Open access peer-reviewed chapter

Kainate Receptors Modulating Glutamate Release in the Cerebellum

Written By

Pilar Losada-Ruiz, Rafael Falcón-Moya and Antonio Rodríguez-Moreno

Submitted: 19 March 2019 Reviewed: 12 June 2019 Published: 24 September 2019

DOI: 10.5772/intechopen.87984

From the Edited Volume

Biogenic Amines in Neurotransmission and Human Disease

Edited by Ahmet Uçar

Chapter metrics overview

1,036 Chapter Downloads

View Full Metrics

Abstract

Glutamate receptors of the kainate type (Kainate receptors, KARs), are mediators of ionotropic postsynaptic synaptic transmission, as well as presynaptic modulators of neurotransmitter release where they show both ionotropic and metabotropic actions regulating glutamate and γ-aminobutiric acid (GABA) release. The mechanisms underlying these modulatory roles are starting to be understood at some brain regions. Here we review the KARs roles and mechanisms involved in the modulation of glutamate release in the cerebellum at parallel fibers (PF)-Purkinje Cells (PuC) synapses. KARs activation mediate a biphasic effect on glutamate release at this synapse, with low kainate (KA) concentrations mediating a facilitation of glutamate release and higher KA concentrations mediating a depression of glutamate release. KA-mediated facilitation is prevented by antagonizing KARs, by inhibition of PKA or stimulation of adenylyl cyclase (AC), by blocking Ca2+ permeant KARs, by depleting intracellular Ca2+ stores and by blocking calmodulin. Thus, at cerebellar parallel fiber-Purkinje cell synapses, presynaptic KARs mediate glutamate release facilitation through Ca2+-calmodulin dependent activation of adenylyl cyclase/cAMP/protein kinase A signaling. KAR-mediated depression of glutamate release involves the AC/cAMP/PKA pathway as for facilitation but not Ca2+-calmodulin, being in this case AC activated by a Gi/o protein to mediate a depression of glutamate release.

Keywords

  • cerebellum
  • KARs
  • glutamate release
  • presynaptic
  • PKA
  • adenylate cyclase
  • Ca2+ calmodulin

1. Introduction

Glutamate is the most abundant excitatory neurotransmitter in the central nervous system (CNS) of mammals. Glutamate mediates its actions by activating glutamate receptors. These receptors participate in normal synaptic transmission at different synapses, in plasticity processes as long-term potentiation (LTP) and long-term depression (LTD) that are considered the cellular and molecular correlation of memory and learning processes and in synaptogenesis and neuronal maturation and, additionally, failure in the functioning of this system can be the origin of some types of epilepsy and may contribute to the development of CNS disorders such as Alzheimer’s disease, Huntington’s Korea, amyotrophic lateral sclerosis, Parkinson’s disease, hypoglycemia, or cerebral ischemia [1, 2, 3].

Glutamate receptors are classically divided into two large families: ionotropic and metabotropic. Ionotropic glutamate receptors (iGluRs) participate in rapid neurotransmission in the nervous system; these ionotropic receptors are classified into three types depending on the agonist that activates them with higher affinity: N-methyl-d-aspartic acid (NMDA) receptors (NMDARs), α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors (AMPARs), and kainate receptors (KARs). These receptors form a channel with different selectivity depending on their subunit composition, all of them being permeable to Na+ and K+ and, additionally, NMDARs are permeable to Ca2+ been some AMPARs and KARs also permeable to Ca2+ depending on subunit composition. They are integral membrane proteins, formed by four subunits (tetramers), being homomers or heteromers [1, 2].

Metabotropic glutamate receptors (mGluRs), which participate also in neurotransmission, are coupled to G proteins and are divided into eight types (mGluR 1–8) and three groups of receptors: group I mGluRs includes mGluR1 and mGluR5 receptors. These receptors are positively coupled to phospholipase C (PLC), which facilitates the conversion of inositol diphosphate (PIP2) to diacylglycerol (DAG) and inositol triphosphate (IP3). DAG activates protein kinase C (PKC) that phosphorylates different substrates and IP3 causes numerous intracellular effects, including the facilitation of Ca2+ release from intracellular stores. Group II mGluRs includes mGluR2 and mGluR3 receptors. These receptors are negatively coupled to adenylate cyclase-mediated AMPc formation, and group III mGluRs includes mGluR4, mGluR6, mGluR7, and mGluR8 receptors. These receptors are negatively coupled to the formation of AMPc mediated by adenylate cyclase [4].

1.1 Kainate receptors

Kainate (KA) is a potent neurotoxin derived from the alga Digenea simplex. The word “Kainic” is derived from the Japanese “Kaininso” (“Makuri”), which means “the ghost of the sea,” and it is an agonist for both KARs and AMPARs (in the same way that the AMPARs agonist AMPA may activate KARs). Kainate is classically known for its potent epileptogenic actions [5, 6].

KA (and other agonists) activates KARs that are tetramers that resulted from different combinations of five subunits called GluK1, GluK2, GluK3, GluK4, and GluK5 (formerly known as GluR5, GluR6, GluR7 and KA1, and KA2, respectively). Of these subunits, GluK1 and GluK3 may form homomeric or heteromeric functional receptors, while GluK4 and GluK5 may only participate in functional receptors when associated with any of the GluK1, GluK2, or GluK3 subunits, but they do not combine with subunits of AMPARs [1, 7, 8].

KARs have been described in different invertebrates such as nematodes and flies [9, 10] and in different species of vertebrates such as amphibia, fish, and birds [11, 12, 13] in addition to mammals. In mammals, KARs have been observed virtually throughout the entire nervous system, although their subcellular location has not been yet fully determined. KARs are widely distributed throughout the CNS and found in the main cells and interneurons of the hippocampus, lateral amygdala, dorsal root ganglia, bipolar cells of the retina, cerebral cortex, and the cerebellum [14, 15].

The lack of knowledge about these receptors compared to other glutamatergic receptors (AMPARs or NMDARs) has been due to the lack of good agonist and antagonist for receptors with particular subunit compositions and to the absence of specific antibodies for the different subunits of KARs, being therefore a significant limitation when exploring the distribution of these receptors. However, by using in situ hybridization techniques, it has been observed that the cells that present a significant expression of the kainate-type subunits GluK1, GluK2, GluK3, and GluK5 are distributed throughout the CNS, including nucleus striatum, hippocampus, cortex, and cerebellum [16]. Likewise, there is a high expression of the GluK4 subunit in the CA3 region of the hippocampus, as well as in the granular neurons of the dentate gyrus. The messenger of the GluK5 subunit, on the other hand, appears more abundantly and more extensively than that of the GluK4 subunit or those of the other subunits [15]. Because the in situ hybridization technique is informative and cannot reveal the subcellular distribution of a specific subunit, and because of pharmacological limitations, there is still much to know about the subcellular location and physiology of these receptors.

Kainate-type glutamate receptors are well established mediators of canonical, ionotropic postsynaptic synaptic transmission and, presynaptically, have a modulatory role in regulating neurotransmitter release. In the latter regard, KARs have been shown to have a noncanonical metabotropic capacity, whereby they affect the control of both glutamate and GABA release, for review see [15, 17, 18, 19, 20, 21, 22]. At some excitatory glutamatergic synapses, KARs’ activation can actually effect biphasic modulation, where low agonist concentrations facilitate glutamate release, while high concentrations decrease the release of the neurotransmitter, for review see [17, 18]. Mechanistic details of how this is achieved are subject of investigation and, indeed, the subcellular location of KARs responsible for presynaptic modulation remains contentious. Different roles of KARs in plasticity have also been described either in LTP or LTD, see [23] for a review of the role of KARs in plasticity.

As other glutamate receptors, KARs are directly or indirectly involved in different diseases, alterations of the nervous system and neurodegeneration and cell death processes. As previously indicated, KA is a potent neurotoxin that directly induces epilepsy and is used as a temporal lobe epilepsy model [5, 6]. Several lines of research indicate that KA directly activating KARs is involved in excitatory and inhibitory imbalances associated with epilepsy. The use of animal models for epilepsy through the use of KA injections has allowed to reproduce in great detail the symptoms observed in humans. The majority of studies of KARs’ involvement in epilepsy have studied acute KA-induced seizures [24, 25, 26, 27]. The best demonstrations of a mechanism for KARs’ involvement in acute epilepsy come from studies of inhibition of GABA release by the activation of presynaptic KA receptors at interneuron-CA1 hippocampal synapses [24, 28, 29]. In chronic epilepsy, a role of KARs has been demonstrated at hippocampal mossy fibers making aberrant synapses onto granule cells of dentate gyrus expressing high number of KARs [30, 31, 32] reviewed in [6, 33]. In humans, genetic studies of members of a family affected by idiopathic juvenile absence epilepsy found elevated levels of Grik1 polymorphisms [34], and in TLE patients, GluK1 subunit containing KARs’ increased levels have also been found [35]. In clinical studies, NS1209 (an AMPA/KARs’ antagonist) has been found to decrease epileptic symptoms [36].

Different studies of neurotoxicity clearly indicate that KARs might be important targets for neuroprotection in neurons and glial cells. The mechanism by which KARs produce excitotoxicity and neuronal cell death is not well understood mainly because of the limitations in appropriate pharmacological tools. Toxicity of KARs involved in multiple sclerosis has also been found onto oligodendrocytes and myelin related to [37, 38], and damage has also been found at axonal levels, where AMPA/KARs’ antagonists prevent it [39]. Interestingly, KARs have also been involved in pain. They are present at dorsal roots activating nociceptors, actually there are clinical trials using KARs’ antagonists to try to prevent pain showing some levels of analgesia [36]. Additionally, KARs have been involved in ischemia [40, 41], migraine pain [36], Alzheimer’s disease [42], Parkinson’s disease [43, 44], Huntington’s Chorea [45, 46, 47], Schizophrenia [48, 49], depression [50], bipolar disorder [51, 52], mental retardation [53], and autism [54, 55] as reviewed in [56]. In general, antagonists of KARs containing particular subunits might be good targets to ameliorate symptoms or treat different CNS diseases and alterations.

Advertisement

2. KARs in the cerebellum

As indicated above, KARs are expressed in the cerebellar cortex [57, 58, 59]. As known, the cerebellum participates in the modulation of movement by modifying the activity patterns of motor neurons. Structurally, the cerebellum is composed of the laminar cerebellar cortex and the deep cerebellar nuclei and has five types of cells: Purkinje, stellate, basket, Golgi, and granule cells. Purkinje cells (PuC) are aligned in front of each other. Their dendritic trees form two 2-dimensional layers through which parallel fibers from the mossy fibers located in the granular layer pass. These parallel fibers (PF) establish excitatory synapses between granular cells and the spines of the PCs dendrites as well as the climbing fibers (CF, originating from the inferior olivary nucleus) with the nearby dendrites and the cellular soma. The parallel fibers pass orthogonally through the dendritic tree of the Purkinje neuron. Up to 200,000 PF form a synapse with a single PuC. Each PuC receives up to 500 synapses of CF, all originated from a single CF. Both basket cells and stellate cells provide an inhibitory (GABAergic) entry to the PuC, with cells in the basket synapse to the initial segment of the PuC axon, and stellate cells to the dendrites [60, 61].

Presynaptic KARs participate in plasticity in the cerebellum where PF synapses onto PuCs mediate a form of LTD that is affected by the paired activation of CFs [62], Table 1; of these two types of fibers (PF and CF synapsing onto the same cells (PC), only PF have presynaptic KARs [62], similar to other brain regions as somatosensory and visual cortices in which fibers containing and noncontaining presynaptic ionotropic glutamate receptors synapse onto the same postsynaptic cell and induce LTD [63, 64, 65, 66, 67, 68, 69]. The exact role and action mechanism of KARs mediating LTD in the cerebellum are not well known yet and await further investigation.

KARs’ activation High concentrations of kainate Depression of glutamatergic synaptic transmission Delaney and Jahr [86]
Low concentrations of kainate Facilitation of glutamatergic synaptic synaptic transmission Falcón-Moya et al. [80]
Ionic imbalance Calcification Korf and Postema [78]
Increase in Ca2+ Neurodegeneration
Nodular cerebellum lesion
Putrescine increase
Histological damage
Ataxia Maiti et al. [72]
de Vera et al. [73]
Yamaguchi et al. [74]
Andoh et al. [75]
Parallel fibers paired with postsynaptic depolarization Presynaptic KARs’ activation Long-term depression Crépel [62]
Increase of GluR6 and GluK2 receptors Reduction of GABAergic activity Schizophrenia Harrison et al. [76]
Bullock et al. [77]

Table 1.

KARs’ actions in the cerebellum.

The proper cerebellum development depends on a precise coordinated sequence of postnatal events, some of which are mediated by glutamate receptors. For example, NMDA receptors have been implicated in the migration of granular cells [70] and in the synaptic pruning of climbing fibers [71]. Although it has recently been shown that KARs are involved in synaptic transmission, little is known about their role in development. However, the expression of kainate-type glutamate receptor subunits in immature granule cells of the outer germinal layer of the developing cerebellum suggests that KARs may also have a role in neuronal maturation. Throughout the maturation process of the cerebellum, the quantity, composition, and function of KARs vary. Initially, cerebellar granular cells have a minimal amount of AMPARs in the postnatal period compared to KARs, which are predominant in immature granule cells. Different studies have shown that KARs composed of subunits GluK1, GluK2, and GluK5 predominate, and over the period of development, an increase in the number of KARs is observed and once the adult stage is reached, the number of KARs containing GluK1 subunits suffers a reduction in their expression in the granular layer, while the GluK2 and GluK5 remain constant, in contrast to AMPARs that increase their number, constituting a very notable majority compared to KARs.

All of these findings suggest that KARs have an important role in the development process of the cerebellum. Some indications suggest that GluK1-containing KARs participate in cerebellar development in the beginning of the differentiation of granular cells.

Additionally, KARs have been involved in some brain alterations in the cerebellum and a direct relationship exists between KA injection and cerebellar ataxia. Thus, the cerebellum is an important target to study functions of KARs and its possible role causing ataxia [72, 73, 74, 75]. Furthermore, in patients with schizophrenia, an increase in KARs containing GluR6 and K2 subunits is observed, which would mediate a reduction in GABAergic transmission [76, 77]. In neurodegeneration, KARs may have a role in calcification of the brain tissue as it has been found that local application of KA in some areas of the cerebellum produces changes in different ion levels, highly increasing Ca2+ levels for more than 8 weeks, which mediate calcification [78] (Table 1). KARs have been described as producing increases in intracellular calcium [79, 80] and seems to signal increasing intracellular calcium without putting the cell at risk due to excitotoxicity, due to its low conductance in contrast to AMPARs. Due to the lack of knowledge on the subject, further exploration is necessary to determine the KARs’ role in cerebellum development and cerebellar alterations.

2.1 KARs modulating glutamate release in the cerebellum: a biphasic effect

KARs are known to be expressed in the cerebellar cortex in the axons of cerebellar granule cells that form PF and make excitatory synapses with PuC [58]. Messenger RNA transcripts encoding for different KAR subunits and functional expression of KAR subtypes have been reported [81, 82, 83, 84]. Biophysical studies with single-channel recording have shown GluK1 activity [85], suggesting these KARs are Ca2+ permeable. A biphasic action of KARs, activated by the agonist domoate, has been shown previously at PF-PuC synapse, with low agonist concentrations, facilitating synaptic transmission and higher concentrations depressing synaptic transmission [86] in agreement with what has been found in the hippocampus [87, 88, 89], cortex [90], amygdala [91], and the thalamus [92]. EPSC trial-to-trial fluctuation analysis, failure rates, as well as paired-pulse ratios have shown that these facilitatory and depressive actions of KARs in the cerebellum are mediated by presynaptic KARs [80]. However, the precise mechanism of action by which KARs mediate potentiation (and depression) of synaptic transmission at PF-PuC synapses has remained elusive until very recently [80] (Table 1).

2.1.1 Action mechanism for KARs-mediated facilitation of glutamate release at cerebellar PF-PuC synapses

We have recently demonstrated that the effect of the KARs’ activation in this synapse requires protein kinase A (PKA) activation, since the inhibition of this protein by cAMP-Rp suppresses the effect of KA in glutamate release [80], in agreement with previous studies in hippocampus and cortex [87, 88, 89]. This congruence between mechanisms at different synapses has also been seen through the inhibition of PKA using H-89, which eliminates KARs-mediated facilitation of glutamate release. Similarly, the direct activation of AC (adenylyl cyclase) using forskolin caused an elimination of facilitation when KARs were activated by KA (with NMDARs and AMPARs blocked). These data indicate that a signaling mediated by AC/cAMP/PKA supports the facilitation of the modulation of synaptic transmission/glutamate release in these cerebellar synapses (Figures 13).

Figure 1.

KAR-mediated facilitation of glutamate release involving activation of adenylyl cyclase (AC) and downstream protein kinase A (PKA) at PF-PuC synapses of the cerebellum. (A) Time course of KA (3 μM) effect on eEPSCs amplitude in the absence (circles) and presence of NBQX (squares). Insets show traces before and after 4 min of KA perfusion in the absence (1, 2) and in the presence of 10 μM NBQX (1′, 2′). (B) Quantification of modulation observed in (A) and dose-response curve. (C) Time course of the effect of KA on eEPSC amplitude in cAMP-Rp-treated slices. (D) Inhibition of PKA by cAMP-Rp (100 μM) or H-89 (2 μM), and activation of AC by forskolin (30 μM) prevented the facilitatory action of KA. Inhibition of PKC with calphostin C (1 μM) has no effect on the KA enhancement of the eEPSC amplitude. The facilitatory effect of KA is not affected in slices treated with pertussis toxin. Modified from [64].

Figure 2.

Facilitation of glutamate release mediated by KAR activation requires an increase of Ca2+ in the cytosol and Ca2+ calmodulin at cerebellar PF-PuC synapses. (A) Time course of KA (3 μM) effect on eEPSCs amplitude in control condition (circles) and in slices treated with philanthotoxin (squares). (B) Quantification of modulation observed in (A). (C) Time course of KA (3 μM) effect on eEPSCs amplitude in control condition (circles) and in the slices treated with 25 μM W-7 (squares). (D) Quantification of modulation observed in (A) and in the presence of 1 μM CMZ. Modified from [64].

Figure 3.

KAR-mediated modulation of glutamate release in the cerebellum. Actions of KARs depressing or facilitating glutamate release at the PF-PuC synapse. KAR activation by high concentrations of KA (>3 μM) depresses glutamate release at PF-PuC synapses, an effect that involves Gi/o protein and the adenylate cyclase/cAMP/protein kinase A (AC/cAMP/PKA) pathway. KAR activation by low concentrations of kainate (<0.3 μM) only facilitates glutamate release following activation of a Ca2+-calmodulin/AC/cAMP/PKA pathway.

As observed in other synapses, Ca2+ seems to play a fundamental role in facilitating glutamate release at PF-PuC synapses. By blocking calcium-permeable KARs by the selective inhibitor philanthotoxin, KAR-mediated synaptic facilitation of glutamate release was prevented, indicating that there is a strict requirement for external Ca2+ entry through KARs to support the facilitation effect observed on glutamate release, indicating that KARs mediating the facilitation of glutamate release are calcium permeable [80].

Additionally, as has been reported at hippocampal synapses, the depletion of intracellular Ca2+ stores by a treatment with thapsigargin (a noncompetitive calcium inhibitor of the sarcoplasmic reticulum ATPase) eliminates the facilitation of glutamate release mediated by KARs’ activation. The same result was found when selectively inhibiting Ca2+-induced calcium release by using ryanodine, indicating that the entry of Ca2+ via KARs induces a mobilization of Ca2+ from the intraterminal Ca2+ reserves to mediate the increase in glutamate release observed [80].

Furthermore, it has been observed that the facilitation of glutamate release mediated by the activation of KARs is sensitive to calmodulin inhibitors. Previous studies showed that the increase of cytosolic calcium levels activates Ca2+ dependent on AC present in the terminals of parallel fibers. Through treatment with the calmodulin inhibitors, W-7 and calmidazolium, it has been recently shown [80] that the inhibition of calcium-calmodulin function prevents KAR-mediated presynaptic facilitation of glutamate release in cerebellar slices, supporting the hypothesis that after KAR activation and cytosolic elevation of Ca2+, a calmodulin-dependent calcium coupling activates AC, which subsequently activates the AC/cAMP/PKA pathway, thus promoting synaptic facilitation through an increase in neurotransmitter release at PF-PuC synapses [80].

2.1.2 Action mechanism for KARs-mediated depression of glutamate release at cerebellar PF-PuC synapses

Recently, in the same study discussed in the previous section [80], a transient synaptic depression of glutamate release with high concentrations of KA (3 μM) was observed as reported for other different brain areas including thalamus, cortex, hippocampus, and amygdala [89, 90, 91, 92]. This depression of glutamate release was prevented in the presence of cAMP-RP (which inhibits the activation of PKA), but was not affected by any other experimental modification discussed above with respect to the facilitation of glutamate release. This fact may indicate that the synaptic depression is probably related to an AC/cAMP/PKA signaling pathway (as for facilitation of glutamate release), but without the coupling of Ca2+ to the AC. Therefore, KA receptors have alternative mechanisms for facilitating and depressing glutamate release at PF-Pu synapses (Figures 13). In previous studies, investigating mossy fiber-CA3 hippocampal synapses [88, 89, 93, 94], as well as the amygdala [37] and cortex [36], a similar mechanism has been observed additionally involving the activation of a G-protein for the depressive effect that may well be also the case for these cerebellar synapses.

Although the presynaptic function of KARs facilitating glutamate release implies an increase in AC/cAMP/PKA signaling induced by the calcium calmodulin complex, KARs appear to be negatively associated with this pathway to carry out synaptic transmission of depression. Previous studies at hippocampal MF-CA3 synapses and thalamocortical synapses, as well as at PF-PuC synapses, have reported that the depression of glutamate release mediated by presynaptic KARs occurs through a negative coupling to the AC/cAMP/PKA pathway, being actually evoked by the action of a PTx-sensitive protein G [80, 88, 92]. Despite the hypotheses discussed above, it is also possible that these observed mechanisms reflect the presence of two different types of KARs, a clear objective being to clarify this hypothesis in future studies.

Advertisement

3. Conclusions

Regarding the role and mechanisms of KARs in the modulation of glutamate release in the cerebellum, new and recent data indicate that the KARs effecting facilitation of glutamate release and synaptic transmission show a mandatory dependence on adenylyl cyclase (AC) and cAMP-mediated protein kinase A (PKA) activity. Furthermore, the KAR-mediated facilitation of transmission is contingent on both external Ca2+ permeation into the cytosol through KAR and repletion of intracellular Ca2+ stores. Finally, a major sensitivity of facilitation to calmodulin inhibition suggests that KARs are coupled through a Ca2+-calmodulin/AC/cAMP/PKA pathway at PF-PuC synapses in the cerebellum. KARs seem to use the inhibition of the AC/cAMP/PKA pathway to mediate a depression of glutamate release at the same synapses, but the activation of the AC does not involve calcium calmodulin and seems to be directly activated by a PTX-sensitive G protein.

Advertisement

Acknowledgments

Work related to kainate receptors in the corresponding author AR-M’s laboratory has been supported by Spanish Ministry of Education and Fundación Rodríguez-Pascual grants.

Advertisement

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. 1. Lerma J, Paternain A, Rodríguez-Moreno A, López-García JC. Molecular physiology of kainate receptors. Physiological Reviews. 2001;81:971-998. DOI: 10.1152/physrev.2001.81.3.971
  2. 2. Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, Ogden KK, et al. Glutamate receptor ion channels: Structure, regulation, and function. Pharmacological Reviews. 2010;62:405-496. DOI: 10.1124/pr.109.002451
  3. 3. Flores G, Negrete-Díaz JV, Carrión M, Andrade-Talavera Y, Bello SA, Sihra TS, et al. Excitatory amino acids in neurological and neurodegenerative disorders. In: Amino Acids in Human Nutrition and Health. Wallingford: CAB International; 2011. pp. 427-453
  4. 4. Niswender CM, Conn PJ. Metabotropic glutamate receptors: Physiology, pharmacology, and disease. Annual Review of Pharmacology and Toxicology. 2010;50:295-322. DOI: 10.1146/annurev.pharmtox.011008.145533
  5. 5. Nadler JV, Perry BW, Cotman CW. Intraventricular kainic acid preferentially destroys hippocampal pyramidal cells. Nature. 1978;271:676-677. DOI: 10.1038/271676a0
  6. 6. Falcón-Moya R, Sihra TS, Rodríguez-Moreno A. Kainate receptors: Role in epilepsy. Frontiers in Molecular Neuroscience. 2018;11:217. DOI: 10.3389/fnmol.2018.00217
  7. 7. Lerma J, Marques JM. Kainate receptors in health and disease. Neuron. 2013;80(2):292-311. DOI: 10.1016/j.neuron.2013.09.045
  8. 8. Paternain AV, Rodríguez-Moreno A, Villarroel A, Lerma J. Activation and desensitization properties of native and recombinant kainate receptors. Neuropharmacology. 1998;37(10-11):1249-1259
  9. 9. Lee DL, editor. The Biology of the Nematodes. Boca Raton: CRC Press. Taylor and Francis Group; 2010
  10. 10. Li Y, Dharkar P, Han TH, Serpe M, Lee CH, Mayer ML. Novel functional properties of Drosophila CNS glutamate receptors. Neuron. 2016;92(5):1036-1048
  11. 11. Somogyi P, Eshhar N, Teichberg VI, Roberts JDB. Subcellular localization of a putative kainate receptor in Bergmann glial cells using a monoclonal antibody in the chick and fish cerebellar cortex. Neuroscience. 1990;35(1):9-30
  12. 12. Atoji Y, Sarkar S. Localization of AMPA, kainate, and NMDA receptor mRNAs in the pigeon cerebellum. Journal of Chemical Neuroanatomy. 2019;98:71-79
  13. 13. Estabel J, König N, Exbrayat JM. AMPA/kainate receptors permeable to divalent cations in amphibian central nervous system. Life Sciences. 1999;64(8):607-616
  14. 14. Huettner JE. Kainate receptors and synaptic transmission. Progress in Neurobiology. 2003;70(5):387-407. DOI: 10.1016/S0301-0082(03)00122-9
  15. 15. Jane DE, Lodge D, Collingridge GL. Kainate receptors: Pharmacology, function and therapeutic potential. Neuropharmacology. 2009;56(1):90-113. DOI: 10.1016/j.neuropharm.2008.08.023
  16. 16. Paternain AV, Herrera MT, Nieto MA, Lerma J. GluR5 and GluR6 kainate receptor subunits coexist in hippocampal neurons and coassemble to form functional receptors. Journal of Neuroscience. 2000;20(1):196-205. DOI: 10.1523/JNEUROSCI.20-01-00196.2000
  17. 17. Rodríguez-Moreno A, Sihra TS. Metabotropic actions of kainate receptors in the CNS. Journal of Neurochemistry. 2007;103(6):2121-2135. DOI: 10.1111/j.1471-4159.2007.04924.x
  18. 18. Rodríguez-Moreno A, Sihra TS. Kainate receptors with a metabotropic modus operandi. Trends in Neurosciences. 2007;30(12):630-637. DOI: 10.1016/j.tins.2007.10.001
  19. 19. Sihra TS, Rodríguez-Moreno A. Metabotropic actions of kainate receptors in the control of GABA release. In: Kainate Receptors. Boston, MA: Springer; 2011. pp. 1-10
  20. 20. Rodríguez-Moreno A, Sihra TS. Metabotropic actions of kainate receptors in the control of glutamate release in the hippocampus. In: Kainate Receptors. Boston, MA: Springer; 2011. pp. 39-48
  21. 21. Sihra TS, Rodríguez-Moreno A. Presynaptic kainate receptor-mediated bidirectional modulatory actions: Mechanisms. Neurochemistry International. 2013;62(7):982-987. DOI: 10.1016/j.neuint.2013.03.012
  22. 22. Negrete-Díaz JV, Sihra TS, Flores G, Rodríguez-Moreno A. Non-canonical mechanisms of presynaptic kainate receptors controlling glutamate release. Frontiers in Molecular Neuroscience. 2018;11:128. DOI: 10.3389/fnmol.2018.00128
  23. 23. Sihra TS, Flores G, Rodríguez-Moreno A. Kainate receptors: Multiple roles in neuronal plasticity. The Neuroscientist. 2014;20(1):29-43. DOI: 10.1177/1073858413478196
  24. 24. Rodríguez-Moreno A, Herreras O, Lerma J. Kainate receptors presynaptically downregulate GABAergic inhibition in the rat hippocampus. Neuron. 1997;19(4):893-901. DOI: 10.1016/S0896-6273(00)80970-8
  25. 25. Mulle C, Sailer A, Pérez-Otaño I, Dickinson-Anson H, Castillo PE, Bureau I, et al. Altered synaptic physiology and reduced susceptibility to kainite-induced seizures in GluR6-deficient mice. Nature. 1998;392:601-605
  26. 26. Smolders I, Bortolotto ZA, Clarke VR, Warre R, Khan GM, O’Neill MJ, et al. Antagonists of GLU(K5)-containing kainate receptors prevent pilocarpine-induced limbic seizures. Nature Neuroscience. 2002;5(8):796-804
  27. 27. Fritsch B, Reis J, Gasior M, Kaminski RM, Rogawski MA. Role of GluK1 kainate receptors in seizures, epileptic discharges, and epileptogenesis. The Journal of Neuroscience. 2014;34(17):5765-5775. DOI: 10.1523/JNEUROSCI.5307-13.2014
  28. 28. Rodríguez-Moreno A, Lerma J. Kainate receptor modulation of GABA release involves a metabotropic function. Neuron. 1998;20(6):1211-1218. DOI: 10.1016/S0896-6273(00)80501-2
  29. 29. Rodríguez-Moreno A, López-García JC, Lerma J. Two populations of kainate receptors with separate signaling mechanisms in hippocampal interneurons. Proceedings of the National Academy of Sciences. 2000;97(3):1293-1298. DOI: 10.1073/pnas.97.3.1293
  30. 30. Artinian J, Peret A, Marti G, Epsztein J, Crépel V. Synaptic kainate receptors in interplay with INaP shift the sparse firing of dentate granule cells to a sustained rhythmic mode in temporal lobe epilepsy. The Journal of Neuroscience. 2011;31(30):10811-10818. DOI: 10.1523/JNEUROSCI.0388-11.2011
  31. 31. Artinian J, Peret A, Mircheva Y, Marti G, Crépel V. Impaired neuronal operation through aberrant intrinsic plasticity in epilepsy. Annals of Neurology. 2015;77(4):592-606. DOI: 10.1002/ana.24348
  32. 32. Peret A, Christie LA, Ouedraogo DW, Gorlewicz A, Epsztein J, Mulle C, et al. Contribution of aberrant GluK2-containing kainate receptors to chronic seizures in temporal lobe epilepsy. Cell Reports. 2014;8(2):347-354. DOI: 10.1016/j.celrep.2014.06.032
  33. 33. Crépel V, Mulle C. Physiopathology of kainate receptors in epilepsy. Current Opinion in Pharmacology. 2015;20:83-88. DOI: 10.1016/j.coph.2014.11.012
  34. 34. Sander T, Hildmann T, Kretz R, Fürst R, Sailer U, Bauer G, et al. Allelic association of juvenile absence epilepsy with a GluR5 kainate receptor gene (GRIK1) polymorphism. American Journal of Medical Genetics. 1997;74:416-421
  35. 35. Li JM, Zeng YJ, Peng F, Li L, Yang TH, Hong Z, et al. Aberrant glutamate receptor 5 expression in temporal lobe epilepsy lesions. Brain Research. 2010;1311:166-174
  36. 36. Swanson GT. Targeting AMPA and kainate receptors in neurological disease: Therapies on the horizon? Neuropsychopharmacology. 2009;34:249-250
  37. 37. Sanchez-Gomez MV, Matute C. AMPA and kainate receptors each mediate excitotoxicity in oligodendroglial cultures. Neurobiology of Disease. 1999;6:475-485. DOI: 10.1006/nbdi.1999.0264
  38. 38. Matute C. Characteristics of acute and chronic kainate excitotoxic damage to the optic nerve. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:10229-10234. DOI: 10.1073/pnas.95.17.10229
  39. 39. Tekkok SB, Goldberg MP. AMPA/kainate receptor activation mediates hypoxic oligodendrocyte death and axonal injury in cerebral white matter. The Journal of Neuroscience. 2001;21:4237-4248. DOI: 10.1523/JNEUROSCI.21-12-04237.2001
  40. 40. Xu J, Liu Y, Zhang GY. Neuroprotection of GluR5-containing kainate receptor activation again ischemic brain injury through decreasing tyrosine phosphorylation of N-methyl-D-aspartate reeptors mediated by SRC kinase. The Journal of Biological Chemistry. 2008;283:29355-29366
  41. 41. O’Neill MJ, Bogaert L, Hicks CA, Bond A, Ward MA, Ebinger G, et al. LY377770, a novel iGlu5 kainate receptor antagonist with neuroprotective effects in global and focal cerebral ischaemia. Neuropharmacology. 2000;39:1575-1588. DOI: 10.1016/S0028-3908(99)00250-6
  42. 42. Aronica E, Dickson DW, Kress Y, Morrison JH, Zukin RS. Non-plaque dystrophic dendrites in Alzheimer hippocampus: A new pathological structure revealed by glutamate receptor immunocytochemistry. Neuroscience. 1998;82:979-991. DOI: 10.1016/S0306-4522(97)00260-1
  43. 43. Luquim MR, Saldise L, Guillén J, et al. Does increased excitatory drive from the subthalamic nucleus contribute to dopaminergic neuronal death in Parkinson’s disease? Experimental Neurology. 2006;201:407-415
  44. 44. Carson KM, Andresen JM, Orr HT. Emerging pathogenic pathways in the spinocerebellar ataxias. Current Opinion in Genetics and Development. 2009;19:247-253
  45. 45. Wagster MV, Hedreen JC, Peyser CE, Folstein SE, Ross CA. Selective loss of [3H] kainic acid and [3H] AMPA binding in layer VI of frontal cortex in Huntington’s disease. Experimental Neurology. 1994;127:70-75. DOI: 10.1006/exnr.1994.1081
  46. 46. Rubinsztein DC, Leggo J, Chiano M, Dodge A, Norbury G, Rosser E, et al. Genotypes at the GluR6 kainate receptor locus are associated with variation in the age of onset of Huntington disease. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:3872-3876. DOI: 10.1073/pnas.94.8.3872
  47. 47. MacDonald ME, Vonsattel JP, Shrinidhi J, Couropmitree NN, Cupples LA, Bird ED, et al. Evidence for the GluR6 gene associated with younger onset age of Huntington’s disease. Neurology. 1999;53:1330-1332. DOI: 10.1212/WNL.53.6.1330
  48. 48. Garey LJ, Von Bussmann KA, Hirsch SR. Decreasednumerical density of kainate receptor-positive neurons in the orbitofrontal cortex of chronic schizophrenics. Experimental Brain Research. 2006;173:234-242. DOI: 10.1007/s00221-006-0396-8
  49. 49. Begni S, Popoli M, Moraschi S, Bignotti S, Tura GB, Gennarelli M. Association between the ionotropic glutamate receptor kainate 3 (GRIK3) ser310ala polymorphism and schizophrenia. Molecular Psychiatry. 2002;7:416-418. DOI: 10.1038/sj.mp.4000987
  50. 50. Schiffer HH, Heinemann SF. Association of the human kainate receptor GluR7 gene (GRIK3) with recurrent major depressive disorder. American Journal of Medical Genetics. Part B, Neuropsychiatric Genetics. 2007;144:20-26. DOI: 10.1002/ajmg.b.30374
  51. 51. Pickard BS, Malloy MP, Christoforou A, et al. Cytogenetic and genetic evidence supports a role for the kainate-type glutamate receptor gene, GRIK4, in schizophrenia and bipolar disorder. Molecular Psychiatry. 2006;11:847-857. DOI: 10.1038/sj.mp.4001867
  52. 52. Wilson GM, Flibotte S, Chopra V, Melnyk BL, Honer WG, Holt RA. DNA copy number analysis in bipolar disorder and schizophrenia reveals aberrations in genes involved in glutamate signaling. Human Molecular Genetics. 2006;15:743-749. DOI: 10.1093/hmg/ddi489
  53. 53. Motazacker MM, Rost BR, Hucho T, Garshasbi M, Kahrizi K, Ullmann R, et al. A defect in the ionotropic glutamate receptor 6 gene (GRIK2) is associated with autosomal recessive mental retardation. American Journal of Human Genetics. 2007;81:792-798. DOI: 10.1086/521275
  54. 54. Jamain S, Betancur C, Mol Quach H, Philippe A, Fellous M, Giros B, et al. Linkage and association of the glutamate receptor 6 gene with autism. Molecular Psychiatry. 2002;7:302-310. DOI: 10.1038/sj.mp.4000979
  55. 55. Bowie D. Ionotropic glutamate receptors & CNS disorders. CNS & Neurological Disorders Drug Targets. 2008;7:129-143
  56. 56. Matute C. Therapeutic potential of kainate receptors. CNS Neuroscience and Therapeutics. 2011;17(6):661-669. DOI: 10.1111/j.1755-5949.2010.00204.x
  57. 57. Pemberton KE, Belcher SM, Ripellino JA, Howe JR. High-affinity kainate-type ion channels in rat cerebellar granule cells. The Journal of Physiology. 1998;510(2):401-420. DOI: 10.1111/j.1469-7793.1998.401bk.x
  58. 58. Smith TC, Wang LY, Howe JR. Distinct kainate receptor phenotypes in immature and mature mouse cerebellar granule cells. The Journal of Physiology. 1999;517(1):51-58. DOI: 10.1111/j.1469-7793.1999.0051z.x
  59. 59. Spiliopoulos K, Fragioudaki K, Giompres P, Kouvelas E, Mitsacos A. Expression of GluR6 kainate receptor subunit in granular layer of weaver mouse cerebellum. Journal of Neural Transmission. 2009;116(4):417-422. DOI: 10.1007/s00702-009-0199-8
  60. 60. Tyrrell T, Willshaw D. Cerebellar cortex: Its simulation and the relevance of Marr’s theory. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences. 1992;336(1277):239-257. DOI: 10.1098/rstb.1992.0059
  61. 61. Wadiche JI, Jahr CE. Multivesicular release at climbing fiber-Purkinje cell synapses. Neuron. 2001;32(2):301-313. DOI: 10.1016/S0896-6273(01)00488-3
  62. 62. Crépel F. Role of presynaptic kainate receptors at parallel-fiber-purkinje cell synapses in induction of cerebellar LTD: Interplay with climbing fiber input. Journal of Neurophysiology. 2009;102:965-973
  63. 63. Rodríguez-Moreno A, Paulsen O. Spike timing-dependent long-term depression requires presynaptic NMDA receptors. Nature Neuroscience. 2008;11:744-745
  64. 64. Banerjee A, Meredith RM, Rodríguez-Moreno A, Mierau SB, Auberson YP, Paulsen O. Double dissociation of spike timing-dependent potentiation and depression by subunit-preferring NMDA receptors antagonists in mouse barrel cortex. Cerebral Cortex. 2009;19:2959-2969
  65. 65. Rodríguez-Moreno A, Banerjee A, Paulsen O. Presynaptic NMDA receptors and spike timing-dependent depression at cortical synapses. Frontiers in Synaptic Neuroscience. 2010;2:18. DOI: 10.3389/fnsyn.2010.00018
  66. 66. Rodríguez-Moreno A, Kohl MM, Reeve J, Eaton TR, Collins HA, Anderson HL, et al. Presynaptic induction and expression of timing-dependent long-term depression demonstrated by compartment specific photorelease of a use-dependent NMDA antagonist. The Journal of Neuroscience. 2011;31:8564-8569
  67. 67. Buchanan KA, Blackman AV, Moreau AW, Elgar D, Costa RP, Lalanne T, et al. Target-specific expression of presynaptic NMDA receptors in neocortical microcircuits. Neuron. 2012;75:451-466
  68. 68. Rodríguez-Moreno A, González-Rueda A, Banerjee A, Upton ML, Craig M, Paulsen O. Presynaptic self-depression at developing neocortical synapses. Neuron. 2013;77:35-42
  69. 69. Banerjee A, González-Rueda A, Sampaio-Baptista C, Paulse 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. Physiological Reports. 2014;2(3):1-11. DOI: 10.1002/phy2.271
  70. 70. Komuro H, Rakic P. Modulation of neuronal migration by NMDA receptors. Science. 1993;260(5104):95-97. DOI: 10.1126/science.8096653
  71. 71. Rabacchi S, Bailly Y, Delhaye-Bouchaud N, Mariani J. Involvement of the N-methyl d-aspartate (NMDA) receptor in synapse elimination during cerebellar development. Science. 1992;256(5065):1823-1825. DOI: 10.1126/science.1352066
  72. 72. Maiti A, Salles KS, Grassi S, Abood LG. Behavior and receptor changes after kainate lesioning of nodular cerebellum. Pharmacology Biochemistry and Behavior. 1986;25(3):589-594. DOI: 10.1016/0091-3057(86)90146-2
  73. 73. de Vera N, Camón L, Martínez E. Cerebral distribution of polyamines in kainic acid-induced models of status epilepticus and ataxia in rats. Overproduction of putrescine and histological damage. European Neuropsychopharmacology. 2002;12(5):397-405. DOI: 10.1016/S0924-977X(02)00050-0
  74. 74. Yamaguchi T, Hayashi K, Murakami H, Maruyama S, Yamaguchi M. Distribution and characterization of the glutamate receptors in the CNS of ataxic mutant mouse. Neurochemical Research. 1984;9(4):497-505. DOI: 10.1007/BF00964376
  75. 75. Andoh T, Kishi H, Motoki K, Nakanishi K, Kuraishi Y, Muraguchi A. Protective effect of IL-18 on kainate-and IL-1 β-induced cerebellar ataxia in mice. Journal of Immunology. 2008;180:2322-2328. DOI: 10.4049/jimmunol.180.4.2322
  76. 76. Harrison PJ, Barton AJ, Najlerahim A, Pearson RC. Distribution of a kainate/AMPA receptor mRNA in normal and Alzheimer brain. Neuroreport. 1990;1(2):149-152
  77. 77. Bullock WM, Cardon K, Bustillo J, Roberts RC, Perrone-Bizzozero NI. Altered expression of genes involved in GABAergic transmission and neuromodulation of granule cell activity in the cerebellum of schizophrenia patients. American Journal of Psychiatry. 2008;165(12):1594-1603. DOI: 10.1176/appi.ajp.2008.07121845
  78. 78. Korf J, Postema F. Regional calcium accumulation and cation shifts in rat brain by kainate. Journal of Neurochemistry. 1984;43(4):1052-1060. DOI: 10.1111/j.1471-4159.1984.tb12843.x
  79. 79. Savidge JR, Bleakman D, Bristow DR. Identification of kainate receptor-mediated intracellular calcium increases in cultured rat cerebellar granule cells. Journal of Neurochemistry. 1997;69(4):1763-1766. DOI: 10.1046/j.1471-4159.1997.69041763.x
  80. 80. Falcón-Moya R, Losada-Ruiz P, Sihra TS, Rodríguez-Moreno A. Cerebellar Kainate receptor-mediated facilitation of glutamate release requires Ca2+−calmodulin and PKA. Frontiers in Molecular Neuroscience. 2018;11:1-10. DOI: 10.3389/fnmol.2018.00195
  81. 81. Bahn S, Volk B, Wisden W. Kainate receptor gene expression in the developing rat brain. Journal of Neuroscience. 1994;14(9):5525-5547. DOI: 10.1523/JNEUROSCI.14-09-05525.1994
  82. 82. Bettler B, Boulter J, Hermans-Borgmeyer I, O’Shea-Greenfield A, Deneris ES, Moll C, et al. Cloning of a novel glutamate receptor subunit, GluR5: Expression in the nervous system during development. Neuron. 1990;5(5):583-595. DOI: 10.1016/0896-6273(90)90213-y
  83. 83. Herb A, Burnashev N, Werner P, Sakmann B, Wisden W, Seeburg PH. The KA-2 subunit of excitatory amino acid receptors shows widespread expression in brain and forms ion channels with distantly related subunits. Neuron. 1992;8(4):775-785. DOI: 10.1016/0896-6273(92)90098-x
  84. 84. Petralia RS, Wang YX, Wenthold RJ. Histological and ultrastructural localization of the kainate receptor subunits, KA2 and GluR6/7, in the rat nervous system using selective antipeptide antibodies. Journal of Comparative Neurology. 1994;349(1):85-110. DOI: 10.1002/cne.903490107
  85. 85. Swanson GT, Feldmeyer D, Kaneda M, Cull-Candy SG. Effect of RNA editing and subunit co-assembly single-channel properties of recombinant kainate receptors. Journal of Physiology. 1996;492:129-142
  86. 86. Delaney AJ, Jahr CE. Kainate receptors differentially regulate release at two parallel fiber synapses. Neuron. 2002;36(3):475-482. DOI: 10.1016/s08966273(02)01008-5
  87. 87. Rodríguez-Moreno A, Sihra TS. Presynaptic kainate receptor facilitation of glutamate release involves protein kinase A in the rat hippocampus. Journal of Physiology. 2004;557(3):733-745. DOI: 10.1113/jphysiol.2004.065029
  88. 88. Negrete-Díaz JV, Sihra TS, Delgado-García JM, Rodríguez-Moreno A. Kainate receptor–mediated inhibition of glutamate release involves protein kinase A in the mouse hippocampus. Journal of Neurophysiology. 2006;96(4):1829-1837. DOI: 10.1152/jn.00280.2006
  89. 89. Andrade-Talavera Y, Duque-Feria P, Negrete-Díaz JV, Sihra TS, Flores G, Rodríguez-Moreno A. Presynaptic kainate receptor-mediated facilitation of glutamate release involves Ca2+-calmodulin at mossy fiber-CA3 synapses. Journal of Neurochemistry. 2012;122(5):891-899. DOI: 10.1111/j.1471-4159.2012.07844.x
  90. 90. Rodríguez-Moreno A, Sihra TS. Presynaptic kainate receptor-mediated facilitation of glutamate release involves Ca2+–calmodulin and PKA in cerebrocortical synaptosomes. FEBS Letters. 2013;587(6):788-792. DOI: 10.1016/j.febslet.2013.01.071
  91. 91. Negrete-Díaz JV, Duque-Feria P, Andrade-Talavera Y, Carrión M, Flores G, Rodríguez-Moreno A. Kainate receptor-mediated depression of glutamatergic transmission involving protein kinase A in the lateral amygdala. Journal of Neurochemistry. 2012;121(1):36-43. DOI: 10.1111/j.1471-4159.2012.07665.x
  92. 92. Andrade-Talavera Y, Duque-Feria P, Sihra TS, Rodríguez-Moreno A. Pre-synaptic kainate receptor-mediated facilitation of glutamate release involves PKA and Ca2+-calmodulin at thalamocortical synapses. Journal of Neurochemistry. 2013;126(5):565-578. DOI: 10.1016/j.febslet.2013.01.071
  93. 93. Negrete-Díaz JV, Sihra TS, Delgado-García JM, 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. Journal of Neural Transmission. 2007;114(11):1425-1431. DOI: 10.1007/s00702-007-0750-4
  94. 94. Lyon L, Borel M, Carrión M, Kew JN, Corti C, Harrison PJ, et al. Hippocampal mossy fiber long-term depression in Grm2/3 double knockout mice. Synapse. 2011;65(9):945-954. DOI: 10.1002/syn.20923

Written By

Pilar Losada-Ruiz, Rafael Falcón-Moya and Antonio Rodríguez-Moreno

Submitted: 19 March 2019 Reviewed: 12 June 2019 Published: 24 September 2019