当前位置: 首页 > 期刊 > 《生理学报》 > 2005年第1期
编号:11331993
Two developmental switches in GABAergic signalling: the K+–Cl– cotransporter KCC2 and carbonic anhydrase CAVII
http://www.100md.com 《生理学报》 2005年第1期
     1 Institute of Biotechnology

    2 Department of Biological and Environmental Sciences

    3 Neuroscience Center, University of Helsinki, FIN-00014 Helsinki, Finland

    Abstract

    GABAergic signalling has the unique property of ‘ionic plasticity’, which is based on short-term and long-term changes in the Cl– and HCO3– ion concentrations in the postsynaptic neurones. While short-term ionic plasticity is caused by activity-dependent, channel-mediated anion shifts, long-term ionic plasticity depends on changes in the expression patterns and kinetic regulation of molecules involved in anion homeostasis. During development the efficacy and also the qualitative nature (depolarization/excitation versus hyperpolarization/inhibition) of GABAergic transmission is influenced by the neuronal expression of two key molecules: the chloride-extruding K+–Cl– cotransporter KCC2, and the cytosolic carbonic anhydrase (CA) isoform CAVII. In rat hippocampal pyramidal neurones, a steep up-regulation of KCC2 accounts for the ‘developmental switch’, which converts depolarizing and excitatory GABA responses of immature neurones to classical hyperpolarizing inhibition by the end of the second postnatal week. The immature hippocampus generates large-scale network activity, which is abolished in parallel by the up-regulation of KCC2 and the consequent increase in the efficacy of neuronal Cl– extrusion. At around postnatal day 12 (P12), an abrupt, steep increase in intrapyramidal CAVII expression takes place, promoting excitatory responses evoked by intense GABAergic activity. This is largely caused by a GABAergic potassium transient resulting in spatially widespread neuronal depolarization and synchronous spike discharges. These facts point to CAVII as a putative target of CA inhibitors that are used as antiepileptic drugs. KCC2 expression in adult rat neurones is down-regulated following epileptiform activity and/or neuronal damage by BDNF/TrkB signalling. The lifetime of membrane-associated KCC2 is very short, in the range of tens of minutes, which makes KCC2 ideally suited for mediating GABAergic ionic plasticity. In addition, factors influencing the trafficking and kinetic modulation of KCC2 as well as activation/deactivation of CAVII are obvious candidates in the ionic modulation of GABAergic responses. The down-regulation of KCC2 under pathophysiological conditions (epilepsy, damage) in mature neurones seems to reflect a ‘recapitulation’ of early developmental mechanisms, which may be a prerequisite for the re-establishment of connectivity in damaged brain tissue.
, http://www.100md.com
    Introduction

    A characteristic feature of all structures in the developing central nervous system, including the spinal cord, sensory systems, brain stem as well as the cortex, is the presence of endogenous large-scale spontaneous activity (Feller, 1999; Penn & Shatz, 1999; Zhang & Poo, 2001; Ben-Ari, 2002). The temporal patterns as well as cellular and network mechanisms related to early network events seem to exhibit considerable variations among distinct neuronal networks as well as during the ontogeny of a given structure. Nevertheless, it is generally believed that this type of endogenous activity has an important role in the activity-dependent wiring of neuronal circuits, and that the maturation of GABAergic inhibition is a crucial factor during later developmental stages when the large-scale endogenous events disappear (Garaschuk et al. 2000; Ben-Ari, 2001; Owens & Kriegstein, 2002; Kandler, 2004; Sernagor et al. 2003).
, 百拇医药
    The aim of the present review is to summarize recent data and conclusions related to the role of gamma-aminobutyric acid (GABA) in early network activity of mammalian cortical structures. Most of this work has been carried out on the rodent hippocampus, where the developmental change in the action of GABA from a depolarizing (and often excitatory) transmitter in immature neurones to a hyperpolarizing and typically inhibitory one has gained an enormous amount of attention. The key mechanisms in such a change in transmitter function must be postsynaptic. In the case of GABA and its sister transmitter glycine, these mechanisms are based on the maturation of neuronal Cl– homeostasis, which produces a negative shift in the equilibrium potential of Cl– during neuronal development and differentiation (Mueller et al. 1984; Ben-Ari et al. 1989; Luhmann & Prince, 1991; Zhang et al. 1991). This early ‘ontogenetic switch’ in ionotropic GABAergic transmission is attributable to the developmental expression of the neurone-specific K+–Cl– cotransporter, KCC2 (Fig. 1A; Rivera et al. 1999; Hubner et al. 2001). In addition to this, there is another molecular switch, the neuronal expression of the carbonic anhydrase isoform VII (CAVII), which can transiently make GABAergic transmission functionally excitatory in mature neurones, especially under conditions of massive activation of GABAA receptors (Ruusuvuori et al. 2004).
, http://www.100md.com
    A, the current across GABAA receptors (GABAAR) is carried by Cl– and HCO3–. While the Cl–-mediated current can be depolarizing or hyperpolarizing, depending on the developmental expression of KCC2, the HCO3– current is depolarizing in all neurones irrespective of developmental stage. In order to have a significant effect, the HCO3– current has to be supported by the conversion of CO2 into HCO3– (and protons) by an intraneuronal carbonic anhydrase, such as the isoform CAVII that is present in cortical principal neurones (see text for further details). B, a scheme showing the developmental patterns of expression of KCC2 and CAVII in rat hippocampal pyramidal neurones. The first main ontogenetic switch from depolarizing to hyperpolarizing GABAA action is caused by the developmental expression of KCC2 that shows a steep increase during the first two postnatal weeks. A sharp increase in the expression of CAVII that takes place at around P12 promotes transient excitatory GABAergic responses which are caused by the GABAA-mediated HCO3– current and, in brain tissue, by the consequent increase in interstitial K+ (see text for detailed explanation).
, 百拇医药
    Neuronal chloride homeostasis

    Adult mammalian central neurones, with some exceptions (Gulacsi et al. 2003; Bartho et al. 2004), are atypical cells in that they maintain a low intracellular Cl– concentration which is, of course, a prerequisite for ‘classical’ hyperpolarizing inhibition mediated by ionotropic GABA and glycine receptors (Kaila, 1994). The evolutionary trade-off for this mode of postsynaptic signalling must have included, among other consequences, a compromised capacity for (and/or a requirement for novel mechanistic designs for) the control of intracellular pH and cellular volume. In most cells, both of these homeostatic functions are largely dependent on a large source of intracellular Cl– that is usually required for the uptake of HCO3– (in the case of pH regulation) or for net efflux of K+ in response to cell swelling (Alvarez-Leefmans & Russell, 1990; Kaila & Ransom, 1998). Hence, the high intracellular Cl– of immature neurones can be considered the rule, rather than the exception, in comparative cellular physiology.
, 百拇医药
    Cl– homeostasis in brain cells is mainly controlled by cation–chloride cotransporters (CCCs), which mediate electrically neutral Cl– uptake fuelled by Na+ (the Na+–K+–2 Cl– cotransporter, NKCC1) or Cl– extrusion fuelled by K+ (the K+–Cl– cotransporters, KCC1–4) (Hiki et al. 1999; Race et al. 1999; Delpire & Mount, 2002; Payne et al. 2003; Mercado et al. 2004). These secondary active transporters do not directly consume ATP, but they derive their energy from the Na+ and K+ gradients generated by the Na+,K+-ATPase. In addition to the CCCs, the Na+-dependent and Na+-independent anion exchangers (NDAE and AE, respectively) have an influence on neuronal Cl– since they mediate an exchange of Cl– for HCO3–. For the present purposes, we will focus on the expression patterns and functions of one member of the KCCs and of NKCCs: these are KCC2 and NKCC1.
, http://www.100md.com
    It is widely believed that uptake of Cl– in immature neurones is mediated by NKCC1 (Payne et al. 2003; Yamada et al. 2004). Interestingly, although GABAA-mediated depolarization is not present in NKCC1 knock-out mice (Sung et al. 2000), no significant phenotype has been observed in the central nervous system.

    Unfortunately, there are no drugs available that would act as specific inhibitors of K+–Cl– cotransport (not to mention an isoform-specific inhibitor!). Two compounds, furosemide and DIOA (R(+)-[(2-n-butyl-6,7-dichloro-2-cyclopentyl-2,3-dihydro-1-oxo-1H-inden-5-yl)oxy] acetic acid), are widely used, but they have effects on a wide variety of enzymes and receptors, which means that careful control experiments must be run to confirm that they are working in a manner that is consistent with any given experimental strategy. In particular, it should be emphasized that both of these drugs are more potent blockers of NKCC1 than of KCC2 (Payne, 1997), and they also inhibit carbonic anhydrases (Supuran et al. 2003; and unpublished observations by K. Kaila, J. Kolehmainen, C.T. Supuran & J. Voipio). Another complication that is often ignored in electrophysiological experiments is that Cs+, a widely used internal cation in whole-cell patch clamp experiments, is a very poor substrate for KCC2, resulting in a block of Cl– extrusion (Kakazu et al. 2000; Williams & Payne, 2004). Unfortunately, the kinetic data in the only published model of neuronal K+–Cl– cotransport in neurones (Staley & Proctor, 1999) were obtained using Cs+ in whole-cell experiments – and hence, in the absence of functional KCC2!
, http://www.100md.com
    Developmental up-regulation of KCC2

    Pioneering work (Deisz & Lux, 1982; Misgeld et al. 1986; Thompson et al. 1988a,b; Thompson & Ghwiler, 1989; Jarolimek et al. 1999) showed that neuronal Cl– extrusion is largely attributable to K+–Cl– cotransport, and that there was a developmentally regulated increase in the efficacy of Cl– extrusion during neuronal maturation (Luhmann & Prince, 1991; Zhang et al. 1991). The molecular identity of the extrusion mechanism became clear in studies on the expression patterns of KCC2 in the immature rat hippocampus (Fig. 2A). The well-known negative shift in the reversal potential of ionotropic GABA action (EGABA) was paralleled by the up-regulation of this molecule as seen in in situ hybridization (Rivera et al. 1999). (A similar result was published by Lu et al. (1999) but, surprisingly, these workers also found KCC2 expression in dorsal root ganglion neurones, which casts doubts on the specificity of their assays.) In addition, KCC2 shows a brain-specific expression pattern (Payne et al. 1996) and the neurone specificity was further confirmed in subsequent work (Rivera et al. 1999; Williams et al. 1999; DeFazio et al. 2000; Hubner et al. 2001; Stein et al. 2004). Obviously, a correlation between the developmental expression patterns of KCC2 and the negative shift of EGABA does not demonstrate a causal relationship. Direct evidence for this was obtained using gene knock-down of KCC2 in slice cultures based on the use of antisense oligonucleotides. This resulted in a block of KCC2 expression and to near abolition of the negative (hyperpolarizing) driving force of the GABAA-mediated current (Rivera et al. 1999). It is likely that NDAE is capable of mediating Cl– extrusion in neurones devoid of KCC2 (see also Kaila, 1994; Gulacsi et al. 2003), which explains the small residual hyperpolarizing GABAergic action. A full KCC2 ‘knock-out’ is lethal, and mice with such a gene disruption die at birth due to the expected malfunction of the respiratory centre (Hubner et al. 2001).
, http://www.100md.com
    A, radioactive in situ hybridization (panels b, d and e) showing steep increase in KCC2 mRNA by the end of the second postnatal week in rat hippocampal principal neurones (eosin haematoxylin counter staining, panels a and c; Th, thalamus; Cx, neocortex; DG, dentate gyrus). B, fluorescent in situ hybridizations indicating a pronounced increase in the expression of CAVII mRNA at around P12. Note the absence of a marked change in the level of the CAII transcript. (A is modified from Rivera et al. 1999 ( Nature Publishing Group (1999); http://www.nature.com/), B from Ruusuvuori et al. 2004 ( 2004 by the Society for Neuroscience)).
, http://www.100md.com
    The developmental patterns of KCC2 expression suggest that it is a molecule that is a very useful indicator of the state of neuronal maturation (see also Shimizu-Okabe et al. 2002; Ikeda et al. 2003; Payne et al. 2003). Here, an interesting finding is that a minority of central neurones do not express KCC2. These include the dopaminergic neurones in the striatum (Gulacsi et al. 2003) as well as a population of neurones in the nucleus reticularis thalami (Bartho et al. 2004). With regard to the dopaminergic neurones, one might postulate that the weak GABAergic inhibition they experience is important for tonic spiking and the consequent secretion of dopamine. Perhaps a lack of KCC2 is also a distinct feature of other neurones that secrete neuromodulatory amines – a question that needs attention in future work.
, http://www.100md.com
    In the developing rat hippocampus, the increase in the efficacy of Cl– extrusion and expression of KCC2 take place by the end of the second postnatal week (Zhang et al. 1991; Rivera et al. 1999). However, one should be careful in comparisons across species, keeping in mind that the time of birth provides no solid point of reference. For instance, the rat is born at a very immature state which makes early postnatal pups useful model organisms in studies of human cortical development, corresponding roughly to the last trimester of pregnancy (cf. Clancy et al. 2001; Avishai-Eliner et al. 2002; Erecinska et al. 2004). Furthermore, the developmental time course of KCC2 expression is neurone-type specific within a given species as seen in the earlier expression of this transporter in, e.g. thalamus than in the cortex (Rivera et al. 1999; Li et al. 2002; Wang et al. 2002). It is also intriguing that KCC2 is highly expressed in dendritic spines (Gulyas et al. 2001), which are major postsynaptic targets of glutamatergic rather than GABAergic inputs (Freund & Buzsaki, 1996). This raises the question whether KCC2 has functions in neuronal signalling which are not directly related to the maintenance of the postsynaptic Cl– gradient underlying ‘classical’ hyperpolarizing inhibition (cf. Kaila, 1994).
, 百拇医药
    What triggers the expression of KCC2 during neuronal development In a recent paper based on hippocampal cultures, Ganguly et al. (2001) suggested that GABA itself is the signal that activates the intracellular cascades controlling KCC2 gene expression. This hypothesis, where the depolarizing action of GABA is shut off in a feed-back manner is an elegant one, but subsequent work has shown that the increase in KCC2 protein expression (Ludwig et al. 2003) as well as the negative shift in EGABA (Titz et al. 2003) can take place in cultures in the constant presence of GABAA antagonists. Factors acting on the tyrosine kinase receptor such as insulin-like growth factor and BDNF have been shown to regulate the expression and function of KCC2 during development (Kelsch et al. 2001; Aguado et al. 2003). In addition, recent work comparing the quantitative expression of KCC2 protein and the efficacy (functionality) of KCC2 in native hippocampal tissue versus cultures suggests the absence of some critical factor(s) in cultures that are needed for the KCC2 transporter to become functional (unpublished observations of S. Khirug, K. Huttu, A. Ludwig, S. Smirnov, J. Voipio, C. Rivera, K. Kaila & L. Khiroug).
, 百拇医药
    Early spontaneous activity and KCC2

    In a highly cited paper published by Ben-Ari and coworkers (Ben-Ari et al. 1989), the shift from depolarizing to hyperpolarizing GABA action in the rat hippocampus was found to take place in parallel with the disappearance of large-scale network events, which were termed ‘giant depolarizing potentials’ (GDPs) in recordings with intracellular sharp microelectrodes. The developmental time scale of these sequences of events at the cellular and network level has been subject to recent revisions (Khazipov et al. 2004). Nevertheless, the intriguing temporal correlation with early endogenous activity and the depolarizing action of GABA has led to the hypothesis that the generation of endogenous patterns of activity in developing networks is set by the action of depolarizing GABA and, hence, by the properties of the interneuronal network.
, 百拇医药
    When examining the literature on developing networks as a whole, it appears that transmitter actions other than, or in addition to, depolarizing GABA are often involved in driving the endogenous activity that is characteristic of immature CNS structures (see Introduction). However, depolarizing GABA seems to play a role during certain developmental time windows in various brain circuits, and in cortical structures the present evidence does suggest an important role for GABA-mediated excitation (Ben-Ari, 2002; Leinekugel et al. 2002; Owens & Kriegstein, 2002). Nevertheless, even in the immature hippocampus it is not clear at all that GABAergic interneurones are responsible for the rhythmogenesis of early spontaneous activity. Recently, we have obtained strong evidence for the hypothesis that the temporal patterns of hippocampal ‘GDPs’ and their counterparts at the network level are shaped by glutamatergic neurones (S. Sipil, K. Huttu, J. Voipio & K. Kaila, unpublished observations).
, http://www.100md.com
    Neuronal damage: ‘recapitulation’ of KCC2 expression patterns and network activity

    The fact that neuronal trauma leads to a de-differentiation of the neuronal phenotype is known from a number of studies, including those on ion transport (Nabekura et al. 2002; Toyoda et al. 2003; Payne et al. 2003). For instance, it is known that changes occur in the GABAA subunit composition in the dentate gyrus that resemble those during early development (Buhl et al. 1996).
, http://www.100md.com
    It is a well-known fact that epileptic activity leads to an increase in the expression of brain-derived neurotrophic factor (BDNF) and its receptor TrkB (Binder et al. 2001; Huang & Reichardt, 2001, 2003). We were intrigued by the observation that following in vivo kindling, the expression of KCC2 showed a rapid, pronounced fall in those regions of the epileptic hippocampus where BNDF–TrkB up-regulation is most salient (Rivera et al. 2002). Further in vitro experiments established a direct causal link from TrkB activation to KCC2 down-regulation (Rivera et al. 2002, 2004). Using transgenic mice with point mutations in their TrkB receptors (see Minichiello et al. 1998), it was found that the down-regulation of KCC2 requires the activation of the two major TrkB-mediated signalling cascades, the PLC and Shc activated pathways. In contrast, activation of the Shc cascade in isolation leads, intriguingly, to an increase in KCC2 expression (Rivera et al. 2004). Hence, these data may shed light also on the involvement of trophic signalling in the developmental up-regulation of KCC2 (see above).
, 百拇医药
    The above observations raise a number of questions and hypotheses: perhaps the fall in KCC2 is part of a wider picture where the afflicted neurones assume neonatal features that are crucial for the de novo targeting of neurones during damage-related rewiring Perhaps a down-regulation of ion extrusion is an energy-saving strategy that will help the neurones to survive a time window where hyperexcitability leads to an enormous energy metabolic load. It is obvious that questions of this kind cannot be resolved just by a series of experiments – they are paradigmatic in nature, and hence require a large bulk of experimental data from future work to provide solid evidence for, or against, definitive hypotheses. However, an intriguing observation in this context is that in the chronically epileptic human hippocampus, the sclerosis of the CA1 area is associated with an enhanced subicular output. In particular, some of the principal neurones in the subiculum have depolarizing and excitatory GABA responses (Cohen et al. 2002, 2003). Perhaps these are indications of a ‘recapitulation’ of a developmental programme (Cohen et al. 2003; Payne et al. 2003), which is necessary for repair mechanisms that operate at the network level. An exciting possibility is that the interictal activity which is a characteristic feature of the epileptic brain shares some fundamental features with the early large-scale spontaneous activity that is implicated in the formation of neuronal networks.
, http://www.100md.com
    In addition to down-regulation, KCC2 expression can be enhanced within a certain time window in response to neuronal hyperexcitability and/or trauma. Here, caution is needed in functional interpretations, since a dramatic enhancement of intrasomatic staining is seen in KCC2 immunohistochemistry (based on a well-characterized, specific antibody that binds to the cytoplasmic C-terminus of this transport protein; Ludwig et al. 2003) in a number of in vivo and in vitro models of neuronal hyperexcitability and damage (Lahtinen et al. 2002, 2004). In order to find out whether an adaptive increase in the expression of functional KCC2 takes place in response to trauma, direct measurements of the capacity of neuronal Cl– extrusion are needed.
, 百拇医药
    Carbonic anhydrase isoform VII (CAVII) and bicarbonate-dependent GABAergic excitation

    Massive, synchronous excitation of hippocampal pyramidal neurones can be achieved by high-frequency stimulation (HFS) of hippocampal interneurones in the absence of ionotropic glutamatergic transmission (Kaila et al. 1997; Smirnov et al. 1999; Ruusuvuori et al. 2004). The ionic mechanism of this type of GABAergic excitation is strictly dependent on the presence of HCO3–, which is the only physiological ion in addition to Cl– that is able to mediate a current across GABAA receptors.
, 百拇医药
    The equilibrium potential of HCO3– (EHCO3) is at a much more positive level (around –10 mV) than that of Cl–, since EHCO3 is set by mechanisms that control intracellular pH regulation (Kaila & Voipio, 1987). Hence, bicarbonate mediates an inward, depolarizing current (Kaila & Voipio, 1987; Kaila et al. 1993; Gulledge & Stuart, 2003). During the GABAA receptor-mediated efflux of HCO3–, the intracellular level of this anion is strongly buffered by the presence of intraneuronal carbonic anhydrase (CA) (Pasternack et al. 1993) that uses CO2 as a substrate for the generation of HCO3– (Kaila et al. 1990). As shown in crayfish muscle fibres (Kaila & Voipio, 1987; Kaila et al. 1989) as well as neurones (Voipio et al. 1991), the depolarization caused by the HCO3– current leads to an accumulation of intraneuronal Cl– and consequently, there is a large positive shift in EGABA.
, http://www.100md.com
    Notably, in mammalian brain tissue, the HCO3–- and CA-dependent anion redistribution induces a profound increase in interstitial K+ during prolonged GABAergic stimulation (also during GABAA agonist application in the absence of spiking; Barolet & Morris, 1991), which results in a long-lasting, non-synaptic depolarization. But why should GABAA receptor activation produce a large potassium transient Here one should note that the only biophysically feasible way to clear the HCO3–-dependent postsynaptic intraneuronal Cl– load is to link the subsequent, homeostatic net efflux of Cl– in a 1 : 1 manner to an efflux of a K+ (Voipio & Kaila, 2000). Whether this coflux is mediated by channels and/or transporters is not known. Nevertheless, the increase in extracellular K+ is an ionic transient that is exactly what is expected to take place in response to massive GABAA receptor activation. In accordance with this, the membrane potential of pyramidal neurones is influenced by the increase in K+, and hence it achieves a level that is much more depolarized than the value of EGABA (Kaila et al. 1997; Smirnov et al. 1999). As might be expected, membrane-permeant CA inhibitors strongly suppress HFS-induced GABAergic excitation, while impermeant CA inhibitors (such as benzolamide) have no effect (Ruusuvuori et al. 2004).
, 百拇医药
    When examining the ontogeny of the HCO3–-dependent GABAergic excitation, we found a sharp expression profile where HFS of GABAergic inputs does not lead to synchronous spiking in the early postnatal hippocampus (Figs 1B and 3; Ruusuvuori et al. 2004). The developmental profiles of action of CA inhibitors on HFS-induced synchronous spiking as well as fluorescence measurements of intracellular pH transients indicated that the expression of intrapyramidal CA, i.e. the ‘developmental switch’ that is necessary for HFS-induced GABAergic excitation, commences at P10–P12.
, http://www.100md.com
    Note emergence of abrupt synchronous spiking evoked by HFS at P12 and the selective sensitivity to the membrane-permeant CA inhibitor ethoxyzolamide (EZA) but not to the impermeant derivative benzolamide (BA). (From Ruusuvuori et al. 2004, 2004 by the Society for Neuroscience).

    From the 14 CA isozymes that are at present known, five show a cytosolic localization (CAs I–III, CAVII and CAXIII; for references, see Supuran et al. 2003; Pastorekova et al. 2004; see also Lakkis et al. 1997). Using in situ hybridization, the intrapyramidal carbonic anhydrase was identified as CAVII (Fig. 2B; Ruusuvuori et al. 2004; see also Lakkis et al. 1997). Considering (1) the role of CAVII in GABAergic excitation and (2) the tight link between epileptiform actitity and interstitial K+ transients, it is intriguing to speculate that CAVII might be a major target of those anticonvulsants that are known to be inhibitors of carbonic anhydrase (Wyllie, 1997; Masereel et al. 2002; Scozzafava et al. 2004). It is of particular interest that CAVII has an extremely high catalytic activity, rivalling the well-known high activity of CAII (Earnhardt et al. 1998). Obviously, CAVII is a promising target for the design of novel anticonvulsant compounds. It is also likely that the catalytic action of CAVII is modulated by a variety of compounds such as monoamines and tricarboxylic acids (Supuran et al. 2003), which suggests that these endogenous modulators would also shape the properties of GABAergic transmission activation/deactivation of intraneuronal CA activity.
, 百拇医药
    Ionic plasticity of GABAergic inhibition

    It is apparent that GABAergic signalling has the unique property of ‘ionic plasticity’ which is based on both short-term and long-term shifts in the concentrations of Cl– and HCO3– in postsynaptic neurones. Short-term ionic plasticity manifests itself not only as a change in the efficacy of GABAergic inhibition, but sometimes as a change from functional inhibition to GABAergic excitation that takes place in response to massive GABAA receptor activation during high-frequency activity of interneuronal networks. This suggests that the qualitative change of action of GABAergic transmission from inhibitory to excitatory is likely to take place during, e.g. epileptiform activity as well as during standard protocols of LTP induction that involve bursts of tetanic stimulation. An obvious task in future work is to examine the role of CAVII in the induction of LTP as well as in epileptogenesis (see also Sun & Alkon, 2002).
, http://www.100md.com
    While short-term ionic plasticity of GABAergic transmission is directly shaped by activity-dependent ion shifts and by the efficacy of transport mechanisms involved in the recovery thereof, long-term ionic plasticity depends on changes in the expression patterns of proteins involved in the regulation of intraneuronal anions. The present review has its focus on two of them, KCC2 and CAVII. In the case of KCC2, there appears to be an intriguing cross-talk among GABAergic transmission, KCC2 expression and neurotrophin signalling during both neuronal development and damage. The lifetime of membrane-associated KCC2 is very short, in the range of tens of minutes, which makes KCC2 ideally suited for mediating GABAergic ionic plasticity (Wardle & Poo, 2003; Rivera et al. 2004; Kovalchuk et al. 2004). In addition, factors influencing the trafficking and kinetic modulation of KCC2 as well as activation/deactivation of CAVII are obvious candidates in the ionic modulation of GABAergic responses. It will be interesting to examine whether the expression of CAVII and other molecules that shape GABAergic transmission via anion homeostasis are interconnected by common intracellular signalling cascades, especially those involving neurotrophin actions. Work along these lines is likely to shed light on the fundamental mechanisms that are involved in the wiring, miswiring and re-wiring of neuronal networks.
, 百拇医药
    Footnotes

    This report was presented at The Journal of Physiology Symposium in honour of the late Eberhard H. Buhl on Structure/Function Correlates in Neurons and Networks, Leeds, UK, 10 September 2004. It was commissioned by the Editorial Board and reflects the views of the authors.

    References

    Aguado F, Carmona MA, Pozas E, Aguilo A, Martinez-Guijarro FJ, Alcantara S, Borrell V, Yuste R, Ibanez CF & Soriano E (2003). BDNF regulates spontaneous correlated activity at early developmental stages by increasing synaptogenesis and expression of the K+/Cl– co-transporter KCC2. Development 130, 1267–1280.
, 百拇医药
    Alvarez-Leefmans FJ & Russell JM (ed.) (1990). Chloride Channels and Carriers in Nerve, Muscle and Glial Cells. Plenum Press, New York.

    Avishai-Eliner S, Brunson KL, Sandman CA & Baram TZ (2002). Stressed-out, or in (utero) Trends Neurosci 25, 518–524.

    Barolet AW & Morris ME (1991). Changes in extracellular K+ evoked by GABA, THIP and baclofen in the guinea-pig hippocampal slice. Exp Brain Res 84, 591–598.
, 百拇医药
    Bartho P, Payne JA, Freund TF & Acsady L (2004). Differential distribution of the KCl cotransporter KCC2 in thalamic relay and reticular nuclei. Eur J Neurosci 20, 965–975.

    Ben-Ari Y (2001). Developing networks play a similar melody. Trends Neurosci 24, 353–360.

    Ben-Ari Y (2002). Excitatory actions of GABA during development: the nature of the nurture. Nat Rev Neurosci 3, 728–739.

    Ben-Ari Y, Cherubini E, Corradetti R & Gaiarsa JL (1989). Giant synaptic potentials in immature rat CA3 hippocampal neurones. J Physiol 416, 303–325.
, http://www.100md.com
    Binder DK, Croll SD, Gall CM & Scharfman HE (2001). BDNF and epilepsy: too much of a good thing Trends Neurosci 24, 47–53 (review).

    Buhl EH, Otis TS & Mody I (1996). Zinc-induced collapse of augmented inhibition by GABA in a temporal lobe epilepsy model. Science 271, 369–373.

    Clancy B, Darlington RB & Finlay BL (2001). Translating developmental time across mammalian species. Neuroscience 105, 7–17.
, http://www.100md.com
    Cohen I, Navarro V, Clemenceau S, Baulac M & Miles R (2002). On the origin of interictal activity in human temporal lobe epilepsy in vitro. Science 298, 1418–1421.

    Cohen I, Navarro V, Le Duigou C & Miles R (2003). Mesial temporal lobe epilepsy: a pathological replay of developmental mechanisms Biol Cell 95, 329–333 (review).

    DeFazio RA, Keros S, Quick MW & Hablitz JJ (2000). Potassium-coupled chloride cotransport controls intracellular chloride in rat neocortical pyramidal neurons. J Neurosci 20, 8069–8076.
, 百拇医药
    Deisz RA & Lux HD (1982). The role of intracellular chloride in hyperpolarizing post-synaptic inhibition of crayfish stretch receptor neurones. J Physiol 326, 123–138.

    Delpire E & Mount DB (2002). Human and murine phenotypes associated with defects in cation-chloride cotransport. Annu Rev Physiol 64, 803–843 (review).

    Earnhardt JN, Qian M, Tu C, Lakkis MM, Bergenhem NC, Laipis PJ, Tashian RE & Silverman DN (1998). The catalytic properties of murine carbonic anhydrase VII. Biochemistry 37, 10837–10845.
, http://www.100md.com
    Erecinska M, Cherian S & Silver IA (2004). Energy metabolism in mammalian brain during development. Prog Neurobiol 73, 397–445.

    Feller MB (1999). Spontaneous correlated activity in developing neural circuits. Neuron 22, 653–656 (review).

    Freund TF & Buzsaki G (1996). Interneurons of the hippocampus. Hippocampus 6, 347–470.

    Ganguly K, Schinder AF, Wong ST & Poo M (2001). GABA itself promotes the developmental switch of neuronal GABAergic responses from excitation to inhibition. Cell 105, 521–532.
, 百拇医药
    Garaschuk O, Linn J, Eilers J & Konnerth A (2000). Large-scale oscillatory calcium waves in the immature cortex. Nat Neurosci 3, 452–459.

    Gulacsi A, Lee CR, Sik A, Viitanen T, Kaila K, Tepper JM & Freund TF (2003). Cell type-specific differences in chloride-regulatory mechanisms and GABAA receptor-mediated inhibition in rat substantia nigra. J Neurosci 23, 8237–8246.

    Gulledge AT & Stuart GJ (2003). Excitatory actions of GABA in the cortex. Neuron 37, 299–309.
, http://www.100md.com
    Gulyas AI, Sik A, Payne JA, Kaila K & Freund TF (2001). The KCl cotransporter, KCC2, is highly expressed in the vicinity of excitatory synapses in the rat hippocampus. Eur J Neurosci 13, 2205–2217.

    Hiki K, D'Andrea RJ, Furze J, Crawford J, Woollatt E, Sutherland GR, Vadas MA & Gamble JR (1999). Cloning, characterization, and chromosomal location of a novel human K+-Cl– cotransporter. J Biol Chem 274, 10661–10667.

    Huang EJ & Reichardt LF (2001). Neurotrophins: roles in neuronal development and function. Annu Rev Neurosci 24, 677–736 (review).
, 百拇医药
    Huang EJ & Reichardt LF (2003). Trk receptors: roles in neuronal signal transduction. Annu Rev Biochem 72, 609–642.

    Hubner CA, Stein V, Hermans-Borgmeyer I, Meyer T, Ballanyi K & Jentsch TJ (2001). Disruption of KCC2 reveals an essential role of K-Cl cotransport already in early synaptic inhibition. Neuron 30, 515–524.

    Ikeda M, Toyoda H, Yamada J, Okabe A, Sato K, Hotta Y & Fukuda A (2003). Differential development of cation-chloride cotransporters and Cl– homeostasis contributes to differential GABAergic actions between developing rat visual cortex and dorsal lateral geniculate nucleus. Brain Res 984, 149–159.
, 百拇医药
    Jarolimek W, Lewen A & Misgeld U (1999). A furosemide-sensitive K+-Cl– cotransporter counteracts intracellular Cl–accumulation and depletion in cultured rat midbrain neurons. J Neurosci 19, 4695–4704.

    Kaila K (1994). Ionic basis of GABAA receptor channel function in the nervous system. Prog Neurobiol 42, 489–537 (review).

    Kaila K, Lamsa K, Smirnov S, Taira T & Voipio J (1997). Long-lasting GABA-mediated depolarization evoked by high-frequency stimulation in pyramidal neurons of rat hippocampal slice is attributable to a network-driven, bicarbonate-dependent K+ transient. J Neuroscience 17, 7662–7672.
, 百拇医药
    Kaila K, Pasternack M, Saarikoski J & Voipio J (1989). Influence of GABA-gated bicarbonate conductance on potential, current and intracellular chloride in crayfish muscle fibres. J Physiol 416, 161–181.

    Kaila K & Ransom B (ed.) (1998). pH and Brain Function. Wiley-Liss, New York.

    Kaila K, Saarikoski J & Voipio J (1990). Mechanism of action of GABA on intracellular pH and on surface pH in crayfish muscle fibres. J Physiol 427, 241–260.
, 百拇医药
    Kaila K & Voipio J (1987). Postsynaptic fall in intracellular pH induced by GABA-activated bicarbonate conductance. Nature 330, 163–165.

    Kaila K, Voipio J, Paalasmaa P, Pasternack M & Deisz RA (1993). The role of bicarbonate in GABAA receptor-mediated IPSPs of rat neocortical neurones. J Physiol 464, 273–289.

    Kakazu Y, Uchida S, Nakagawa T, Akaike N & Nabekura J (2000). Reversibility and cation selectivity of the K+-Cl– cotransport in rat central neurons. J Neurophysiol 84, 281–288.
, 百拇医药
    Kandler K (2004). Activity-dependent organization of inhibitory circuits: lessons from the auditory system. Cur Opin Neurobiol 14, 96–104.

    Kelsch W, Hormuzdi S, Straube E, Lewen A, Monyer H & Misgeld U (2001). Insulin-like growth factor 1 and a cytosolic tyrosine kinase activate chloride outward transport during maturation of hippocampal neurons. J Neurosci 21, 8339–8347.

    Khazipov R, Khalilov I, Tyzio R, Morozova E, Ben-Ari Y & Holmes GL (2004). Developmental changes in GABAergic actions and seizure susceptibility in the rat hippocampus. Eur J Neurosci 19, 590–600.
, 百拇医药
    Kovalchuk Y, Holthoff K & Konnerth A (2004). Neurotrophin action on a rapid timescale. Curr Opin Neurobiol 14, 558–563.

    Lahtinen H, Ludwig A, Karhunen T, Panula P, Kaila K & Rivera C (2004). Fast dendritic retraction and intrasomatic accumulation of KCC2 protein following neuronal trauma. The 3rd INMED/TINS Symposium, The Multiple Facets of GABAergic Neurotransmission, La Ciotat, France, 15–18 September, 2004.

    Lahtinen H, Rivera C & Kaila K (2002). An excitotoxic insult leads to fast changes in the transcription of the neuronal K-Cl cotransporter, KCC2, in hippocampal slice cultures. FENS Abstr 1, A013.7
, 百拇医药
    Lakkis MM, O'Shea KS & Tashian RE (1997). Differential expression of the carbonic anhydrase genes for CA VII (Car7) and CA-RP VIII (Car8) in mouse brain. J Histochem Cytochem 45, 657–662.

    Leinekugel X, Khazipov R, Cannon R, Hirase H, Ben-Ari Y & Buzsaki G (2002). Correlated bursts of activity in the neonatal hippocampus in vivo. Science 296, 2049–2052.

    Li H, Tornberg J, Kaila K, Airaksinen MS & Rivera C (2002). Patterns of cation-chloride cotransporter expression during embryonic rodent CNS development. Eur J Neurosci 16, 2358–2370.
, 百拇医药
    Lu J, Karadsheh M & Delpire E (1999). Developmental regulation of the neuronal-specific isoform of K-Cl cotransporter KCC2 in postnatal rat brains. J Neurobiol 39, 558–568.

    Ludwig A, Li H, Saarma M, Kaila K & Rivera C (2003). Developmental upregulation of KCC2 in the absence of GABAergic and glutamatergic transmission. Eur J Neurosci 18, 3199–3206.

    Luhmann HJ & Prince DA (1991). Postnatal maturation of the GABAergic system in rat neocortex. J Neurophysiol 65, 247–263.
, 百拇医药
    Masereel B, Rolin S, Abbate F, Scozzafava A & Supuran CT (2002). Carbonic anhydrase inhibitors: anticonvulsant sulfonamides incorporating valproyl and other lipophilic moieties. J Med Chem 45, 312–320.

    Mercado A, Mount DB & Gamba G (2004). Electroneutral cation-chloride cotransporters in the central nervous system. Neurochem Res 29, 17–25.

    Minichiello L, Casagranda F, Tatche RS, Stucky CL, Postigo A, Lewin GR, Davies AM & Klein R (1998). Point mutation in trkB causes loss of NT4-dependent neurons without major effects on diverse BDNF responses. Neuron 21, 335–345.
, 百拇医药
    Misgeld U, Deisz RA, Dodt HU & Lux HD (1986). The role of chloride transport in postsynaptic inhibition of hippocampal neurons. Science 232, 1413–1415.

    Mueller AL, Taube JS & Schwartzkroin PA (1984). Development of hyperpolarizing inhibitory postsynaptic potentials and hyperpolarizing response to gamma-aminobutyric acid in rabbit hippocampus studied in vitro. J Neurosci 4, 860–867.

    Nabekura J, Ueno T, Okabe A, Furuta A, Iwaki T, Shimizu-Okabe C, Fukuda A & Akaike N (2002). Reduction of KCC2 expression and GABAA receptor-mediated excitation after in vivo axonal injury. J Neurosci 22, 4412–4417.
, 百拇医药
    Owens DF & Kriegstein AR (2002). Is there more to GABA than synaptic inhibition Nat Rev Neurosci 3, 715–727.

    Pasternack M, Voipio J & Kaila K (1993). Intracellular carbonic anhydrase activity and its role in GABA-induced acidosis in isolated rat hippocampal pyramidal neurones. Acta Physiol Scand 148, 229–231.

    Pastorekova S, Parkkila S, Pastorek J & Supuran CT (2004). Carbonic anhydrases: current state of the art, therapeutic applications and future prospects. J Enzyme Inhib Med Chem 19, 199–229.
, 百拇医药
    Payne JA (1997). Functional characterization of the neuronal-specific K-Cl cotransporter: implications for [K+]o regulation. Am J Physiol 273, C1516–C1525.

    Payne JA, Rivera C, Voipio J & Kaila K (2003). Cation-chloride co-transporters in neuronal communication, development and trauma. Trends Neurosci 26, 199–206 (review).

    Payne JA, Stevenson TJ & Donaldson LF (1996). Molecular characterization of a putative K-Cl cotransporter in rat brain. A neuronal-specific isoform. J Biol Chem 271, 16245–16252.
, 百拇医药
    Penn AA & Shatz CJ (1999). Brain waves and brain wiring: the role of endogenous and sensory-driven neural activity in development. Pediatr Res 45, 447–458 (review).

    Race JE, Makhlouf FN, Logue PJ, Wilson FH, Dunham PB & Holtzman EJ (1999). Molecular cloning and functional characterization of KCC3, a new K-Cl cotransporter. Am J Physiol 277, C1210–C1219.

    Rivera C, Li H, Thomas-Crusells J, Lahtinen H, Viitanen T, Nanobashvili A, Kokaia Z, Airaksinen MS, Voipio J, Kaila K & Saarma M (2002). BDNF-induced TrkB activation down-regulates the K+-Cl– cotransporter KCC2 and impairs neuronal Cl– extrusion. J Cell Biol 159, 747–752.
, http://www.100md.com
    Rivera C, Voipio J, Payne JA, Ruusuvuori E, Lahtinen H, Lamsa K, Pirvola U, Saarma M & Kaila K (1999). The K+/Cl– co-transporter KCC2 renders GABA hyperpolarizing during neuronal maturation. Nature 397, 251–255.

    Rivera C, Voipio J, Thomas-Crusells J, Li H, Emri Z, Sipila S, Payne JA, Minichiello L, Saarma M & Kaila K (2004). Mechanism of activity-dependent downregulation of the neuron-specific K-Cl cotransporter KCC2. J Neurosci 24, 4683–4691.
, 百拇医药
    Ruusuvuori E, Li H, Huttu K, Palva JM, Smirnov S, Rivera C, Kaila K & Voipio J (2004). Carbonic anhydrase isoform VII acts as a molecular switch in the development of synchronous gamma-frequency firing of hippocampal CA1 pyramidal cells. J Neurosci 24, 2699–2707.

    Scozzafava A, Mastrolorenzo A & Supuran CT (2004). Modulation of carbonic anhydrase activity and its applications in therapy. Expert Opin Ther Patents 14, 667–702.
, 百拇医药
    Sernagor E, Young C & Eglen SJ (2003). Developmental modulation of retinal wave dynamics: shedding light on the GABA saga. J Neurosci 23, 7621–7629.

    Shimizu-Okabe C, Yokokura M, Okabe A, Ikeda M, Sato K, Kilb W, Luhmann HJ & Fukuda A (2002). Layer-specific expression of Cl– transporters and differential [Cl–]i in newborn rat cortex. Neuroreport 13, 2433–2437.

    Smirnov S, Paalasmaa P, Uusisaari M, Voipio J & Kaila K (1999). Pharmacological isolation of the synaptic and nonsynaptic components of the GABA-mediated biphasic response in rat CA1 hippocampal pyramidal cells. J Neurosci 19, 9252–9260.
, 百拇医药
    Staley KJ & Proctor WR (1999). Modulation of mammalian dendritic GABAA receptor function by the kinetics of Cl– and HCO3– transport. J Physiol 519, 693–712.

    Stein V, Hermans-Borgmeyer I, Jentsch TJ & Hubner CA (2004). Expression of the KCl cotransporter KCC2 parallels neuronal maturation and the emergence of low intracellular chloride. J Comp Neurol 468, 57–64.

    Sun MK & Alkon DL (2002). Carbonic anhydrase gating of attention: memory therapy and enhancement. Trends Pharmacol Sci 23, 83–89.
, 百拇医药
    Sung KW, Kirby M, McDonald MP, Lovinger DM & Delpire E (2000). Abnormal GABAA receptor-mediated currents in dorsal root ganglion neurons isolated from Na-K-2Cl cotransporter null mice. J Neurosci 20, 7531–7538.

    Supuran CT, Scozzafava A & Casini A (2003). Carbonic anhydrase inhibitors. Med Res Rev 23, 146–189.

    Thompson SM, Deisz RA & Prince DA (1988a). Relative contributions of passive equilibrium and active transport to the distribution of chloride in mammalian cortical neurons. J Neurophysiol 60, 105–124.
, http://www.100md.com
    Thompson SM, Deisz RA & Prince DA (1988b). Outward chloride/cation co-transport in mammalian cortical neurons. Neurosci Lett 89, 49–54.

    Thompson SM & Ghwiler BH (1989). Activity-dependent disinhibition. II. Effects of extracellular potassium, furosemide, and membrane potential on ECl– in hippocampal CA3 neurons. J Neurophysiol 61, 512–523.

    Titz S, Hans M, Kelsch W, Lewen A, Swandulla D & Misgeld U (2003). Hyperpolarizing inhibition develops without trophic support by GABA in cultured rat midbrain neurons. J Physiol 550, 719–730.
, 百拇医药
    Toyoda H, Ohno K, Yamada J, Ikeda M, Okabe A, Sato K, Hashimoto K & Fukuda A (2003). Induction of NMDA and GABAA receptor-mediated Ca2+ oscillations with KCC2 mRNA downregulation in injured facial motoneurons. J Neurophysiol 89, 1353–1362.

    Voipio J & Kaila K (2000). GABAergic excitation and K+-mediated volume transmission in the hippocampus. Progr Brain Res 125, 323–332.

    Voipio J, Pasternack M, Rydqvist B & Kaila K (1991). Effect of gamma-aminobutyric acid (GABA) on intracellular pH in the crayfish stretch-receptor neurone. J Exp Biol 156, 349–361.
, http://www.100md.com
    Wang C, Shimizu-Okabe C, Watanabe K, Okabe A, Matsuzaki H, Ogawa T, Mori N, Fukuda A & Sato K (2002). Developmental changes in KCC1, KCC2, and NKCC1 mRNA expressions in the rat brain. Brain Res Dev Brain Res 139, 59–66.

    Wardle RA & Poo MM (2003). Brain-derived neurotrophic factor modulation of GABAergic synapses by postsynaptic regulation of chloride transport. J Neurosci 23, 8722–8732.

    Williams JR & Payne JA (2004). Cation transport by the neuronal K+-Cl– cotransporter KCC2: thermodynamics and kinetics of alternate transport modes. Am J Physiol Cell Physiol 287, C919–C931.
, 百拇医药
    Williams JR, Sharp JW, Kumari VG, Wilson M & Payne JA (1999). The neuron-specific K-Cl cotransporter, KCC2. Antibody development and initial characterization of the protein. J Biol Chem 274, 12656–12664.

    Wyllie E (1997). The Treatment of Epilepsy: Principles and Practice. Williams & Wilkin, Baltimore.

    Yamada J, Okabe A, Toyoda H, Kilb W, Luhmann HJ & Fukuda A (2004). Cl– uptake promoting depolarizing GABA actions in immature rat neocortical neurones is mediated by NKCC1. J Physiol 557, 829–841.

    Zhang LI & Poo M-M (2001). Electrical activity and development of neural circuits. Nature Neurosci 4(Suppl.), 1207–1214 (review).

    Zhang L, Spigelman I & Carlen PL (1991). Development of GABA-mediated, chloride-dependent inhibition in CA1 pyramidal neurones of immature rat hippocampal slices. J Physiol 444, 25–49., 百拇医药(Claudio Rivera, Juha Voip)