A Review on the Molecular Causes of Epilepsy

  • Journal List
  • HHS Writer Manuscripts
  • PMC4409128

Nat Neurosci. Author manuscript; available in PMC 2016 Mar i.

Published in terminal edited form every bit:

PMCID: PMC4409128

NIHMSID: NIHMS679541

Molecular mechanisms of epilepsy

Abstract

Decades of experimental work take established an imbalance of excitation and inhibition equally the leading mechanism of the transition from normal brain function to seizure. In epilepsy, these transitions are rare and abrupt. Transition processes incorporating positive feedback, such as activity-dependent disinhibition, could provide these unique timing features. A rapidly expanding array of genetic etiologies will help delineate the molecular machinery(due south). This depiction will entail quite a bit of prison cell biological science. The genes discovered to appointment are currently more than remarkable for their diversity than their similarities.

Introduction

Epileptic activeness can be induced acutely by blocking synaptic and voltage-gated inhibitory conductances1, 2, or by activating synaptic and voltage-gated excitatory conductances3. Seizures are blocked past the contrary manipulations: increasing inhibition4 or decreasing excitation5. Several decades of these types of pharmacological experiments take established the idea that an imbalance between inhibitory and excitatory conductances leads to seizures (i.e. is ictogenic) in otherwise normal encephalon tissue6. This imbalance is most clearly embodied clinically in toxic exposures such as domoic acid, which activates excitatory GluK1 glutamate receptors, or past overdoses of theophylline, which blocks the inhibitory adenosine A1 receptor7, 8. In these cases, immediate, repeated, and medically intractable seizure action is induced in otherwise normal subjects. An imbalance between excitation and inhibition is thus a validated ictogenic mechanism. Difficulties arise in extending this mechanism to epileptogenesis, that is, equally a mechanism that creates a persistent increase in the probability of spontaneous seizures.

Chronic epilepsy rather than toxic exposure is responsible for the vast majority of seizures in man patients, and this condition requires an expansion of the theory of an imbalance between inhibitory and excitatory conductances. This expansion is needed for two reasons. Showtime, unlike the acute exposures, the timing of seizures in chronic epilepsy is unpredictable and seizures are relatively rare, representing much less than 1 % of the total brain activity except in the almost severe epileptic encephalopathiesnine. Thus in chronic epilepsy, not just is an ictogenic machinery required, but this mechanism or an additional mechanism must too explain the timing of episodic transitions from normal activeness to seizures.

The 2d surface area of difficulty in applying a theory of imbalanced inhibition and excitation is that the etiology of epilepsy does non usually suggest such an imbalance. In epilepsies arising from a genetic cause, analyses of the genetic etiology accept occasionally found causal loss-of-functions mutations in inhibitory conductancesx, simply loss-of-function mutations are also found in several excitatory conductances11, 12, and the majority of causal mutations involve loss of function of genes that practise not directly modify the remainder of inhibition and excitationthirteen. Commercially available diagnostic genetic sequencing services at present feature panels of several hundred genes that have been associated with epilepsy. About of these genes exercise not encode membrane conductances. This is in line with the intermittent nature of seizures described in a higher place; genetic abnormalities that compromise important inhibitory conductances would be expected to continuously alter brain function. Thus such mutations are most frequently associated with severe epileptic encephalopathies, in which there are no normal epochs of brain activity, and frequent seizuresfourteen, 15, 16.

In the acquired epilepsies, spontaneous seizures brainstorm later on injury to a normal brain as a consequence of trauma, stroke, infection or status epilepticus. Steady-country imbalances in excitation vs. inhibition are difficult to demonstrate in established animal models of acquired epilepsy. Damage to inhibitory neurons is compensated by increases in GABAergic synaptogenesis before the onset of seizures in the pilocarpine model17. Compensatory glutamatergic synaptogenesis also occurs18, but steady-land network imbalances in excitation vs. inhibition are not apparent in experimental and human epilepsy19.

Thus we need to expand the experimentally-derived idea of fourth dimension-invariant ictogenic imbalances between inhibition and excitation to encompass both the timing of seizures and the broad variety of etiologies of human epilepsy.

Timing of seizures

I explanation for the timing of seizures in chronic epilepsy is an episodic shift in the balance of inhibition and excitation, which begs the question as to the mechanisms of the episodic shifts in this balance. Other possibilities include seizure mechanisms that are of low probability, for instance circular or reentrant activity that can occur simply in network states that exist very rarely. The probability of entry into a seizure under such weather condition may not exist a straight consequence of the molecular mechanism of ictogenesis. For example, in autosomal dominant nocturnal frontal lobe epilepsy (ADFNLE) arising from gain-of-function mutations in a nicotinic cholinergic receptor, seizures only occur in not rapid middle movement (nonREM) slumberxx (Ferini-Strambi et al. 2012). Thus network (brain) states may exist an of import determinant of seizure timing. Network states that are less probable than nonREM slumber may exist responsible for proportionately lower frequencies of seizures: for example, in catamenial epilepsy, seizures occur predominantly at specific stages of the menstrual bike21. While global brain states may help explain changes in seizure probability, they do not directly explain seizure timing. For both ADFNLE and catamenial epilepsy, most of the at-hazard periods are characterized by the absence of seizure activity. Thus in the loftier-risk periods, additional mechanisms must induce farther, ictogenic alterations of the rest of inhibition and excitation.

Unpredictable and low-frequency events tin can be readily generated by low-probability state transitions in nonlinear systems22, 23. While this trunk of analysis does non lend itself to elucidating molecular mechanisms, it does emphasize that the improbable land itself must be at to the lowest degree temporarily stable. If the ictal land were non stable, there would exist continual exits at random times from the seizure state, which generally does not occur: both experimental and human seizure durations are stereotyped, ranging from seconds for absence epilepsy to one or two minutes for chief and secondarily generalized tonic clonic seizures.

One way to explain both low-probability ictal transitions and ictal stability is to incorporate positive feedback into the ictogenic mechanism (Figure). For example, if transient imbalances in inhibition and excitation engendered farther imbalance, this imbalance could occasionally build to the indicate of engendering a seizure. Activeness-dependent reductions in the efficacy of neuronal inhibition, or activeness-dependent disinhibition, comprises ane such mechanism, or prepare of mechanisms24, 25, 26. If disinhibition occurred only at extremes of local network activeness, so the probability of ictal transition would depend on the probability of that exceptional level of local network activity. Presumably these levels of activity could occur over the same range of probabilities as those characterizing spontaneous seizures. Once a seizure was induced by an activity-dependent reduction in the residuum of inhibition to excitation, the seizure itself could continue to produce activity levels that were sufficient to suppress inhibition or enhance excitation, providing the necessary positive feedback to sustain the seizure.

An external file that holds a picture, illustration, etc.  Object name is nihms-679541-f0001.jpg

Episodic surges in network activity may rarely cross a "seizure threshold" of activeness level at which positive feedback mechanisms such as activity-dependent disinhibition boss network dynamics. The variances of the summed inputs to the epileptic network is proportional to the sum of the input variances, leading to rare surges of the intensity of input. In the case of activeness-dependent disinhibition, levels of input above this seizure threshold rapidly degrade inhibition in the epileptic network, leading to further increases network activity. This process culiminates in a seizure.

Molecular mechanisms of activity-dependent shifts in the residue of inhibition and excitation

One manner to delimit candidate ictogenic mechanisms is to consider the duration of the epochs of elevated network activity that lead to seizures. Some candidate mechanisms can be excluded if nosotros assume that the ictogenic mechanism must operate on the fourth dimension scale relevant to the epoch of increased activity. Unfortunately, we don't really know how long these high-activity epochs might be. Clues include induction of seizures in experimental animals undergoing kindling27, and in patients undergoing electroshock handling28; for such subjects, stimuli of a few seconds are sufficient to engender seizures. The duration of these stimuli represents the lower bound of epoch duration, because the applied stimuli synchronously activate many afferents, and such synchrony is unlikely to exist engendered by physiological activity.

On a fourth dimension calibration of seconds, short-term synaptic plasticity could be a relevant machinery of ictogenesis. Potentially ictogenic mechanisms of short-term synaptic plasticity include depression at inhibitory GABAergic synapses onto principal neurons24, and depression of glutamatergic synapses onto inhibitory neurons29, as well as facilitation of glutamatergic synapses between chief cellsthirty, and facilitation of GABAergic inhibitory synapses onto inhibitory interneurons 31. In brusk, either depression or facilitation, with loci at either GABAergic or glutamatergic synapses, could be proconvulsant, depending on the blazon of postsynaptic neuron. Neurotransmitter vesicle fusion with the presynaptic membrane requires the coordinate activity of a set of proteins subserving docking, priming, and fusion of the vesicles as well equally calcium entry, sensing, and export 32. Sustained fusion of neurotransmitter vesicles with the presynaptic membrane requires efficient neurotransmitter and vesicle recycling. Mutations in many of these presynaptic proteins have been demonstrated to be associated with epilepsy33, 34, whereas the anticonvulsants pregabalin and levetiracetam interact with presynaptic calcium channels and docking proteins 35. While these findings support the idea that abnormal short-term synaptic plasticity could underlie ictogenesis, it has non yet been demonstrated that abnormal short-term synaptic plasticity is necessary or sufficient to engender seizures or epilepsy 36.

Postsynaptic mechanisms of synaptic plasticity interim on slightly longer time scales that are still likely to be relevant to epochs of increased network activity include include neurotransmitter reuptake, receptor desensitization, receptor membrane trafficking, and post-translational modifications of receptor subunits affecting transmitter analogousness and channel gating. Mutations in proteins subserving some of these functions have been associated with epilepsy. For example epilepsy is engendered by mutations in the gene encoding for the stargazin fellow member of the transmembrane AMPA receptor regulatory protein (TARP) family that modulates many aspects of AMPA-type glutamate receptor subunit membrane trafficking 37, 38 likewise as by mutations in genes encoding for the LGI1 / ADAM22 proteins that link stargazin to the mail service synaptic density 39. Some of these mechanisms, such as membrane trafficking of NMDA and GABAA receptors, accept been implicated in the pathological prolongation of seizure activeness initiated in normal animals by chemic or electrical stimulation 40, 41. Intracellular signaling networks may participate in these activeness-dependent shifts in the balance of excitation and inhibition 42, 43.

Collectively the epileptogenic mutations in pre and postsynaptic proteins support the idea that functional shifts in the residuum of inhibition vs. excitation arising from abnormal synaptic plasticity could transform high-normal levels of network activeness into ictal activeness. Another prominent machinery of activity-dependent disinhibition involves dysregulation of ionic concentrations in the intra and extracellular spaces. Ions that accept been implicated in action-dependent ictogenesis include potassium, calcium, protons, and chloride 44, 45, 46, 47. The relevant proteins involved in ionic homeostasis are broadly distributed in the brain, and include transmembrane channels and transporters in the membranes of the neuronal cytoplasm and subcellular organelles. Mutations in many of the genes encoding these proteins take been associated with epilepsy 33, 15, 48. A dandy deal of experimental work has gone into testing the thought that activity-dependent shifts in ion concentrations contributes to ictogenesis. For instance, tetanic stimulation of afferents every bit well as chemically induced seizure activity results in increases in K+ o equally potassium efflux overwhelms the ability of NaKATPase to maintain the appropriate cation gradients 46. The increment in K+ o shifts the potassium reversal potential to depolarized values, so that the resting membrane potential of neurons and cotransport of other ions is compromised. Similarly, large influxes of Ca2+ through voltage and ligand-gated conductances overwhelms the capacity for restorative calcium ship, resulting in reductions in extracellular calcium 49, fifty, altering both neurotransmitter release probability and the magnitude of calcium-dependent potassium conductances.

Prolonged activation of inhibitory GABAergic conductances can also dethrone inhibition as a upshot of excitatory shifts in the GABAA reversal potential. The GABA receptor is permeable to chloride and bicarbonate, although nether near circumstances the higher concentration and permeability of chloride make it the dominant accuse carrier 51. When high levels of GABAergic synaptic activity are sustained for periods on the society of one second, the chloride flux through the GABAA receptor channel overwhelms the transport capacity of the cation-chloride cotransporters that maintain the steady-land transmembrane chloride gradients 52, 53. This is due to the finite maximum velocity of the transporters equally well as the finite velocity of NaKATPase, which maintains the appropriate driving forces for the cations whose diffusion provides the free energy for chloride cotransport 54. Nether these circumstances, chloride becomes relatively passively distributed, and then that the driving force for net ion movement through the GABAA aqueduct is largely due to bicarbonate. The bicarbonate slope does not degrade with activity because the gratis improvidence of COtwo across the membrane and subsequent hydration/rehydration via carbonic anhydrase provides a strong buffer for the bicarbonate concentration. Complimentary CO2 diffusion keeps the bicarbonate approximately symmetrically distributed on both sides of the membrane, so that bicarbonate currents reverse near 0 mV. Thus the shift to net bicarbonate flux causes the GABA conductance to become very strongly depolarizing, so that instead of inhibition, GABAergic circuits subserve excitation as long as the GABA conductance is in excess of cotransport chapters 52, 53, 55.

The activity-dependent synaptic and ionic mechanisms of disinhibition described to a higher place can result in ictal action in normal tissue subject to sufficiently prolonged synchronous input in vivo and in vitro 56, 57. Presumably, normal subjects exercise not experience seizures considering the activeness-dependent changes in synapses and ion concentrations practise not cross the threshold needed to generate a cocky-reinforcing, positive feedback cycle of increased action, disinhibition, and consequent farther increases in activeness. However, after injury a number of changes may brand such cycles more than likely. In the undercut model of post-traumatic epilepsy, NaKATPase levels and cotransport chapters are diminished, rendering ion concentrations less stable 58. Loss of function mutations in NaKATpase are associated with epilepsy 59 and pharmacological reductions in NaKATPase activity engenders seizure activeness in normal tissue 60, supporting the thought that degradations in ionic homeostasis during periods of increased network activeness is a key mechanism of ictogenesis.

Setting the stage for activity-dependent shifts in the I/E balance

If rare, loftier levels of network activity create positive feedback cycles that reduce the balance of inhibition and excitation, why isn't anybody epileptic? Has the preceding word just described the processes underlying the generalization of seizures, rather than ictogenesis? These questions suggest two possibilities: epilepsy may arise from uniquely high levels of network activity, or from unique vulnerabilities to action-dependent shifts in the balance of inhibition to excitation. In the preceding section we focused on unique vulnerabilities to high-normal levels of synaptic activeness. In this section, we consider mechanisms that may lead to uniquely high levels of network activeness.

High, ictogenic levels of network activity could occur more than readily in circuits with abnormal feedback due to cortical dysgenesis or the compensatory rewiring that occurs after brain injury. These circuit alterations are discussed in detail elsewhere in this upshot. Hither we consider the molecular mechanisms that might drive the formation of such an imbalanced excursion. A number of mutations in the PI3K / IGF / mTOR pathway have been reported in association with brain malformations that are strongly associated with epilepsy, including focal cortical dysplasias, the gliotic tubers of tuberous sclerosis, and hemimegalencephaly 61, 62. The mTOR cascade is an important anabolic cellular signaling pathway that is activated by growth factors such as IGF besides as abundance of metabolic substrates. Activation of this pathway may exist a necessary step in the outgrowth of afferent and efferent neural processes 63, 64, and then that over-activation of the mTOR pathway may pb to unnecessary neuronal connectivity and epilepsy. Supporting this hypothesis is the more robust repression of seizure action by mTOR inhibitors early in evolution 65, 63 when process outgrowth may exist more active and potentially rate-limited by substrate availability, making synaptogenesis sensitive to anabolic cues, vs. more than pocket-size effects observed in the mature brain in which process outgrowth is much more than express 66.

The status of a sufficiently anabolic land is perhaps the nigh basic limitation on the development of synaptic connectivity, which is farther influenced past a broad variety of cues that vary with location, prison cell type, and developmental stage. Accordingly, epileptic circuitry tin can be engendered past a host of defects that touch on neuronal migration 61, procedure outgrowth 67 and synaptic plasticity 68, 69, 70. For example, the protein product of ARX, or aristaless related homeobox gene, is a transcription factor that is an important determinant of neuronal migration. ARX mutations have been associated with failure of interneuron migration to their proper cortical target 71. If a population of interneurons practise not successfully complete tangential migration to a cortical circuit due to mutations in ARX, it is likely that the cortical circuit will exist chronically mis-inhibited. This is because the populations of interneurons that do successfully innervate this surface area may not exist able to presume all of the duties of the interneurons that didn't attain their target. The successfully migrating neurons may non have the advisable firing backdrop, calcium buffers, or synaptic connectivity to successfully inhibit the network under all circumstances 72. Such a network would exist decumbent to disinhibition during peaks of network activity, engendering seizures either in the mis-inhibited circuit itself, or in the downstream synaptic targets of the mis-inhibited circuit. A host of signals regulate interneuron differentiation and migration 73, and mutations in many of these signals are associated with astringent epileptic encephalopathies 14.

Interneurons may migrate to the right place in time to course the proper connections, but nonetheless be unable to provide the levels of negative feedback necessary to prevent ictogenic levels of network activity. An example of this may be the autosomal dominant SCN1A sodium channel mutations that underlies Dravet syndrome, a serious epileptic encephalopathy of childhood. A leading mechanistic hypothesis is that haploinsufficiency of the SCN1A gene results in lower densities of the Nav1.1 sodium channel, and that the concomitant reduction in sodium currents and action potential firing causes the most significant functional compromise in fast-firing GABAergic basket cells 11. In this scheme, handbasket cell function is sufficient to regulate network action under all but the college range of normal network activeness.

Neuronal homeostasis

There are many epilepsies associated with neuronal loss, either due to brain injury or to degenerative disorders. The strategies that neurons employ to rewire circuits are even less well understood than the strategies used to wire the original circuits during evolution. Neuronal homeostasis refers to the procedure by which neurons regulate their excitability. Thus in conditions that engender excess action, neurons downregulate excitatory conductances and upregulate inhibitory conductances; the reverse is true nether conditions of chronically reduced excitation 74. After injury to a brain circuit, these mechanisms would be an important means to foreclose the development of over-active neural networks. Nether most circumstances, this is successful: fifty-fifty subsequently severe brain injury, the bulk of patients practise non develop epilepsy 75. I reason that some patients develop epilepsy may be that the mechanisms underlying neuronal homeostasis fail to forbid the evolution of a seizure-prone network 76, 58.

While in that location is substantial empirical data supporting the idea of neuronal homeostasis, the molecular mechanisms are not well understood 74. Presumably the wiring and re-wiring of neuronal circuits are governed by neuronal homeostasis. Just as neurons may modify their complement of voltage-gated membrane proteins to maintain firing afterwards deafferentation, the formation and weighting of new synapses later injury to inputs may be driven past like mechanisms. Considering these mechanisms of neuronal homeostasis are unknown, we don't know how, or if, seizures impact neuronal homeostasis. Although ictal activity levels are high, the rarity and brevity of seizures may reduce their impact on the feedback mechanisms governing synaptic homeostasis. For example, a daily seizure lasting 1 minute may be a catastrophe for the patient, but implies an ictal duty cycle of less than 0.1%. The duration of epochs of activity necessary to influence neuronal homeostasis are not known. If the epileptic circuit is significantly deafferented as a consequence of injury or dysgenesis, such that inter-ictal activity is too depression, it may be that the overall activity level of the epileptic circuit (including seizures and epochs of postictal depression) averages to a level that is within the "acceptable" range of the homeostatic mechanisms. Or it may exist that ictal hyperactivity and the characteristic postictal low average to an acceptable level of action, and the processes that regulate neuronal homeostasis average the two periods together. Epilepsy might also arise from defects in the homeostatic mechanisms, for example an inability to assign an appropriately negative weight to synaptic solutions that result in seizures.

Because we don't accept a good thought as to what intracellular signaling networks regulate neuronal action levels, genetic epilepsies may provide our all-time clues equally to what some of the necessary elements of this network might be. Many of the genes whose mutations are linked to epilepsy are involved in intracellular signaling, and thus could be involved in neuronal homeostasis (Ran et al. 2014).

Neuronal homeostasis gone awry

A second approach to deciphering molecular mechanisms of neuronal homeostasis is to consider genes for channels that must exist exquisitely regulated in order to maintain normal neuronal excitability. The concept of neuronal homeostasis was starting time derived by considering the consequences of transient pharmacological derangement of voltage-gated sodium channels and ligand-gated GABAA channels 77. Are there genetic "experiments of nature" that correspond to these ideas? Dravet syndrome was mentioned earlier as an early epileptic encephalopathy resulting from mutations in genes encoding ion channels, near unremarkably SCN1A gene encoding Nav1.1 sodium channels 11 although causal genes may likewise include presynaptic proteins and ligand-gated GABAA channels 78.

Loss of office mutations in sodium channel genes would exist expected to underlie disorders of sensation, movement, and knowledge rather than epilepsy. However, a wide range of loss-of-role mutations in sodium channels upshot in epilepsy. Every bit discussed earlier, a clever proposed pathogenesis is that fast-firing inhibitory interneurons are unduly affected by some of these mutations, and substantial experimental data support this idea 11. However, why loss of a single allele is sufficient to cause disease is not known; at a molecular level, there is no evidence that the truncated protein exerts a ascendant negative event in experimental models 79. This raises the possibility that the mutation alters the expression of sodium channels in the cytoplasmic membrane by interference with the feedback mechanism neurons use to sense and conform the sodium current density. There is some prove for such a mechanism, in that sodium currents are elevated in iPSC-derived neurons bearing human sodium channel mutations, although the developmental stage at which currents were measured was all the same very far from mature and the sodium electric current density was much lower than in the mature animal 80(Liu et al. 2013). Nonetheless this experiment raises the possibility that haploinsufficiency is not the core trouble, but rather dysregulation of sodium current density that results in college currents at some developmental stages, and lower currents at others. This modulation of intrinsic excitability, along with the circuit alterations discussed before, are key elements of neuronal homeostasis 74. It is possible that conscientious consideration of the molecular pathogenesis of epilepsies such as Dravet Syndrome will provide of import clues to the nature of the mechanisms of neuronal homeostasis, as well as the means by which they may be altered in some genetic forms of epilepsy.

The variety of etiologies of epilepsies

At the fourth dimension of the last major review of the molecular mechanisms of epileptogenesis, the number of genes known to cause epilepsy was xiii, and all of these genes coded for ion channels 68. This was prior to the era of rapid, inexpensive cistron sequencing, so the predominance of ion channels was in hindsight a function of what genes were being sequenced. Unbiased sequencing and copy number studies take uncovered over 400 genes closely associated with epilepsy xiii. Many of these are discussed in the chapter on the genetics of epilepsy. The majority of the mutations associated with epilepsy are not in inhibitory or excitatory ion channels, so the mechanisms by which mutations lead to seizures and epilepsy need to be broadened. It is possible that entirely new mechanisms of epilepsy volition exist discovered as a outcome of investigations into the cell biology of these newly-discovered gene defects. However, it can also be argued that we will find instead a wider variety of means of engendering the core mechanism of episodic, self-reinforcing shifts in the residuum of excitation and inhibition. In this scenario, new genes volition be institute to act, both singly and in concert with other factors, to influence network structure in a way that leads to episodic disinhibition and seizures. For example, there are many factors that make up one's mind the stability of network connectivity, including the survival of neurons, that may be linked to epilepsy. As an example, the intractable epilepsies associated with neurodegenerative diseases 81 such every bit the neuronal ceroid lipofuscinoses 82 can be virtually readily explained as network instabilities engendered by loss of connectivity, including loss of neurons. Thus the molecular mechanisms of the epilepsies may be considered as a diverse set of pathways leading to a concluding common pathophysiology of a self-reinforcing bike of activity-dependent disinhibition.

These genetic "experiments of nature" have provided us with a wide variety of pathways to epilepsy. Where does that exit neuroscientists doing "experiments of the laboratory"? The positive feedback mechanisms that underlie the distinctive timing features of seizures can now be approached by network studies fabricated possible by new technologies in microscopy, action-dependent biomarkers, optogenetics, and computation. The timing of seizures is also ripe for report with improvements in electrode applied science, telemetry, signal assay, and seizure detection. Finally, every gene discovered has a complex cellular biological science that will take some time to unravel. With luck, the large numbers of discovered genetic mutations tin can exist exploited to develop categories of genetic etiologies. Those categories in turn may be very helpful in focusing our pathophysiological investigations. I will not try to gauge those categories at this stage of gene discovery, but I look forward to the point where categorization becomes viable.

References

1. Matsumoto H, Ajmonemarsan C. Cellular mechanisms in experimental epileptic seizures. Science. 1964 Apr 10;144(3615):193–4. [PubMed] [Google Scholar]

2. Prince DA, Wilder BJ. Control mechanisms in cortical epileptogenic foci. "Surroundings" inhibition. Arch Neurol. 1967 February;xvi(2):194–202. [PubMed] [Google Scholar]

3. Walther H, Lambert JD, Jones RS, Heinemann U, Hamon B. Epileptiform activeness in combined slices of the hippocampus, subiculum and entorhinal cortex during perfusion with low magnesium medium. Neurosci Lett. 1986 Aug 29;69(two):156–61. [PubMed] [Google Scholar]

4. Wiechert P, Herbst A. Provocation of cerebral seizures past derangement of the natural balance between glutamic acid and gamma-aminobutyric acrid. J Neurochem. 1966 Feb;xiii(2):59–64. [PubMed] [Google Scholar]

5. Croucher MJ, Collins JF, Meldrum BS. Anticonvulsant action of excitatory amino acid antagonists. Science. 1982 May 21;216(4548):899–901. [PubMed] [Google Scholar]

seven. Jett DA. Chemic toxins that cause seizures. Neurotoxicology. 2012 Dec;33(half dozen):1473–5. [PubMed] [Google Scholar]

9. Moran NF, Poole K, Bell G, Solomon J, Kendall S, McCarthy M, McCormick D, Nashef L, Sander J, Shorvon SD. Epilepsy in the Great britain: seizure frequency and severity, anti-epileptic drug utilization and bear upon on life in 1652 people with epilepsy. Seizure. 2004 Sep;13(6):425–33. [PubMed] [Google Scholar]

x. Macdonald RL, Kang JQ. mRNA surveillance and endoplasmic reticulum quality control processes alter biogenesis of mutant GABAA receptor subunits associated with genetic epilepsies. Epilepsia. 2012 Dec;53(Suppl 9):59–70. [PMC costless article] [PubMed] [Google Scholar]

11. Yu FH, Mantegazza Thou, Westenbroek RE, Robbins CA, Kalume F, Burton KA, Spain WJ, McKnight GS, Scheuer T, Catterall WA. Reduced sodium current in GABAergic interneurons in a mouse model of astringent myoclonic epilepsy in infancy. Nat Neurosci. 2006 Sep;nine(9):1142–9. [PubMed] [Google Scholar]

12. Carvill GL, Weckhuysen South, McMahon JM, Hartmann C, Møller RS, Hjalgrim H, Melt J, Geraghty Due east, O'Roak BJ, Petrou S, Clarke A, Gill D, Sadleir LG, Muhle H, von Spiczak Due south, Nikanorova M, Hodgson BL, Gazina EV, Suls A, Shendure J, Dibbens LM, De Jonghe P, Helbig I, Berkovic SF, Scheffer IE, Mefford HC. GABRA1 and STXBP1: novel genetic causes of Dravet syndrome. Neurology. 2014 Apr viii;82(14):1245–53. [PMC costless article] [PubMed] [Google Scholar]

13. Ran X, Li J, Shao Q, Chen H, Lin Z, Sun ZS, Wu J. EpilepsyGene: a genetic resource for genes and mutations related to epilepsy. Nucleic Acids Res. 2014 Oct xvi; PMID: 25324312. [PMC free article] [PubMed] [Google Scholar]

14. Epi4K Consortium; Epilepsy Phenome/Genome Project. Allen AS, et al. De novo mutations in epileptic encephalopathies. Nature. 2013 Sep 12;501(7466):217–21. [PMC complimentary article] [PubMed] [Google Scholar]

15. Veeramah KR, Johnstone L, Karafet TM, Wolf D, Sprissler R, Salogiannis J, Barth-Maron A, Greenberg ME, Stuhlmann T, Weinert Southward, Jentsch TJ, Pazzi M, Restifo LL, Talwar D, Erickson RP, Hammer MF. Exome sequencing reveals new causal mutations in children with epileptic encephalopathies. Epilepsia. 2013 Jul;54(seven):1270–81. [PMC free article] [PubMed] [Google Scholar]

16. Weckhuysen Due south, et al. KCNQ2 Study Group. Extending the KCNQ2 encephalopathy spectrum: clinical and neuroimaging findings in 17 patients. Neurology. 2013 Nov 5;81(19):1697–703. [PMC free commodity] [PubMed] [Google Scholar]

17. Zhang Due west, Yamawaki R, Wen 10, Uhl J, Diaz J, Prince DA, Buckmaster PS. Surviving hilar somatostatin interneurons enlarge, sprout axons, and form new synapses with granule cells in a mouse model of temporal lobe epilepsy. J Neurosci. 2009 Nov 11;29(45):14247–56. [PMC free article] [PubMed] [Google Scholar]

18. Buckmaster PS. Does mossy fiber sprouting requite rise to the epileptic state? Adv Exp Med Biol. 2014;813:161–8. [PubMed] [Google Scholar]

xix. Sutula TP, Dudek FE. Unmasking recurrent excitation generated by mossy fiber sprouting in the epileptic dentate gyrus: an emergent property of a circuitous system. Prog Encephalon Res. 2007;163:541–63. [PubMed] [Google Scholar]

20. Ferini-Strambi 50, Sansoni V, Combi R. Nocturnal frontal lobe epilepsy and the acetylcholine receptor. Neurologist. 2012 Nov;18(6):343–9. [PubMed] [Google Scholar]

21. Herzog AG. Catamenial epilepsy: definition, prevalence pathophysiology and handling. Seizure. 2008 Mar;17(2):151–ix. [PubMed] [Google Scholar]

22. Lopes da Silva F, Blanes W, Kalitzin SN, Parra J, Suffczynski P, Velis DN. Epilepsies every bit dynamical diseases of brain systems: bones models of the transition betwixt normal and epileptic action. Epilepsia. 2003;44(Suppl 12):72–83. [PubMed] [Google Scholar]

23. Jirsa VK, Stacey WC, Quilichini PP, Ivanov AI, Bernard C. On the nature of seizure dynamics. Brain. 2014 Aug;137(Pt 8):2210–30. [PMC free article] [PubMed] [Google Scholar]

24. Bracci E, Vreugdenhil G, Hack SP, Jefferys JG. Dynamic modulation of excitation and inhibition during stimulation at gamma and beta frequencies in the CA1 hippocampal region. J Neurophysiol. 2001 Jun;85(6):2412–22. [PubMed] [Google Scholar]

25. Patenaude C, Massicotte G, Lacaille JC. Prison cell-type specific GABA synaptic transmission and activity-dependent plasticity in rat hippocampal stratum radiatum interneurons. Eur J Neurosci. 2005 Jul;22(1):179–88. [PubMed] [Google Scholar]

26. Wester JC, McBain CJ. Behavioral country-dependent modulation of singled-out interneuron subtypes and consequences for excursion role. Curr Opin Neurobiol. 2014 Jul 22;29C:118–125. [PMC gratis article] [PubMed] [Google Scholar]

27. Morimoto One thousand, Fahnestock M, Racine RJ. Kindling and status epilepticus models of epilepsy: rewiring the encephalon. Prog Neurobiol. 2004 May;73(one):one–sixty. [PubMed] [Google Scholar]

28. Valentine M, Keddie KM, Dunne D. A comparison of techniques in electro-convulsive therapy. Br J Psychiatry. 1968 Aug;114(513):989–96. [PubMed] [Google Scholar]

29. Le Duigou C, Holden T, Kullmann DM. Short- and long-term depression at glutamatergic synapses on hippocampal interneurons by group I mGluR activation. Neuropharmacology. 2011 April;threescore(5):748–56. [PubMed] [Google Scholar]

30. Jane DE, Lodge D, Collingridge GL. Kainate receptors: pharmacology, function and therapeutic potential. Neuropharmacology. 2009 Jan;56(1):90–113. [PubMed] [Google Scholar]

31. Méndez P, Bacci A. Assortment of GABAergic plasticity in the cortical interneuron melting pot. Neural Plast. 2011;2011:976856. doi: x.1155/2011/976856. [PMC free article] [PubMed] [Google Scholar]

32. Kaeser PS, Regehr WG. Molecular mechanisms for synchronous, asynchronous, and spontaneous neurotransmitter release. Annu Rev Physiol. 2014;76:333–63. [PMC gratis article] [PubMed] [Google Scholar]

33. Rajakulendran Southward, Kaski D, Hanna MG. Neuronal P/Q-type calcium channel dysfunction in inherited disorders of the CNS. Nat Rev Neurol. 2012 January 17;eight(2):86–96. [PubMed] [Google Scholar]

34. Lazarevic Five, Pothula South, Andres-Alonso Thousand, Fejtova A. Molecular mechanisms driving homeostatic plasticity of neurotransmitter release. Front Jail cell Neurosci. 2013 Dec 3;7:244. [PMC gratuitous article] [PubMed] [Google Scholar]

35. Casillas-Espinosa PM, Powell KL, O'Brien TJ. Regulators of synaptic manual: roles in the pathogenesis and treatment of epilepsy. Epilepsia. 2012 Dec;53(Suppl 9):41–58. [PubMed] [Google Scholar]

36. Meier JC, Semtner One thousand, Winkelmann A, Wolfart J. Presynaptic mechanisms of neuronal plasticity and their role in epilepsy. Front Cell Neurosci. 2014 Jun 17;8:164. [PMC free commodity] [PubMed] [Google Scholar]

37. Letts VA, Felix R, Biddlecome GH, Arikkath J, Mahaffey CL, Valenzuela A, Bartlett FS, 2nd, Mori Y, Campbell KP, Frankel WN. The mouse stargazer gene encodes a neuronal Ca2+-channel gamma subunit. Nat Genet. 1998 Aug;19(4):340–7. [PubMed] [Google Scholar]

38. Yokoi N, Fukata M, Fukata Y. Synaptic plasticity regulated by protein-poly peptide interactions and posttranslational modifications. Int Rev Cell Mol Biol. 2012;297:i–43. [PubMed] [Google Scholar]

39. Fukata Y, Adesnik H, Iwanaga T, Bredt DS, Nicoll RA, Fukata M. Epilepsy-related ligand/receptor complex LGI1 and ADAM22 regulate synaptic manual. Science. 2006 Sep 22;313(5794):1792–5. [PubMed] [Google Scholar]

forty. Goodkin HP, Joshi Due south, Mtchedlishvili Z, Brar J, Kapur J. Subunit-specific trafficking of GABA(A) receptors during status epilepticus. J Neurosci. 2008 Mar v;28(x):2527–38. [PMC free commodity] [PubMed] [Google Scholar]

41. Naylor DE, Liu H, Niquet J, Wasterlain CG. Rapid surface accumulation of NMDA receptors increases glutamatergic excitation during status epilepticus. Neurobiol Dis. 2013 Jun;54:225–38. [PMC complimentary commodity] [PubMed] [Google Scholar]

42. Kovács R, Rabanus A, Otáhal J, Patzak A, Kardos J, Albus K, Heinemann U, Kann O. Endogenous nitric oxide is a key promoting factor for initiation of seizure-like events in hippocampal and entorhinal cortex slices. J Neurosci. 2009 Jul 1;29(26):8565–77. [PMC gratis commodity] [PubMed] [Google Scholar]

43. Chang P, Walker MC, Williams RS. Seizure-induced reduction in PIP3 levels contributes to seizure-activity and is rescued by valproic acrid. Neurobiol Dis. 2014 Feb;62:296–306. [PMC free commodity] [PubMed] [Google Scholar]

44. Fisher RS, Pedley TA, Moody WJ, Jr, Prince DA. The office of extracellular potassium in hippocampal epilepsy. Arch Neurol. 1976 Feb;33(2):76–83. [PubMed] [Google Scholar]

45. Staley KJ, Soldo BL, Proctor WR. Ionic mechanisms of neuronal excitation by inhibitory GABAA receptors. Science. 1995 Aug 18;269(5226):977–81. [PubMed] [Google Scholar]

46. Bribe CB, Ransom BR, Sontheimer H. Activity-dependent extracellular 1000+ accumulation in rat optic nerve: the role of glial and axonal Na+ pumps. J Physiol. 2000 February 1;522(Pt three):427–42. [PMC gratuitous article] [PubMed] [Google Scholar]

47. Raimondo JV, Joyce B, Kay L, Schlagheck T, Newey SE, Srinivas S, Akerman CJ. A genetically-encoded chloride and pH sensor for dissociating ion dynamics in the nervous organisation. Forepart Jail cell Neurosci. 2013 Nov thirteen;7:202. [PMC free article] [PubMed] [Google Scholar]

48. Deng H, Xiu X, Song Z. The molecular biology of genetic-based epilepsies. Mol Neurobiol. 2014 Feb;49(i):352–67. [PubMed] [Google Scholar]

49. Heinemann U, Konnerth A, Pumain R, Wadman WJ. Extracellular calcium and potassium concentration changes in chronic epileptic brain tissue. Adv Neurol. 1986;44:641–61. [PubMed] [Google Scholar]

50. Kovács R, Kardos J, Heinemann U, Kann O. Mitochondrial calcium ion and membrane potential transients follow the blueprint of epileptiform discharges in hippocampal slice cultures. J Neurosci. 2005 Apr 27;25(17):4260–ix. [PMC free article] [PubMed] [Google Scholar]

51. Macdonald RL, Botzlakis EJ. Physiology and Pathology of chloride transporters and channels in the nervous organisation. In: Alvarez-Leefmans F, Delpire Eric, editors. From molecules to diseases. Elsevier; London UK: 2009. [Google Scholar]

52. Staley KJ, Proctor WR. Modulation of mammalian dendritic GABA(A) receptor function past the kinetics of Cl- and HCO3- ship. J Physiol. 1999 Sep 15;519(Pt three):693–712. [PMC free article] [PubMed] [Google Scholar]

53. Kahle KT, Staley KJ, Nahed BV, Gamba G, Hebert SC, Lifton RP, Mountain DB. Roles of the cation-chloride cotransporters in neurological disease. Nat Clin Pract Neurol. 2008 Sep;4(ix):490–503. [PubMed] [Google Scholar]

54. Larsen BR, Assentoft M, Cotrina ML, Hua SZ, Nedergaard Chiliad, Kaila M, Voipio J, MacAulay N. Contributions of the Na+/Thousand+-ATPase, NKCC1, and Kir4.ane to hippocampal K+ clearance and volume responses. Glia. 2014 Apr;62(4):608–22. [PMC free article] [PubMed] [Google Scholar]

55. Lillis KP, Kramer MA, Mertz J, Staley KJ, White JA. Pyramidal cells accrue chloride at seizure onset. Neurobiol Dis. 2012 Sep;47(three):358–66. [PMC free commodity] [PubMed] [Google Scholar]

56. Albensi BC, Oliver DR, Toupin J, Odero Thousand. Electric stimulation protocols for hippocampal synaptic plasticity and neuronal hyper-excitability: are they effective or relevant? Exp Neurol. 2007 Mar;204(one):ane–13. [PubMed] [Google Scholar]

57. Bertram E. The relevance of kindling for man epilepsy. Epilepsia. 2007;48(Suppl 2):65–74. [PubMed] [Google Scholar]

58. Prince DA, Parada I, Graber K. Traumatic Brain Injury and Posttraumatic Epilepsy. In: Noebels JL, Avoli One thousand, Rogawski MA, Olsen RW, Delgado-Escueta AV, editors. Jasper'south Basic Mechanisms of the Epilepsies [Internet] 4th edition National Center for Biotechnology Data (United states of america); Bethesda (Doc): 2012. [Google Scholar]

59. Heinzen EL, et al. ATP1A3 Working Grouping Singled-out neurological disorders with ATP1A3 mutations. Lancet Neurol. 2014 May;13(5):503–xiv. [PMC free article] [PubMed] [Google Scholar]

threescore. Vaillend C, Mason SE, Cuttle MF, Alger BE. Mechanisms of neuronal hyperexcitability acquired past partial inhibition of Na+-Grand+-ATPases in the rat CA1 hippocampal region. J Neurophysiol. 2002 Dec;88(6):2963–78. [PubMed] [Google Scholar]

61. Guerrini R, Dobyns WB. Malformations of cortical development: clinical features and genetic causes. Lancet Neurol. 2014 Jul;xiii(7):710–26. [PMC free article] [PubMed] [Google Scholar]

62. Mirzaa GM, Poduri A. Megalencephaly and hemimegalencephaly: breakthroughs in molecular etiology. Am J Med Genet C Semin Med Genet. 2014 Jun;166C(two):156–72. [PubMed] [Google Scholar]

63. Berdichevsky Y, Dryer AM, Saponjian Y, Mahoney MM, Pimentel CA, Lucini CA, Usenovic Chiliad, Staley KJ. PI3K-Akt signaling activates mTOR-mediated epileptogenesis in organotypic hippocampal culture model of post-traumatic epilepsy. J Neurosci. 2013 May 22;33(21):9056–67. [PMC costless article] [PubMed] [Google Scholar]

64. Lasarge CL, Danzer SC. Mechanisms regulating neuronal excitability and seizure evolution following mTOR pathway hyperactivation. Front end Mol Neurosci. 2014 Mar xiv;7:18. [PMC complimentary commodity] [PubMed] [Google Scholar]

65. Wong Yard. A critical review of mTOR inhibitors and epilepsy: from basic scientific discipline to clinical trials. Expert Rev Neurother. 2013 Jun;13(6):657–69. [PMC free article] [PubMed] [Google Scholar]

66. Buckmaster PS, Lew FH. Rapamycin suppresses mossy fiber sprouting but non seizure frequency in a mouse model of temporal lobe epilepsy. J Neurosci. 2011 February 9;31(6):2337–47. [PMC free commodity] [PubMed] [Google Scholar]

67. Gardiner J, Marc J. Disruption of normal cytoskeletal dynamics may play a key function in the pathogenesis of epilepsy. Neuroscientist. 2010 Feb;16(1):28–39. [PubMed] [Google Scholar]

68. McNamara JO, Huang YZ, Leonard AS. Molecular signaling mechanisms underlying epileptogenesis. Sci STKE. 2006 Oct 10;2006(356):re12. [PubMed] [Google Scholar]

69. Fernández East, Rajan Northward, Bagni C. The FMRP regulon: from targets to disease convergence. Front Neurosci. 2013 October 24;7:191. [PMC free article] [PubMed] [Google Scholar]

70. Na ES, Nelson ED, Kavalali ET, Monteggia LM. The impact of MeCP2 loss- or gain-of-role on synaptic plasticity. Neuropsychopharmacology. 2013 Jan;38(1):212–ix. [PMC free article] [PubMed] [Google Scholar]

71. Marsh ED, Golden JA. Developing Models of Aristaless-related homeobox mutations. In: Noebels JL, Avoli M, Rogawski MA, Olsen RW, Delgado-Escueta AV, editors. Jasper's Basic Mechanisms of the Epilepsies [Internet] 4th edition National Center for Biotechnology Data (US); Bethesda (Physician): 2012. [PubMed] [Google Scholar]

72. Kato Grand, Dobyns WB. Ten-linked lissencephaly with abnormal genitalia every bit a tangential migration disorder causing intractable epilepsy: proposal for a new term, "interneuronopathy" J Child Neurol. 2005 Apr;20(four):392–seven. [PubMed] [Google Scholar]

73. Southwell DG, Nicholas CR, Basbaum AI, Stryker MP, Kriegstein AR, Rubenstein JL, Alvarez-Buylla A. Interneurons from embryonic development to cell-based therapy. Science. 2014 Apr 11;344(6180):1240622. [PMC gratuitous article] [PubMed] [Google Scholar]

74. O'Leary T, Wyllie DJ. Neuronal homeostasis: time for a alter? J Physiol. 2011 Oct fifteen;589(Pt 20):4811–26. [PMC free article] [PubMed] [Google Scholar]

75. Frey LC. Epidemiology of posttraumatic epilepsy: a critical review. Epilepsia. 2003;44(Suppl 10):11–7. [PubMed] [Google Scholar]

76. Volman V, Bazhenov One thousand, Sejnowski TJ. Pattern of trauma determines the threshold for epileptic activity in a model of cortical deafferentation. Proc Natl Acad Sci U S A. 2011 Sep 13;108(37):15402–7. [PMC free commodity] [PubMed] [Google Scholar]

77. Turrigiano GG, Leslie KR, Desai NS, Rutherford LC, Nelson SB. Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature. 1998 February 26;391(6670):892–6. [PubMed] [Google Scholar]

78. Carvill GL, et al. GRIN2A mutations cause epilepsy-aphasia spectrum disorders. Nat Genet. 2013 Sep;45(ix):1073–6. [PMC free commodity] [PubMed] [Google Scholar]

79. Ogiwara I, Miyamoto H, Morita N, Atapour N, Mazaki E, Inoue I, Takeuchi T, Itohara South, Yanagawa Y, Obata G, Furuichi T, Hensch TK, Yamakawa K. Nav1.1 localizes to axons of parvalbumin-positive inhibitory interneurons: a excursion basis for epileptic seizures in mice carrying an Scn1a cistron mutation. J Neurosci. 2007 May 30;27(22):5903–14. [PMC free article] [PubMed] [Google Scholar]

80. Liu Y, Lopez-Santiago LF, Yuan Y, Jones JM, Zhang H, O'Malley HA, Patino GA, O'Brien JE, Rusconi R, Gupta A, Thompson RC, Natowicz MR, Meisler MH, Isom LL, Parent JM. Dravet syndrome patient-derived neurons suggest a novel epilepsy mechanism. Ann Neurol. 2013 Jul;74(ane):128–39. [PMC free article] [PubMed] [Google Scholar]

81. Girard JM, Turnbull J, Ramachandran North, Minassian BA. Progressive myoclonus epilepsy. Handb Clin Neurol. 2013;113:1731–6. [PubMed] [Google Scholar]

82. Chabrol B, Caillaud C, Minassian B. Neuronal ceroid lipofuscinoses. Handb Clin Neurol. 2013;113:1701–half-dozen. [PubMed] [Google Scholar]

hollinstheivein88.blogspot.com

Source: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4409128/

0 Response to "A Review on the Molecular Causes of Epilepsy"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel