligin-1 via postsynaptic NMDA receptors. Proc. Natl. Acad. Sci. USA 2010, 107, 4710–4715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lu, J.T.; Li, C.Y.; Zhao, J.P.; Poo, M.M.; Zhang, X.H. Spike-timing-dependent plasticity of neocortical excitatory synapses on inhibitory interneurons depends on target cell type. J. Neurosci. 2007, 27, 9711–9720. [Google Scholar] [CrossRef] [PubMed]
  53. Shipman, M.L.; Madasu, S.C.; Morielli, A.D.; Green, J.T. Intracerebellar infusion of an mGluR1/5 agonist enhances eyeblink conditioning. Behav. Neurosci. 2021, 135, 336–342. [Google Scholar] [CrossRef] [PubMed]
  54. Kwag, J.; Paulsen, O. Gating of NMDA receptor-mediated hippocampal spike timing-dependent potentiation by mGluR5. Neuropharmacology 2012, 63, 701–709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Wong, H.H.W.; Rannio, S.; Jones, V.; Thomazeau, A.; Sjöström, P.J. NMDA receptors in axons: There’s no coincidence. J. Physiol. 2021, 599, 367–387. [Google Scholar] [CrossRef]
  56. Bi, G.-Q.; Rubin, J. Timing in synaptic plasticity: from detection to integration. Trends Neurosci. 2005, 28, 222–228. [Google Scholar] [CrossRef] [PubMed]
  57. Froemke, R.C.; Poo, M.-M.; Dan, Y. Spike-timing-dependent synaptic plasticity depends on dendritic location. Nature 2005, 434, 221–225. [Google Scholar] [CrossRef]
  58. Nishiyama, M.; Hong, K.; Mikoshiba, K.; Poo, M.M.; Kato, K. Calcium stores regulate the polarity and input specificity of synaptic modification. Nature 2000, 408, 584–588. [Google Scholar] [CrossRef]
  59. Shouval, H.Z.; Bear, M.F.; Cooper, L.N. A unified model of NMDA receptor-dependent bidirectional synaptic plasticity. Proc. Natl. Acad. Sci. USA 2002, 99, 10831–10836. [Google Scholar] [CrossRef] [Green Version]
  60. Rodriguez-Moreno, A.; Kohl, M.M.; Reeve, J.E.; Eaton, T.R.; Collins, H.A.; Anderson, H.L.; Paulsen, O.; Rodríguez-Moreno, A.; Kohl, M.M.; Reeve, J.E.; et al. Presynaptic induction and expression of timing-dependent long-term depression demonstrated by compartment-specific photorelease of a use-dependent NMDA receptor antagonist. J. Neurosci. 2011, 31, 8564–8569. [Google Scholar] [CrossRef]
  61. Egger, V.; Feldmeyer, D.; Sakmann, B. Coincidence detection and changes of synaptic efficacy in spiny stellate neurons in rat barrel cortex. Nat. Neurosci. 1999, 2, 1098–1105. [Google Scholar] [CrossRef] [PubMed]
  62. Bouvier, G.; Larsen, R.S.; Rodríguez-Moreno, A.; Paulsen, O.; Sjöström, P.J. Towards resolving the presynaptic NMDA receptor debate. Curr. Opin. Neurobiol. 2018, 51, 1–7. [Google Scholar] [CrossRef]
  63. Pérez-Otaño, I.; Rodríguez-Moreno, A. Presynaptic NMDARs and astrocytes ally to control circuit-specific information flow. Proc. Natl. Acad. Sci. USA 2019, 116, 13166–13168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Hlavackova, V.; Zabel, U.; Frankova, D.; Bätz, J.; Hoffmann, C.; Prezeau, L.; Pin, J.-P.; Blahos, J.; Lohse, M.J. Sequential inter- and intrasubunit rearrangements during activation of dimeric metabotropic glutamate receptor 1. Sci. Signal. 2012, 5, ra59. [Google Scholar] [CrossRef] [PubMed]
  65. Berridge, M.J. Neuronal calcium signaling. Neuron 1998, 21, 13–26. [Google Scholar] [CrossRef] [Green Version]
  66. Inglebert, Y.; Debanne, D. Calcium and Spike Timing-Dependent Plasticity. Front. Cell. Neurosci. 2021, 15, 727336. [Google Scholar] [CrossRef]
  67. Nevian, T.; Sakmann, B. Spine Ca2+ Signaling in Spike-Timing-Dependent Plasticity. J. Neurosci. 2006, 26, 11001–11013. [Google Scholar] [CrossRef] [Green Version]
  68. Mateos-Aparicio, P.; Rodríguez-Moreno, A. Calcium Dynamics and Synaptic Plasticity. In Advances in Experimental Medicine and Biology; Springer: Cham, Switzerland, 2020; pp. 965–984. ISBN 978-3-030-12457-1. [Google Scholar]
  69. Cepeda-Prado, E.A.; Khodaie, B.; Quiceno, G.D.; Beythien, S.; Edelmann, E.; Lessmann, V. Calcium-Permeable AMPA Receptors Mediate Timing-Dependent LTP Elicited by Low Repeat Coincident Pre- and Postsynaptic Activity at Schaffer Collateral-CA1 Synapses. Cereb. Cortex 2021, 32, 1682–1703. [Google Scholar] [CrossRef]
  70. Hashimotodani, Y.; Ohno-Shosaku, T.; Tsubokawa, H.; Ogata, H.; Emoto, K.; Maejima, T.; Araishi, K.; Shin, H.-S.; Kano, M. Phospholipase Cbeta serves as a coincidence detector through its Ca2+ dependency for triggering retrograde endocannabinoid signal. Neuron 2005, 45, 257–268. [Google Scholar] [CrossRef] [Green Version]
  71. Sabatini, B.L.; Maravall, M.; Svoboda, K. Ca(2+) signaling in dendritic spines. Curr. Opin. Neurobiol. 2001, 11, 349–356. [Google Scholar] [CrossRef]
  72. Bloodgood, B.L.; Sabatini, B.L. Ca(2+) signaling in dendritic spines. Curr. Opin. Neurobiol. 2007, 17, 345–351. [Google Scholar] [CrossRef] [PubMed]
  73. Kreitzer, A.C.; Malenka, R.C. Dopamine modulation of state-dependent endocannabinoid release and long-term depression in the striatum. J. Neurosci. 2005, 25, 10537–10545. [Google Scholar] [CrossRef] [PubMed]
  74. Lorenzon, P.; Zacchetti, D.; Codazzi, F.; Fumagalli, G.; Meldolesi, J.; Grohovaz, F. Ca2+ waves in PC12 neurites: a bidirectional, receptor-oriented form of Ca2+ signaling. J. Cell Biol. 1995, 129, 797–804. [Google Scholar] [CrossRef] [PubMed]
  75. Rodríguez-Moreno, A.; Sistiaga, A.; Lerma, J.; Sánchez-Prieto, J. Switch from facilitation to inhibition of excitatory synaptic transmission by group I mGluR desensitization. Neuron 1998, 21, 1477–1486. [Google Scholar] [CrossRef]
  76. 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]
  77. Perea, G.; Araque, A. Astrocytes Potentiate Transmitter Release at Single Hippocampal Synapses. Science 2007, 317, 1083–1086. [Google Scholar] [CrossRef]
  78. Buzsáki, G.; Draguhn, A. Neuronal oscillations in cortical networks. Science 2004, 304, 1926–1929. [Google Scholar] [CrossRef] [Green Version]
  79. McBain, C.J.; Fisahn, A. Interneurons unbound. Nat. Rev. Neurosci. 2001, 2, 11–23. [Google Scholar] [CrossRef]
  80. Gray, E.E.; Murphy, J.G.; Liu, Y.; Trang, I.; Tabor, G.T.; Lin, L.; Hoffman, D.A. Disruption of GpI mGluR-Dependent Cav2.3 Translation in a Mouse Model of Fragile X Syndrome. J. Neurosci. 2019, 39, 7453–7464. [Google Scholar] [CrossRef]
  81. Crupi, R.; Impellizzeri, D.; Cuzzocrea, S. Role of Metabotropic Glutamate Receptors in Neurological Disorders. Front. Mol. Neurosci. 2019, 12, 20. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic summarizing the role of mGluR I in different forms of STDP, showing that two different large sets of STDP forms could be proposed according to the underlying mechanism for coincidence detection: (A) In a classic model of Hebbian STDP, postsynaptic NMDARs are the main coincidence detectors (providing strong and brief Ca2+ signals) that drive postsynaptic forms of t-LTP and could also involve postsynaptic VGCCs, postsynaptic mGluRs and IP3R-mediated increase in postsynaptic Ca2+ [30,33,34,47,50]. This model does not fully support the mechanism underlying postsynaptic NMDAR-dependent t-LTD at horizontal layer 2/3-layer 2/3 synapses of the primary somatosensory cortex [36]. Consequently, more research needs to be performed. In turn, other findings support the proposed model involving the mGluR-VGCC-IP3R pathway in presynaptic forms of t-LTD as shown in B and t-LTP as shown in C as representative examples. In these studies, presynaptic forms of STDP either involve non-postsynaptic and likely presynaptic NMDARs and eCBs as retrograde signal driving to t-LTD, as represented in B, [34,47] or NO retrograde signal driving to NMDAR-independent t-LTP, as represented in C [44]. In addition, astrocytes release D-serine for t-LTD at P13-P21 (B) and ATP/adenosine for t-LTP at P35-42 (C). The presence and involvement of astrocytic mGluRs in diverse forms of STDP should not be discarded and deserve further experimental efforts. Note that at hippocampal area CA1, a developmental switch occurs: when the postsynaptic activity precedes the presynaptic activity (∆t < 0) at P13-21, presynaptic t-LTD is induced (B) whereas presynaptic t-LTP is induced by the same protocol at older ages (P35-P42) (C) with astrocytes commanding the closing and opening of these plasticity windows [44,46]. Such general models represented in (AC) are constantly evolving and, therefore, must be revisited considering future advances. (-): decrease in glutamate release in the presynaptic form of t-LTD, (+): pathways that are activated, (?): mechanistic insights that are yet to be demonstrated.
Figure 1. Schematic summarizing the role of mGluR I in different forms of STDP, showing that two different large sets of STDP forms could be proposed according to the underlying mechanism for coincidence detection: (A) In a classic model of Hebbian STDP, postsynaptic NMDARs are the main coincidence detectors (providing strong and brief Ca2+ signals) that drive postsynaptic forms of t-LTP and could also involve postsynaptic VGCCs, postsynaptic mGluRs and IP3R-mediated increase in postsynaptic Ca2+ [30,33,34,47,50]. This model does not fully support the mechanism underlying postsynaptic NMDAR-dependent t-LTD at horizontal layer 2/3-layer 2/3 synapses of the primary somatosensory cortex [36]. Consequently, more research needs to be performed. In turn, other findings support the proposed model involving the mGluR-VGCC-IP3R pathway in presynaptic forms of t-LTD as shown in B and t-LTP as shown in C as representative examples. In these studies, presynaptic forms of STDP either involve non-postsynaptic and likely presynaptic NMDARs and eCBs as retrograde signal driving to t-LTD, as represented in B, [34,47] or NO retrograde signal driving to NMDAR-independent t-LTP, as represented in C [44]. In addition, astrocytes release D-serine for t-LTD at P13-P21 (B) and ATP/adenosine for t-LTP at P35-42 (C). The presence and involvement of astrocytic mGluRs in diverse forms of STDP should not be discarded and deserve further experimental efforts. Note that at hippocampal area CA1, a developmental switch occurs: when the postsynaptic activity precedes the presynaptic activity (∆t < 0) at P13-21, presynaptic t-LTD is induced (B) whereas presynaptic t-LTP is induced by the same protocol at older ages (P35-P42) (C) with astrocytes commanding the closing and opening of these plasticity windows [44,46]. Such general models represented in (AC) are constantly evolving and, therefore, must be revisited considering future advances. (-): decrease in glutamate release in the presynaptic form of t-LTD, (+): pathways that are activated, (?): mechanistic insights that are yet to be demonstrated.
Ijms 23 07807 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínez-Gallego, I.; Rodríguez-Moreno, A.; Andrade-Talavera, Y. Role of Group I Metabotropic Glutamate Receptors in Spike Timing-Dependent Plasticity. Int. J. Mol. Sci. 2022, 23, 7807. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23147807

AMA Style

Martínez-Gallego I, Rodríguez-Moreno A, Andrade-Talavera Y. Role of Group I Metabotropic Glutamate Receptors in Spike Timing-Dependent Plasticity. International Journal of Molecular Sciences. 2022; 23(14):7807. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23147807

Chicago/Turabian Style

Martínez-Gallego, Irene, Antonio Rodríguez-Moreno, and Yuniesky Andrade-Talavera. 2022. "Role of Group I Metabotropic Glutamate Receptors in Spike Timing-Dependent Plasticity" International Journal of Molecular Sciences 23, no. 14: 7807. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23147807

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