Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2012 2024-01-30 02:21:51 |
2 Reference format revised. Meta information modification 2012 2024-01-30 03:07:03 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Villalobos, N. GABA Levels and GABAB Receptors on Network Disinhibition. Encyclopedia. Available online: https://encyclopedia.pub/entry/54502 (accessed on 06 July 2024).
Villalobos N. GABA Levels and GABAB Receptors on Network Disinhibition. Encyclopedia. Available at: https://encyclopedia.pub/entry/54502. Accessed July 06, 2024.
Villalobos, Nelson. "GABA Levels and GABAB Receptors on Network Disinhibition" Encyclopedia, https://encyclopedia.pub/entry/54502 (accessed July 06, 2024).
Villalobos, N. (2024, January 30). GABA Levels and GABAB Receptors on Network Disinhibition. In Encyclopedia. https://encyclopedia.pub/entry/54502
Villalobos, Nelson. "GABA Levels and GABAB Receptors on Network Disinhibition." Encyclopedia. Web. 30 January, 2024.
GABA Levels and GABAB Receptors on Network Disinhibition
Edit

Network dynamics are crucial for action and sensation. Changes in synaptic physiology lead to the reorganization of local microcircuits. Consequently, the functional state of the network impacts the output signal depending on the firing patterns of its units. Networks exhibit steady states in which neurons show various activities, producing many networks with diverse properties. Transitions between network states determine the output signal generated and its functional results. The temporal dynamics of excitation/inhibition allow a shift between states in an operational network. Therefore, a process capable of modulating the dynamics of excitation/inhibition may be functionally important. 

circuit neurophysiology oscillation GABA B receptor GABA levels

1. Introduction

The neural network is the substrate of brain function. The functional dynamics of neural networks modify the synaptic physiology, leading to the reorganization of these circuits [1]. These modifications modulate the firing of local units that determine the network’s functional state [2]. In the network’s functional state, neurons respond to both convergent and divergent inputs (from the network itself, recurrent local networks, or even distal networks), which can be excitatory or inhibitory. These dynamics effectively influence the output signals by modulating selective connections, leading to constant changes in the network pathways [3]. Thus, the balance between excitation and inhibition (E/I) is a fundamental parameter that influences the network’s output signal [4]. An imbalance in this parameter impacts the progression of various diseases [5][6].
Networks have a stable state in which they are not entirely inactive. In this state, the neuronal activity (including synaptic strength, intrinsic properties, and E/I balance) produces a broad spectrum of networks with various properties (size, dynamics, and linkage frameworks [7]). In an operational network, the temporal evolution between E/I is proportional among individual neurons and across the global network. This feature allows rapid transitions between states that translate into functional output signals [3]. Therefore, a process able to modulate the network state by controlling the temporal dynamics between E/I is functionally essential. This process is known as disinhibition.
Another essential signal modulation mechanism is gain control, which involves regulating the amplitude caused by the E/I balance [8]. In this framework, inhibition in the perisomatic region controls the gain of projection neuron responses, which modulates the arrangement of synaptic inputs [9]. In the network, gain control induces normalization of the average neuron firing rate according to the inputs [9][10][11]; therefore, inhibition adjusts the gain. Consequently, a regenerative depolarization that begins in the dendrites is the process by which distal excitatory inputs lead to neural activation; this process has been described in relation to disinhibition [12].
Neural network oscillations are involved in multiple physiological processes and are accepted as tools for communication among brain regions. Thus, oscillations lead to the organization and coordination of information due to the precise interactions among the activities of different neurons over time. The emergence of oscillations and their frequency ranges depend on both the intrinsic neuronal properties and the network properties [2][13].
Gamma-aminobutyric acid (GABA) is a canonical inhibitory neurotransmitter. GABA performs its function through three types of receptors: GABAA, GABAB, and GABAC. The versatility of GABA is demonstrated not only in the diversity of receptors but also in how it establishes inhibition, with phasic and tonic forms. In the phasic form, the postsynaptic GABAA receptor is activated by an increase in GABA in the synaptic cleft after release. In the tonic form, GABA overflows from the synaptic cleft and activates extrasynaptic GABA A receptors (located in the presynaptic terminal and synapses with adjacent neurons), allowing temporally and spatially slow transmission. The extrasynaptic receptors show a high affinity to GABA and mainly constitute the δ subunits assembled at α4 or α6 [14][15][16][17].

2. Involvement of GABA B Receptors

GABAB Rs are ubiquitous metabotropic receptors in the brain. These receptors couple with the Gi protein. Consequently, the Gαi/Gαo subunits inhibit adenylyl cyclase (AC), which reduces cAMP levels and thus decreases protein kinase A (PKA) activity [17]. The Gβγ subunits inhibit the Ca2+ channels and activate the GIRK and TREK-type K+ channels [18]. The affinity of the Gβγ subunit to GIRKs increases due to the Na+ that enters during the action potential; this increase in affinity leads to the opening of the channel [19].
GABAB Rs are expressed at both inhibitory and excitatory synapses. They require two different subunits for their function: GABAB1 and GABAB2. GABA binds to the GABAB1 subunit and the GABAB2 subunit causes signaling [20]. Due to gene splicing, GABAB1 has two isoforms, GABAB1a and GABAB1b, which have different N-N-terminal regions. This region determines that the GABA B1a isoform is expressed at the presynaptic terminal in excitatory synapses, thus modulating glutamate release [20][21][22].

2.1. Presynaptic Modulation

The critical presynaptic function of GABAB Rs is inhibiting the release of neurotransmitters by restricting the entry of Ca2+ into the terminal, followed by the inhibition of VGCC by the Gβγ subunit. However, this is not the only mechanism of GABAB Rs. GABAB Rs also modulate the release mechanism at various levels. For example, they modulate the SNAP-25 protein by decreasing cAMP levels, which reduces vesicular priming. During this process, the Gβγ subunit modulates neurotransmitter release through direct interaction with the SNAP-25 protein, and this interaction is modulated by synaptotagmin-1 [23].
GABAB Rs exert gain modulation at the network level through their presynaptic control. In olfactory inputs, they control differential presynaptic gain [24]. In the presynaptic neurons of the inferior colliculus, GABAB Rs control the excitability gain, and blockading them increases adaptation to the stimulus [25]

2.2. Postsynaptic Modulation

The predominant effect at the postsynaptic level is the inhibition mediated by the activation of the GIRK channels by the Gβγ subunit [26]. Once active, the Gβγ subunit forms a complex with the RGS protein, which binds to the GIRKs and accelerates their kinetics [27]. At the same time, the RGS protein increases the GTPase activity of the Gα subunit, causing rapid desensitization of the K+ current [28], which reduces neuronal excitability and inhibits action potential backpropagation.

2.3. Modulation of Network Dynamics

GABAB Rs modulate the dynamics in the network by decreasing the output current. Once activated, GABAB Rs inhibit the N-type calcium and BK-type potassium channels. Consequently, the neurons increase their degree of depolarization, incrementing their excitation level [18]. It has been suggested that inhibiting the N-type channels leads to operational advantages by expanding the transmission dynamics without influencing neurotransmitter release [18]. In the lateral geniculate nucleus, the activation of GABAB Rs leads to strong hyperpolarization, followed by rebound firing as mediated by the T-type Ca2+ channels, which improves the signal detectability while altering sensory discrimination [29]. GABAB R activation in the network is related to burst firing and constant rhythmic activity. In this way, GABAB Rs may temporally modulate slow network activity, the strength of fast activity, and the relative firing during network oscillations [30].

3. GABA Levels

Recent evidence suggests that the synaptic levels of GABA are essential to synaptic transmission and network physiology. Quantitative reports suggest that the synaptic GABA levels are in the mM range, while the extrasynaptic levels are in the μM range. However, in the nuclei of specific neurons, these levels are higher at rest and during physiological processes [26].

Accepted functional roles of the GABA levels in both synaptic transmission and the network dynamics include the generation of tonic current and the activation of extrasynaptic GABAA and GABAB Rs (including subtypes of both receptors sensitive to low GABA concentrations [31]). 

It is evident that GABA (by controlling the generation of membrane potential with different forms in various regions) modulates the temporal integration of synaptic inputs and, consequently, the activity patterns of the neuronal populations that make up the circuits, with the tonic current playing a fundamental role. Thus, the stimulating current is modulated by different processes, including the sustained activation of individual cells, the coordination of presynaptic events of different interneurons, an increase in the current density at the release site, and the reuptake mechanisms, which depend on the transporters.
GABA transporters (GATs) are widely expressed in multiple nuclei, compartments, and neuronal glia, with a higher expression in axons than dendrites [32]. The function of GATs depends on several factors. First, the amplitude of the receptor activation is transiently modulated by the GABA concentration. Second, GATs control the receptor kinetics during the recruitment of neighboring synapses. Third, GATs are involved in spatial transmission; they convert spatially confined inhibitory signals into waves without spatial restrictions, which can activate type A or B receptors at both the presynaptic and postsynaptic terminals [31][33].
The contribution of GABA Rs to tonic current has been described. Studies on neurons in the locus ceruleus suggest that GABAB Rs generate tonic currents through ERK1-dependent activation [34]. In the basal amygdala, an increase in tonic inhibition related to presynaptic regulation was reported [35].

4. Disinhibition

Functionally, a microcircuit may have a “disinhibiting” motif, including serial connections between two inhibitory interneurons and a principal excitatory neuron [36]. These circuits are involved in the response to social fear. In this process, the increase in the activity of somatostatin (SST)-positive interneurons (INs) inhibits parvalbumin-positive (PV+) INs, which causes disinhibition of the principal cells of the dorsomedial prefrontal cortex [37].
An interesting disinhibitory process was described in the lateral entorhinal cortex. In it, optogenetic silencing of VIP INs (positive for vasoactive intestinal peptide) significantly decreased the incidence of the dendritic spikes driven by the lateral entorhinal cortex, suggesting a disinhibitory effect being exerted on the dendritic activity by the INs [38]. A disinhibitory process at the dendritic level was described in the hippocampus associated with the LTP process [39]. The VIP INs in the hippocampus regulate LTP through disinhibition by activating the VPAC1 receptors (G protein-coupled receptors of the VIP/PACAP family). This molecular mechanism affects the expression and phosphorylation of the Kv4.2 K+ channels in the dendrites of the hippocampal pyramidal cells [40].
During attentional selection, selective disinhibition improved the target firing rates with resemblance to multiplicative input gain, another commonly reported effect of attention on neural responses [41]. In the cerebellum, the activation of GABAB Rs in the dentate gyrus (DG) improves the granule cell (GC) activity by reducing the synaptic inhibition imposed by hilar INs. This disinhibitory action facilitates the transfer of signals from the hippocampus.
Tourette’s syndrome (TS) is a hyperkinetic disorder characterized by motor and phonetic tics [42][43]. However, its clinical presentation is commonly accompanied by various abnormal behaviors (motor, sensory, and complex behavioral), such as inappropriate non-obscene behaviors, impulsivity, obsessive-compulsive disorder, and attention deficit hyperactivity disorder [43] The central pathophysiological substrate is disinhibition [42][43][44]. This originates from alterations in the GABAergic signaling and a loss of E/I balance in the cortex–basal ganglia–thalamus–cortex circuit associated with the modulation of motor output, motor learning, and action selection.

5. Oscillation

Various compelling studies have described the modulatory participation of GABAB Rs in slow cortical oscillations. During slow wave oscillations (SWOs), GABAB R blockade modifies three important aspects of the SWO cycle. First, GABAB R blockade increases the number of “up” states; second, GABAB R blockade affects the subsequent duration of the “down” state; and third, GABAB R activation desynchronizes the SWOs [45]. The participation of GABAB Rs according to their synaptic location has also been described. At the presynaptic level, GABAB Rs contribute to spontaneous transitions from the down state, while postsynaptic receptors are essential for the afferent termination of the up state. Thus, GABAB Rs containing the GABA subunit contribute to spontaneous termination of the up state, and GABAB Rs containing the GABAB1b subunit are essential for afferent evoked termination of the upstate [46].
Thalamic oscillations of 3 to 5 Hz are characteristic of absence epilepsy. In a rodent model, selectively blocking the GABA transporters ([GAT] to GAT-1 or GAT-3) was shown to prolong the oscillations; however, blocking both transporters inhibited the oscillations. In this same model, it was reported that extending the activity in a narrow range of GABAB Rs promoted the opening of T-type channels and intensified the oscillations [47].
In addition, tonic inhibition and disinhibition have been suggested to regulate motor activity [48]. An analysis of the cortical oscillations after GAT-1 blockade demonstrated that high GABA levels influence the beta oscillations related to movement [49].
Recently, it was reported that high GABA levels and the participation of GABAB Rs in the external globus pallidus disinhibit the RTn and thus desynchronize the beta oscillations in the motor cortex [65,66]. 

References

  1. Avena-Koenigsberger, A.; Misic, B.; Sporns, O. Communication dynamics in complex brain networks. Nat. Rev. Neurosci. 2017, 19, 17–33.
  2. Acosta-Mejia, M.T.; Villalobos, N. Neurophysiology of Brain Networks Underlies Symptoms of Parkinson’s Disease: A Basis for Diagnosis and Management. Diagnostics 2023, 13, 2394.
  3. Hahn, G.; Ponce-Alvarez, A.; Deco, G.; Aertsen, A.; Kumar, A. Portraits of communication in neuronal networks. Nat. Rev. Neurosci. 2019, 20, 117–127.
  4. Yang, W.; Sun, Q.Q. Circuit-specific and neuronal subcellular-wide E-I balance in cortical pyramidal cells. Sci. Rep. 2018, 8, 3971.
  5. Sohal, V.S.; Rubenstein, J.L.R. Excitation-inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders. Mol. Psychiatry 2019, 24, 1248–1257.
  6. Eichler, S.A.; Meier, J.C. E-I balance and human diseases—From molecules to networking. Front. Mol. Neurosci. 2008, 1, 2.
  7. Barral, J.; DReyes, A. Synaptic scaling rule preserves excitatory-inhibitory balance and salient neuronal network dynamics. Nat. Neurosci. 2016, 19, 1690–1696.
  8. Kremkow, J.; Aertsen, A.; Kumar, A. Gating of signal propagation in spiking neural networks by balanced and correlated excitation and inhibition. J. Neurosci. 2010, 30, 15760–15768.
  9. Letzkus, J.J.; Wolff, S.B.; Lüthi, A. Disinhibition, a Circuit Mechanism for Associative Learning and Memory. Neuron 2015, 88, 264–276.
  10. Salinas, E.; Thier, P. Gain modulation: A major computational principle of the central nervous system. Neuron 2000, 27, 15–21.
  11. Louie, K.; Khaw, M.W.; Glimcher, P.W. Normalization is a general neural mechanism for context-dependent decision making. Proc. Natl. Acad. Sci. USA 2013, 110, 6139–6144.
  12. Yang, G.R.; Murray, J.D.; Wang, X.J. A dendritic disinhibitory circuit mechanism for pathway-specific gating. Nat. Commun. 2016, 7, 12815.
  13. Singer, W. Neuronal oscillations: Unavoidable and useful? Eur. J. Neurosci. 2018, 48, 2389–2398.
  14. Farrant, M.; Nusser, Z. Variations on an inhibitory theme: Phasic and tonic activation of GABA(A) receptors. Nat. Rev. Neurosci. 2005, 6, 215–229.
  15. Farrant, M.; Kaila, K. The cellular, molecular and ionic basis of GABA(A) receptor signalling. Prog. Brain Res. 2007, 160, 59–87.
  16. Kasaragod, V.B.; Mortensen, M.; Hardwick, S.W.; Wahid, A.A.; Dorovykh, V.; Chirgadze, D.Y.; Smart, T.G.; Miller, P.S. Mechanisms of inhibition and activation of extrasynaptic αβ GABAA receptors. Nature 2022, 602, 529–533.
  17. Evenseth, L.S.M.; Gabrielsen, M.; Sylte, I. The GABAB Receptor-Structure, Ligand Binding and Drug Development. Molecules 2020, 25, 3093.
  18. Garaycochea, J.; Slaughter, M.M. GABAB receptors enhance excitatory responses in isolated rat retinal ganglion cells. J. Physiol. 2016, 594, 5543–5554.
  19. Wang, W.; Touhara, K.K.; Weir, K.; Bean, B.P.; MacKinnon, R. Cooperative regulation by G proteins and Na+ of neuronal GIRK2 K+ channels. eLife 2016, 5, e15751.
  20. Chalifoux, J.R.; Carter, A.G. GABAB receptor modulation of synaptic function. Curr. Opin. Neurobiol. 2011, 21, 339–344.
  21. Frangaj, A.; Fan, Q.R. Structural biology of GABAB receptor. Neuropharmacology 2018, 136 Pt A, 68–79.
  22. Shaye, H.; Stauch, B.; Gati, C.; Cherezov, V. Molecular mechanisms of metabotropic GABAB receptor function. Sci. Adv. 2021, 7, eabg3362.
  23. Alten, B.; Guzikowski, N.J.; Zurawski, Z.; Hamm, H.E.; Kavalali, E.T. Presynaptic mechanisms underlying GABAB-receptor-mediated inhibition of spontaneous neurotransmitter release. Cell Rep. 2022, 38, 110255.
  24. Root, C.M.; Masuyama, K.; Green, D.S.; Enell, L.E.; Nässel, D.R.; Lee, C.H.; Wang, J.W. A presynaptic gain control mechanism fine-tunes olfactory behavior. Neuron 2008, 59, 311–321.
  25. Ayala, Y.A.; Malmierca, M.S. The effect of inhibition on stimulus-specific adaptation in the inferior colliculus. Brain Struct. Funct. 2018, 223, 1391–1407.
  26. Hung, W.C.; Chu, Y.L.; Tsai, M.L.; Wong, S.B.; Min, M.Y.; Chen, R.F.; Yang, H.W. GABAB receptor-mediated tonic inhibition of locus coeruleus neurons plays a role in deep anesthesia induced by isoflurane. Neuroreport 2020, 31, 557–564.
  27. Luo, H.; Marron Fernandez de Velasco, E.; Wickman, K. Neuronal G protein-gated K+ channels. Am. J. Physiol. Cell Physiol. 2022, 323, C439–C460.
  28. Turecek, R.; Schwenk, J.; Fritzius, T.; Ivankova, K.; Zolles, G.; Adelfinger, L.; Jacquier, V.; Besseyrias, V.; Gassmann, M.; Schulte, U.; et al. Auxiliary GABAB receptor subunits uncouple G protein βγ subunits from effector channels to induce desensitization. Neuron 2014, 82, 1032–1044.
  29. Wang, X.X.; Jin, Y.; Sun, H.; Ma, C.; Zhang, J.; Wang, M.; Chen, L. Characterization of Rebound Depolarization in Neurons of the Rat Medial Geniculate Body In Vitro. Neurosci. Bull. 2016, 32, 16–26.
  30. Kohl, M.M.; Paulsen, O. The roles of GABAB receptors in cortical network activity. Adv. Pharmacol. 2010, 58, 205–229.
  31. Khatri, S.N.; Wu, W.C.; Yang, Y.; Pugh, J.R. Mechanisms of GABAB receptor enhancement of extrasynaptic GABAA receptor currents in cerebellar granule cells. Sci. Rep. 2019, 9, 16683.
  32. Melone, M.; Ciappelloni, S.; Conti, F. A quantitative analysis of cellular and synaptic localization of GAT-1 and GAT-3 in rat neocortex. Brain Struct. Funct. 2015, 220, 885–897.
  33. Scimemi, A. Structure, function, and plasticity of GABA transporters. Front. Cell. Neurosci. 2014, 8, 161.
  34. Wu, R.N.; Kuo, C.C.; Min, M.Y.; Chen, R.F.; Yang, H.W. Extracellular Signal-Regulated Kinases Mediate an Autoregulation of GABAB-Receptor-Activated Whole-Cell Current in Locus Coeruleus Neurons. Sci. Rep. 2020, 10, 7869.
  35. Meis, S.; Endres, T.; Munsch, T.; Lessmann, V. Presynaptic Regulation of Tonic Inhibition by Neuromodulatory Transmitters in the Basal Amygdala. Mol. Neurobiol. 2018, 55, 8509–8521.
  36. Chen, Q.; Smith, R.G.; Huang, X.; Wei, W. Preserving inhibition with a disinhibitory microcircuit in the retina. eLife 2020, 9, e62618.
  37. Xu, H.; Liu, L.; Tian, Y.; Wang, J.; Li, J.; Zheng, J.; Zhao, H.; He, M.; Xu, T.L.; Duan, S.; et al. A Disinhibitory Microcircuit Mediates Conditioned Social Fear in the Prefrontal Cortex. Neuron 2019, 102, 668–682.e5.
  38. Bilash, O.M.; Chavlis, S.; Johnson, C.D.; Poirazi, P.; Basu, J. Lateral entorhinal cortex inputs modulate hippocampal dendritic excitability by recruiting a local disinhibitory microcircuit. Cell Rep. 2023, 42, 111962.
  39. Caulino-Rocha, A.; Rodrigues, N.C.; Ribeiro, J.A.; Cunha-Reis, D. Endogenous VIP VPAC1 Receptor Activation Modulates Hippocampal Theta Burst Induced LTP: Transduction Pathways and GABAergic Mechanisms. Biology 2022, 11, 627.
  40. Artinian, J.; Lacaille, J.C. Disinhibition in learning and memory circuits: New vistas for somatostatin interneurons and long-term synaptic plasticity. Brain Res. Bull. 2018, 141, 20–26.
  41. Sridharan, D.; Knudsen, E.I. Selective disinhibition: A unified neural mechanism for predictive and post hoc attentional selection. Vis. Res. 2015, 116 Pt B, 194–209.
  42. Jackson, G.M.; Draper, A.; Dyke, K.; Pépés, S.E.; Jackson, S.R. Inhibition, Disinhibition, and the Control of Action in Tourette Syndrome. Trends Cogn. Sci. 2015, 19, 655–665.
  43. Kurvits, L.; Martino, D.; Ganos, C. Clinical Features That Evoke the Concept of Disinhibition in Tourette Syndrome. Front. Psychiatry 2020, 11, 21.
  44. Ganos, C.; Rothwell, J.; Haggard, P. Voluntary inhibitory motor control over involuntary tic movements. Mov. Disord. 2018, 33, 937–946.
  45. Perez-Zabalza, M.; Reig, R.; Manrique, J.; Jercog, D.; Winograd, M.; Parga, N.; Sanchez-Vives, M.V. Modulation of cortical slow oscillatory rhythm by GABAB receptors: An in vitro experimental and computational study. J. Physiol. 2020, 598, 3439–3457.
  46. Craig, M.T.; Mayne, E.W.; Bettler, B.; Paulsen, O.; McBain, C.J. Distinct roles of GABAB1a- and GABAB1b-containing GABAB receptors in spontaneous and evoked termination of persistent cortical activity. J. Physiol. 2013, 591, 835–843.
  47. Lu, A.C.; Lee, C.K.; Kleiman-Weiner, M.; Truong, B.; Wang, M.; Huguenard, J.R.; Beenhakker, M.P. Nonlinearities between inhibition and T-type calcium channel activity bidirectionally regulate thalamic oscillations. eLife 2020, 9, e59548.
  48. Benjamin, P.R.; Staras, K.; Kemenes, G. What roles do tonic inhibition and disinhibition play in the control of motor programs? Front. Behav. Neurosci. 2010, 4, 30.
  49. Muthukumaraswamy, S.D.; Myers, J.F.; Wilson, S.J.; Nutt, D.J.; Lingford-Hughes, A.; Singh, K.D.; Hamandi, K. The effects of elevated endogenous GABA levels on movement-related network oscillations. NeuroImage 2013, 66, 36–41.
More
Information
Subjects: Neurosciences
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 160
Revisions: 2 times (View History)
Update Date: 30 Jan 2024
1000/1000
Video Production Service