U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Noebels JL, Avoli M, Rogawski MA, et al., editors. Jasper's Basic Mechanisms of the Epilepsies [Internet]. 4th edition. Bethesda (MD): National Center for Biotechnology Information (US); 2012.

  • This title is an author manuscript version first made accessible on the NCBI Bookshelf website July 2, 2012.

This title is an author manuscript version first made accessible on the NCBI Bookshelf website July 2, 2012.

Cover of Jasper's Basic Mechanisms of the Epilepsies

Jasper's Basic Mechanisms of the Epilepsies [Internet]. 4th edition.

Show details

Histopathology of Human Epilepsy

, , and .

Author Information and Affiliations

Histopathologic lesions associated with seizures are found in a variety of disease conditions. The histopathology of temporal lobe epilepsy (TLE) is the most extensively investigated. In TLE with hippocampal sclerosis, several lines of evidence implicate the hippocampus as the predominant focus of seizure origin and maintenance In sclerotic hippocampi, two major pathological processes are observed. (1) The reorganization of the neural circuitry of the dentate gyrus that favors the hyperexcitability of dentate granule cells. These changes include the loss of populations of subgranular hilar interneurons, hilar mossy cells and granule cells. Additionally there is sprouting of granule cell mossy fiber recurrent collaterals into the dentate inner molecular layer and the sprouting of neuropeptide Y, somatostatin and substance P containing axons throughout the molecular layer. (2) The second process consists of glial activation and proliferation throughout Ammon’s horn along with neuronal loss. These astrocytes show a number of changes -- increased expression of specific glutamate receptors, redistribution of the water transporter AQP4, increased expression of Na+ and Ca2+ ion channels with down regulation of the Kir4.1 channels. The enzymes glutamine synthetase and lactate dehydrogenase are reduced whereas a number of molecules associated with immune function are upregulated. The density of the microvasculature is increased. The “astrocytes” associated with these changes may be subcategorized into two populations -- the strongly GFAP positive astrocytes and weakly GFAP reactive cells, which resemble the GluR astrocytes also known as NG2 cells. The functions of these two subpopulations may favor the creation of a high extracellular glutamate and K+ containing environment in the sclerotic hippocampus, which in turn may trigger seizure spread through a fairly normal subiculum. These changes may even provide an excitable substrate through which the hyperexcitability of dentate granule cells may spread out of the hippocampus to generate seizures.

INTRODUCTION

Histopathological lesions that are associated with seizures are described in several neurological disease states. Such histopathological lesions are found in developmental disorders, neoplasms, microbial diseases, cerebrovascular diseases, trauma, and immune-mediated disorders, and in disorders primarily related to seizures such as in temporal lobe epilepsy (TLE). General descriptions of these lesions have been the subject of several reviews. 1, 2 Temporal lobe epilepsy, a major seizure disorder which affects over 40 million people worldwide, has received the most study in relation to its histopathology, and the histological lesion is thus best understood in relation to epilepsy. This chapter will review the state of our knowledge in the histopathology of TLE.

Pierre Gloor provides an excellent review of the historical studies of TLE. 3 However, some of the more significant of these studies bear revisiting for our current appreciation of the histopathology of TLE. The earliest systematic attempt at understanding the histopathology of TLE was by Sommer. 4 He reviewed all previously published papers on the pathology of epilepsy and examined (macroscopically) a few of his own cases; he estimated that pathology in Ammon’s horn was found in about 30% of cases and that the pathology was more often unilateral. Studying the histology of only one case, he described diffuse cell loss limited to a sector of the hippocampus, field H1 of Rose (area CA1), which subsequently became known as the Sommer sector. This condition, which manifests as an increased hardening of the tissue, was first noted by Bouchet and Cazauvieilh 5 and termed Ammon’s horn sclerosis. Bratz, 6 studying autopsied brains of epilepsy patients, confirmed the loss of neurons in the Ammon’s horn in about 50% of the brains, and also noted that such loss was predominantly unilateral. He provided a more complete description of a sclerotic hippocampus, with accompanying woodcuts to illustrate his findings. Bratz observed the following: almost complete pyramidal cell loss in CA1 and a sharp boundary with the subiculum which showed no cell loss; preservation of spindle shaped cells in the stratum oriens; less complete cell loss in the hilus (end folium) and dentate gyrus; preservation of pyramidal cells in area CA2; and an abundance of blood vessels in the atrophic sector (area CA1). These detailed observations have held up as accurately describing the gross histology of a sclerotic hippocampus. 7, 8

Whether hippocampal pathology is specially associated with TLE is still a topic of considerable discussion. To date, several lines of evidence suggest that the hippocampus does play an important role in TLE. Positron emission tomography (PET) studies demonstrate that the most significant hypometabolism is seen in the lateral temporal and medial temporal regions, with the degree of hypometabolism correlating with the severity of hippocampal pathology. 9 Hippocampal cell densities are significantly related to the amount of hypometabolism in the thalamus and basal ganglia in TLE patients. 10 Further, anteromedial temporal lobectomy, with removal of the hippocampus and amygdala, controls seizures and leads to improvement of the pre-operative hypometabolism in the remaining areas of the temporo-limbic network (e.g. inferior frontal temporal lobe, ipsilateral temporal neocortex and medial thalamus). 11 Stauder 12 was the first to associate hippocampal sclerosis with TLE. He made a list of ictal symptoms and signs attributable unequivocally to temporal lobe disease (olfactory and gustatory auras, sensory aphasia and paraphasia, vertigo, auditory hallucinations or illusions and dreamy states) and correlated these with the postmortem pathology of 53 cases of “genuine epilepsy.” Sixty eight percent of the brains showed hippocampal sclerosis, of which 80% had definite and another 11% probable symptoms of “genuine epilepsy,” whereas none of those without hippocampal sclerosis showed symptoms. Falconer 13 who developed the technique of en bloc resection of the temporal lobe for treatment of drug-resistant epilepsy, found a high correlation of good surgical outcome with mesial temporal sclerosis in the resected temporal lobe. Further, intracranial EEG recordings in patients with TLE associated with hippocampal sclerosis show that seizures originate in the sclerotic hippocampus. 14 More recently, a multicenter study of seizure outcome after medial temporal resective surgery reports that hippocampal atrophy (sclerosis) and a history of absence of tonic-clonic seizures were the sole predictors of remission of seizures 2 years after surgery. 15

Bratz 6 suggested that the hippocampal pathology in TLE cannot be assumed to be a consequence of seizures but must play a pathogenetic role. Whether hippocampal sclerosis is the cause or a consequence of seizures, has remained a topic of lively debate over the years. Several more recent observations indicate that hippocampal sclerosis is probably not a consequence of seizures. MRI volume measurements show that the degree of hippocampal atrophy is not correlated with the duration and severity of seizures. 16 Some patients with multiple daily seizures since infancy do not have hippocampal sclerosis. 17 Likewise, a study of post-mortem hippocampi from patients with poorly controlled seizures has confirmed a subgroup with an absence of significant neuronal loss despite decades of seizures. 18 Additionally, hippocampi surgically removed from a large series of patients with temporal lobe epilepsy show that in about 15% of them (paradoxical temporal lobe epilepsy) there is no hippocampal sclerosis or evidence of extra hippocampal mass lesions, though their seizure semiology and history is indistinguishable from those with hippocampal sclerosis. 7 Thus, seizures themselves do not result in hippocampal sclerosis. A more detailed and careful investigation of the histopathology of the sclerotic hippocampus and its relationship to the pathogenesis of human TLE is important, and since the 1980’s there has been resurgence in investigations of the histopathology of the hippocampus in TLE. This renewed effort was ignited by a wider use of the technique of en bloc resection of the hippocampus by neurosurgeons 19, 20 and the availability of new methods for histological analysis. This chapter will attempt to summarize this considerable body of recent literature, various aspects of which have been the subject of other detailed reviews. 7, 21, 22

HISTOPATHOLOGICAL VARIATIONS IN TLE HIPPOCAMPUS

In their now classic study, Margerison and Corsellis 23 examined the brains of 55 epileptic patients postmortem, in which clinical and electroencephalographic assessments of seizures during life were available. The hippocampi of this group showed varying degrees of histological change. Thirty five percent of the hippocampi showed no evidence of neuronal loss bilaterally while the remainder showed varying degrees of hippocampal sclerosis, ranging from classical hippocampal sclerosis (40%) to end folium sclerosis (25%). With the advent of surgery for the control of intractable epilepsy, further analyses of excised hippocampi were undertaken. Bruton 24 identifies a variety of hippocampal pathologies using neuronal and myelin stains. 23 More recently, in another surgical series the use of neuronal cell counts with immunohistochemical stains for neurotransmitters, has led to the identification of several variations in the histopathology of the hippocampus in patients with medically intractable TLE. 7, 25 Based on conventional histopathologic criteria, hippocampi were divided into two groups – those without classical hippocampal sclerosis (compared to autopsy controls from neurologically normal subjects) and those with hippocampal sclerosis (Figure 1). Those without classical hippocampal sclerosis had a small loss (<25%) of neurons throughout the hippocampus, with hippocampal volumes similar to autopsy controls and none of the immunohistochemical changes observed in the sclerotic hippocampi. However, these non-sclerotic hippocampi are divisible into two sub-groups based on epilepsy etiology: the mass-associated TLE group (MaTLE) had an extrahippocampal temporal lobe mass lesion (low grade gliomas, cavernomas), as opposed to a group described as “paradoxical” temporal lobe epilepsy (PTLE) in which such mass lesions were absent and there was no history of febrile seizure or other identifiable etiological factors. Distinctive patterns correlating neuronal densities of the dentate gyrus and all areas of Ammon’s horn, electrophysiologic measures, and power spectral densities further differentiate these sclerotic and non-sclerotic groups. 7, 25, 26 The hippocampi with hippocampal sclerosis (MTLE) show more extensive neuronal loss and some distinctive histopathological features. These features could be grouped under two broad categories: First, features associated with the neural reorganization of the dentate gyrus; and the second, the proliferation of astrocytes accompanying neuron loss in the Ammon’s horn (Cornu Ammonis; CA), especially noticeable in area CA1. The details of these changes are described in the sections below.

Figure 1. Photomicrographs of coronal sections of the hippocampus from a non-sclerotic (A) and sclerotic (MTLE) hippocampus stained with Cresyl Violet.

Figure 1

Photomicrographs of coronal sections of the hippocampus from a non-sclerotic (A) and sclerotic (MTLE) hippocampus stained with Cresyl Violet. (A) The non-sclerotic hippocampus is from a temporal lobe epilepsy patient with an extrahippocampal mass lesion (more...)

REORGANIZATION OF THE DENTATE GYRUS IN SCLEROTIC HIPPOCAMPUS

Neuronal Changes in Dentate Gyrus

Granule cells

The principal neurons of the dentate gyrus are the granule cells, which form a densely packed C-shaped band of cells in a normal hippocampus. The granule cell has a cone-shaped tree of apical dendrites, which extend through the dentate molecular layer, 27 and in the human there are basal dendrites, which extend into the hilus as well as the molecular layer. 28 The granule cell axon, called the mossy fiber, extends into the hilus where its collaterals synapse with hilar interneurons, continuing on to synapse on pyramidal neurons in area CA3. 27 Historically, the granule cells are described as a band of cells that are relatively resistant to injury and loss in the sclerotic hippocampus. However, such resistance is visible only in a small proportion of TLE hippocampi, such as in a group of patients showing neuronal loss only in area CA1, as well as MaTLE and PTLE groups. 7 In most of the MTLE hippocampi there is loss of granular neurons resulting in a thinning of the layer and gaps in the cell band. 29 In other cases there is a vertical spread or dispersion of remaining granular neurons into the dentate molecular layer. 21 Reduced reelin mRNA levels are associated with granule cell dispersion, suggesting that the dispersion may be a developmental abnormality associated with granule cell migration. 30 Nevertheless, granule cell dispersion is not accompanied by enhanced neurogenesis in TLE patients. 31 Further, granule cell dispersion is not exclusive to hippocampi with sclerosis. 21, 32 Granule cell dispersion seems more correlated with memory and learning changes than with seizures, 21 though some studies suggest it is correlated with seizure frequency. 33 The surviving dentate granule cells in the sclerotic hippocampus show distinctive morphological alterations compared to those in non-sclerotic hippocampi. The apical dendrites of the granule cells in sclerotic hippocampi have a more limited spread in the longitudinal axis, a significant increase in the length of the portion of dendrites in the inner molecular layer, a decrease in the dendritic portion in the outer third, and an increase in spine density in both regions. 34 A proportion of the granule cells in sclerotic hippocampi also has axon collaterals from their mossy fibers extending into the molecular layer 35 and these neurons have a significantly higher spine density in the proximal portion of the dendrite located in the inner molecular layer. 36

Granule cells have a complex neurochemical profile. While glutamate is an important neurotransmitter produced by these cells, they also contain the neuropeptide dynorphin, 29 which increases in TLE hippocampi, 37 the calcium binding protein calbindin 38 and the hyperpolarization-activated cyclic nucleotide-gated cation channel 1 (HCN1), which is markedly increased when the granule cell density is <50%. 39

Hilar interneurons

The hilar region of the normal human hippocampus has many types of neurons. Among these are the mossy cells, dentate pyramidal basket cells, and a variety of other interneurons characterized by their neurochemical content. 25 One of the early observations in relation to the histopathology of TLE is that interneurons mostly located in the subgranular region of the hilus and containing neuropeptide Y (NPY), somatostatin (SOM) and substance P (SP) were significantly reduced in number in the sclerotic hippocampus. 40 Many of these neurons also co-localize GABA and their loss may result is some loss of GABAergic inhibition. The dentate pyramidal basket cells, which are located on the hilar edge of the granule cell layer and are also GABAergic, are preserved in sclerotic hippocampi. 41

The mossy cell is a neuron type of particular importance in TLE. 42 Large and complex spines called thorny excrescences, which cover their proximal dendrites and from which the cell derives its name 43 characterize these neurons, which are often triangular or multipolar in shape. These spines are the sites of termination of the granule cell mossy fiber axon. Mossy cells are excitatory glutamatergic neurons. 44 They can inhibit granule cells by directly exciting the pyramidal basket cells, which inhibit the granule cells. 45 In the rodent, the mossy cells are a very abundant cell type in the hilus. Mossy cells in primates (humans) and rodent show several differences. In the human they are not the most abundant or largest cell type in the hilus, and the dendrites of most mossy cells penetrate into the dentate molecular layer and receive inputs from the perforant path. 46 In rodent models of seizures, the mossy cells are selectively vulnerable, and it is proposed that their death results in cellular reorganization favoring seizures. 42 In the human TLE hippocampus, mossy cells do not appear to be as vulnerable. Though the number of mossy cells is reduced, they have been identified in sclerotic hippocampi by increased expression of markers such as dynorphin, 47 GluR1 receptor 48 and cocaine and amphetamine-regulated transcript peptide. 49 In several sclerotic hippocampi large numbers of mossy cells are observed even when most pyramidal neurons in CA1 and CA3 are lost.47

Plasticity of Neuronal Fibers

In addition to the loss of neurons in the TLE hippocampus, there is considerable evidence for the growth of new fiber systems or “sprouting.” These changes are seen in the MTLE or sclerotic hippocampus. The most distinctive of these is the growth of recurrent collaterals from granule cell mossy axons into the inner molecular layer. Such sprouting is recognized by increased Timm stain, which visualizes the zinc contained in mossy fibers 50, 51 and by immunostaining for dynorphin (a peptide contained exclusively in dentate granule cells in non-sclerotic hippocampi). 40, 52 These recurrent collaterals form synaptic contacts with the proximal portions of the apical dendrites of granule cells. 50, 53 Enhanced immunoreactivity for dynorphin is also seen in surviving mossy cells and their thorny excrescences only in sclerotic hippocampi. 47 Four other examples of sprouting are observed in the MTLE hippocampus compared to other patient groups and autopsy controls. 25, 40 The most striking of these examples is the increased density of NPY immunoreactive fibers throughout the dentate molecular layer. Somatostatin and substance P fibers show a similar increase. Ultrastructural studies show that all these fibers are axons, which form synaptic contacts with the granule cell soma and dendrites. 53 Acetylcholinesterase staining shows decreased staining in the inner molecular layer (IML) and increased staining in the outer molecular layer of MTLE hippocampi compared to non-sclerotic hippocampi. 54

Changes in Neurotransmitter Receptors

Associated with the reorganization of neurons and neural circuits in the dentate gyrus of the sclerotic hippocampus (MTLE) are several changes in the distribution of neurotransmitter receptors within the dentate gyrus as compared to non-sclerotic surgically removed hippocampi (MaTLE and PTLE) and neurologically normal autopsy controls. These changes are summarized in Table 1. Several receptor subtypes that mediate glutamatergic transmission are changed. The AMPA type receptor GluR 1 and GluR2/3 subunits are increased on the apical dendrites of the granule cells throughout the IML while the GluR2/3 subunits are also increased on the granule cell bodies. 55 Such an increase in glutamate receptors may facilitate the excitation of granule cells through glutamatergic inputs such as the recurrent collaterals from mossy fibers or glutamatergic inputs from surviving hilar mossy cells (which terminate in the IML), and entorhinal inputs (which synapse on dendrites in the outer ML). 22 Up regulation of these AMPA type receptor subunits on the thorny excrescences of mossy cells 56 suggests a facilitation of glutamatergic transmission from remaining granule cells through surviving mossy cells. The role of the metabotropic receptor subtypes (Table 1), which are also upregulated in the molecular layer and the granule cell layer, 57–59 is unclear at present but may modulate excitatory and/or inhibitory effects. 60 Concomitant with up regulation of these glutamate receptors, there is also evidence of upregulation of inhibitory receptors. Thus, the GABAA receptor subtypes α1 and α2 61 along with GABAB receptors 62, 63 are upregulated on the granule cells somata, with the GABAA α2 subtype significantly upregulated also on the proximal portion of the granule cells dendrite in the IML. 61 GABAA and GABAB receptor subtypes are also upregulated in the hilus, though their exact cellular location is unclear. 61 The receptors for somatostatin are upregulated in the granule cell layer and dentate molecular layer, particularly the IML as well as the hilus, 64, 65 and are consistent with the sprouting of somatostatin containing axons into these regions. 40 Likewise, the NPY-Y2 receptors are upregulated on the granule cell layer and the IML as well as the hilus, 65, 66 consistent with the sprouting of NPY positive axons in these regions. The dopamine D2 receptors are upregulated in the granule cell layer and IML 67 while VIP receptors are upregulated in the granule cell layer and hilus. 68 Several of these receptor changes -- somatostatin, NPY-Y2, GABA, dopamine D2 and some of the mGluR receptors may be involved in increasing the inhibition of the excitability of surviving granule cells. 69 It thus appears that in the sclerotic hippocampus, along with processes that facilitate hyperexcitability, there are also several reorganizational changes that seem to reestablish inhibition, though perhaps inadequately.

TABLE 1. Neurotransmitter Receptor Localization in the Dentate Gyrus of the Sclerotic (MTLE) Hippocampus.

TABLE 1

Neurotransmitter Receptor Localization in the Dentate Gyrus of the Sclerotic (MTLE) Hippocampus.

Implications of Reorganizational Changes in the Dentate Gyrus

The implications of this wide array of histopathological changes in the dentate gyrus for epileptogenesis or the maintenance of seizures in patients with TLE associated with hippocampal sclerosis is difficult to evaluate with any certainty, though several attempts have been made (see 22, 70). However, though it is commonly thought that the dentate gyrus granule cells are relatively resistant to injury and thus play a role in seizure generation, study of a large series of sclerotic hippocampi reveal cases in which there are extremely few granule cells remaining 29 – and so are unlikely to be of functional significance. Yet these patients had intractable epilepsy. In an intermediate state of neuronal loss, it is likely that the dentate gyrus contributes to the epileptic state. Electrophysiological recordings from granule cells in such cases indicate that these neurons are hyperexcitable. 71 A more integrated understanding of these several changes, through experimental investigation, is needed.

CHANGES IN AMMON’S HORN AND THE ROLE OF ASTROCYTES

Neurons

The Ammon’s horn (the hippocampus proper) is divided into three distinct fields identified as CA3, CA2 and CA1. Viewed in cross section, it is composed of a single band of pyramidal cells (the stratum pyramidale) whose apical dendrites extend into a stratum radiatum. Between the pyramidal layer and the alveus is a network of small neurons forming the stratum oriens. Ammon’s horn is a region most vulnerable to neuronal loss in TLE patients. 23, 72, 73 The pyramidal neurons are the cell type most prominently lost. The degree of neuronal loss can vary from patient to patient, and usually ranges from greater than 50% compared to neurologically normal controls 8 to almost complete loss. In some TLE patients, neuronal loss is seen only in the Ammon’s horn, with the dentate gyrus being intact (CA only; see 7). In a small proportion of patients, neuronal loss is confined to the dentate hilus and area CA3, called end-folium sclerosis. 23 In the non-sclerotic patient groups MaTLE and PTLE (see above), Ammon’s horn appears intact on merely visual inspection, but neuronal counts show about a 20% loss of neurons, as in the dentate gyrus, and are qualitatively indistinguishable from hippocampi of neurologically normal subjects. 7

In the sclerotic hippocampi, even in those with extensive pyramidal cell loss, several populations of neurochemically-defined interneurons remain. These neurons are especially prominent in the stratum oriens/alvear region where they appear as characteristic horizontal neurons. In MTLE hippocampi these neurons are immunoreactive for GABA, neuropeptide Y, somatostatin, substance P (see Figure 6, 74) and express glutamate receptor protein subunits GluR1, GluR2/3, mGluR1 and mGluR5. 57, 75 These stratum oriens neurons, which appear to form a network throughout the stratum oriens, are also more strongly immunoreactive for the enzyme phosphate-activated glutaminase. 76. Stratum oriens neurons are characterized by having soma and dendrites which extend horizontally in the stratum oriens; their axonal arbors show distinctive projections, which serve as the basis on which they are subcategorized. 77 The physiological properties of these neurons have been studied extensively in normal animals, where they have been shown to modulate the excitability of Ammon’s horn pyramidal cells. 78 While these neurons have been studied in animal models of epilepsy and reported to be even more excitable than pyramidal cells 79 their functions in human sclerotic hippocampi are completely unknown.

Astrocytes

The most recognized feature in the histopathology of the sclerotic hippocampus is the proliferation of astrocytes in the neuron-depleted Ammon’s horn. 4 These astrocytes exhibit many unusual properties compared to astrocytes from non-epileptic brain regions, and have been the subject of previous reviews. 80, 81 These changes are seen in their membrane proteins (receptors, transporters, ion channels), membrane physiology, enzymes, gene expression and neurovascular relationships.

Astrocytes from sclerotic hippocampi have increased expression of the glutamatergic receptors mGluR2/3, mGluR4 and mGluR8. 82 Activation of these receptors is known to lead to intracellular Ca2+ release and Ca2+ wave propagation, a phenomenon demonstrated in sclerotic astrocytes. 83 It is suggested that such Ca2+ release leads to glutamate release from astrocytes, 84 but this remains controversial. An elevated flip-to-flop mRNA ratio of the GluR1 subunit of the AMPA type receptor is also reported 85, 86 suggesting an increased responsiveness to glutamate. Patch- and voltage-clamp studies demonstrate increased density of membrane sodium channels. 87–89 There is also strong upregulation of the α1C subunit of the voltage dependent calcium channel. 90 The α1C subunit contributes to L-type calcium currents, suggesting a change in the current characteristic of these cells. The Kir channel on astrocytes, which normally helps astrocytes remove extracellular K+, is also defective in astrocytes from sclerotic tissue. 91, 92 A large proportion (~60%) of astrocytes in primary cultures from sclerotic seizure foci are capable of generating action potential-like responses when depolarized 89 compared to those from non-sclerotic foci. This ability to produce action potentials may be related to their more depolarized resting membrane potential (~ −55 mV) and increased Na+ channel densities.

Astrocytes possess some key enzymes not found in neurons. Astrocytes from sclerotic foci show changes in these enzymes as well. Most prominent is the down regulation of glutamine synthetase in astrocytes in the CA fields but not in the subiculum of the sclerotic hippocampus. 93, 94 Glutamine synthetase catalyzes the conversion of glutamate, removed from the extracellular space, into glutamine in a process that utilizes ammonia. Indeed, astrocytes from a sclerotic hippocampus seem to have a reduced capacity for glutamine synthesis and ammonia detoxification, 95 and cellular glutamine 96, 97 is reduced in the sclerotic hippocampus. Additionally, glutamate dehydrogenase (GDH) activity is increased while aspartate amino transferase (AAT) activity is reduced in the sclerotic hippocampus. 98 GDH catalyzes the inter-conversion of glutamate to α-ketoglutarate. Lactate dehydrogenase (LDH) activity levels, normalized to citrate synthase activity levels, are also decreased in the sclerotic hippocampus. 98

The distribution of some transporter molecules on the astrocyte plasma membrane is changed in sclerotic hippocampi. Aquaporin 4 (AQP4) is a water transporter molecule on astrocyte membranes. These molecules are more densely located on the perivascular end feet than on the rest of the cell. In sclerotic hippocampi the density on the perivascular end feet is reduced, but unchanged on the rest of the cell membrane. 99 The glutamate transporters EAAT1 and EAAT2 are reported as down regulated by some 100, 101 but not by others. 102 The degree of change does not seems to be an inadequate explanation of the high extracellular glutamate levels observed in sclerotic seizure foci. 103 The expression of the γ-amino butyric acid (GABA) transporter GAT3 is increased on astrocytes in the sclerotic hippocampus. 104 The GAT3 expression is confined to cells resembling protoplasmic astrocytes, which are located in regions of relative neuronal sparing such as the dentate gyrus and hilus. The increased expression of GAT3 may contribute to the increase in removal of GABA and thus reduced extracellular levels during the ictal state.105

Astrocytes and the microvasculature

It was reported over 100 years ago that there is a proliferation of the microvasculature in the sclerotic hippocampus 6 and since confirmed. 106, 107 It is also reported that the blood-brain-barrier may become leaky during a seizure, releasing substances such as albumin from the blood to the brain. 108 Such albumen may, by binding to TGFβ-receptors on astrocytes, trigger transcriptional activation of downstream pathways resulting in down-regulation of inward rectifying potassium channels (Kir 4.1), increased inflammatory responses and reduced inhibitory transmission. 109, 110

Several changes in the expression of molecules at the BBB-astrocyte interface are also observed. The erythropoietin receptor (EPO-r) shows increased expression on the capillaries and perivascular astrocytic end feet of sclerotic hippocampi, particularly in regions of neuronal loss and gliosis (CA3 and CA1). 106 Likewise, the multiple drug resistance gene (MDR1) mRNA expression and the protein it encodes, P-glycoprotein, shows increased expression in the capillary endothelial cells in a majority of patients with intractable TLE. 111 Among the other molecules located on the astrocytic end feet, AQP4 and dystrophin are decreased in expression, whereas CD44 and Plectin 1 expression are increased. 112 These molecular changes suggest that there are functional changes in the BBB of sclerotic hippocampi. The significance of these changes for epileptogensis are poorly understood.

Astrocyte Types

In the sclerotic areas of the seizure foci, while many of cells are strongly GFAP-positive reactive astrocytes, accumulating evidence suggests that the astrocytes may not constitute a homogeneous population. Two subtypes of cells have been distinguished: a GluR cell which is characterized by the presence of AMPA type GluR receptors but does not express glutamate transporters, and a GluT cell which expresses glutamate transporters but does not express glutamate receptors. 113 These cells are recognized in cultures and in slices from hippocampi, and are non-overlapping populations. The GluR cells lack gap junctions while the GluT cells are extensively coupled. 114 GluR cells seem identical with a cell type described as the Neuron-Glial cell (NG2 cell) or polydendrocyte. 115 In the sclerotic hippocampus, several lines of evidence suggest that in addition to GFAP positive astrocytes there is also a population of NG2 cells 80 These NG2 cells appear to be those that have increased calcium oscillations and calcium waves, and are also those that can be depolarized to generate action potentials. 89 It is reported that there is an almost complete loss of GluT cells in the sclerotic hippocampus, with only the GluR cells remaining, and that it is these cells that had the flip isoform of GluR1. 85, 91 There is, however, a population of GFAP positive astrocytes in the sclerotic hippocampus that show down regulation of glutamine synthetase and defective Kir channels. 95

Probable Roles of Astrocytes in a Seizure Focus

The foregoing changes in astrocytes at the sclerotic hippocampal seizure focus may influence epileptogenesis by contributing to high extracellular glutamate levels, to high extracellular K+ levels, and/or to the spread of excitation through the hippocampus.

The down regulation of the enzyme glutamine synthetase may produce an increase in extracellular glutamate due to inadequate clearance of the glutamate released at synapses, 93 while the increase in GDH may produce an increase in the observed cellular glutamate. 116 Further, evidence of intracellular Ca2+ release and Ca2+ oscillations in NG2-like cells in the sclerotic focus (see above), along with evidence of up-regulation of the synaptic vesicle protein SNAP 23 112, suggests the possibility that the sclerotic hippocampus may have cells capable of Ca2+ - dependent exocytotic glutamate release. 117 Additionally, astrocytic swelling by accumulation of water - due to inadequate clearance because of changes in the membrane distribution of AQP4 - may also facilitate astrocyte release of glutamate. 118

A diminished capacity to buffer K+ in the sclerotic hippocampus 119 may result from impaired Kir channels. 92 Since the buffering of K+ by the Kir channel depends on a parallel flux of water through the plasma membranes of these cells, the loss of AQP4 from the perivascular astrocytic membrane could result in diminished extrusion of water from astrocytes and thus concentration of extracellular K+.

The probable presence of a population of NG2 like cells in the sclerotic hippocampus, in addition to their role in Ca2+ dependent glutamate release, may directly contribute to the excitability of the seizure focus. They may contribute to the excitability of the seizure focus because these cells are capable of being depolarized to generate action potentials 89 and have GluR1 receptors with elevated flip to flop ratios which facilitate and prolong depolarization. 85 Such cells may facilitate the generation or spread of waves of depolarization from hyperexcitabe granule cells 120 to the subiculum without synaptic pathways between the two regions. 80 Alternatively, the NG2 cell action potentials may just mediate Na+ ion influx, to increase astrocytic [Na+]I that stimulates the activity of Na/K ATPase. An increase in Na/K ATPase in astrocytes may play a role in buffering extracellular K+, compensating to some degree for the decreased Kir function in these astrocytes and so may decrease excitability. The function of NG2 cells in the sclerotic hippocampus needs further study.

GENE EXPRESSION IN SEIZURE FOCUS

In recent years, several high throughput gene expression analyses of hippocampal tissue from TLE patients have been reported. 112, 121, 122 The most interesting of these studies are those that have compared the expression patterns in sclerotic hippocampi with those of non-sclerotic hippocampi. Unbiased hierarchical cluster analysis of expression data reveals that the gene expression patterns of sclerotic hippocampi from TLE patients usually cluster closely together suggesting that they have molecular characteristics distinguishable from other TLE subtypes (Figure 2). 112, 122 While there is variability among studies on the specific genes that are selectively expressed, there is considerable agreement on some of the functional groups of genes that are differentially expressed in the sclerotic hippocampus. In our own laboratory, we compared the CA1 region from TLE patients with hippocampal sclerosis (medial temporal lobe epilepsy with sclerosis, MTLE) with those from two groups of patients without sclerosis (MaTLE and PTLE, see 7). Comparison of this study with other studies is published elsewhere. 112 On examination of the genes selectively expressed in all MTLE hippocampi, increased expression of genes belonging to three important categories are recognized – astrocyte associated genes, immune and inflammatory response genes, and genes associated with the endothelial cell-astrocyte interface.

Figure 2. The comparison of the gene expression patterns in 20 hippocampi from patients with TLE.

Figure 2

The comparison of the gene expression patterns in 20 hippocampi from patients with TLE. The dendrogram was obtained by unsupervised hierarchical cluster analysis using Gene Spring software. The analysis included all genes expressed in a microarray analysis (more...)

Among the astrocyte related genes upregulated in MTLE are: (1) those associated with astrocyte morphology -- GFAP is associated with astrocyte process extension, paladin (KIAA0992) regulates astrocyte shape, and plectin 1 (PLEC1) serves as a structural protein. The genes ezrin, radixin and moesin control three closely related proteins (ERM proteins), which constitute the machinery for association of actin filaments to the plasma membrane. (2) the CD44 family of surface glycoproteins whose cytoplasmic domain is directly associated with the ERM proteins. (3) the extracellular matrix proteins tenascin (TNC) and chondroitin sulphate phosphoglycan 2 (CSPG2). (4) S100A6 and S100B - members of a family of proteins involved in Ca2+ regulation with AHNAK being a target for S100B.

In addition to the molecules associated with the microvasculature described above (see above, “astrocytes and vasculature”), the ligands CCL3 and CCL2 for the chemokine receptors CCR1 and CCR2 found on the microvasculature 123 are increased in MTLE. CCL4 is also upregulated on astrocytes. 112, 122 The binding of these ligands to their receptors on blood vessels may influence their permeability and perhaps leakage during seizures.

Several genes involved with immune and inflammatory responses are also upregulated. These include those regulating chemokines and their receptors, cytokines and their receptors, signal transduction proteins, transcription factors, transcription factor related genes, and class II major histocompatibility complex genes. The particular genes belonging to these categories are listed in Lee et al. 112 Many of these immune and inflammatory factors are probably produced by astrocytes. 124, 125 Indeed, immunohistochemical localization of interleukin (IL)-1β and IL-1R in sclerotic hippocampi reveals their expression on astrocytes in areas of prominent gliosis and neuronal loss (CA1 – CA3 and hilus) particularly in perivascular end feet. 126 Likewise, immunohistochemical localization of complement C1q, C3c, C3d, and C5b – C9 show that C1q, C3c and C3d are localized in astrocyte-like cells only in sclerotic hippocampi. C5 – C9 are not on astrocytes. Experimental studies in animals indicate that molecules such as IL-1β and complement are closely associated with seizure generation. 127

SUBICULUM

The subiculum is the region of the hippocampal formation from which major efferent pathways originate 128 and is involved in the spread of electrical activity from the hippocampus to other parts of the brain generating behavioral activity. Even in sclerotic hippocampi in MTLE patients, the subiculum remains largely intact with no detectable neuron loss. 8, 129, 130 However, Cohen et al. 131 report that the subiculum of the sclerotic hippocampus initiates spontaneous interictal discharges that originate in a minority of subicular neurons, including interneurons and a subset of GABAergic pyramidal neurons in the subiculum and its transitional area with CA1. This paradoxical behavior of GABAergic neurons, which are normally inhibitory but behave as excitatory cells in the subiculum, has found an explanation in recent research. This switch of GABAergic neurons from inhibitory to excitatory relates to relative expression of two Cl cotransporters – Na+K+, 2Cl cotransporters such as NKCC1, which accumulate Cl in the cell, and Cl extruding K+, Cl co-transporters like KCC2. Alterations in the balance of NKCC1 and KCC2 determine the switch from hyperpolarizing to depolarizing effects of GABA. Immunohistochemical localization and quantitative RT-PCR analysis mRNA extracted from subiculum have suggested that greater than 20% of NKCC1-expressing pyramidal neurons in the sclerotic subicular/CA1 transitional zone not only lacked KCC2 but exhibited increased NKCC1. This pattern contrasts with about 95% co-localization in non-sclerotic patients. 132–134 Increased Cl in cells produces depolarization when GABA-mediated chloride channels are opened. Thus, it is suggested that altered Cl transport renders GABA neurons aberrantly excitable, contributing to the precipitation of seizures. 134 However, Wozny and coworkers 135 provide evidence, also obtained from slices of surgically removed hippocampal tissue, that interictal spontaneous activity is seen not only in sclerotic hippocampi but also non-sclerotic hippocampi. Thus, the subiculum of the sclerotic hippocampus may not be involved in a epileptogenesis uniquely in sclerotic hippocampus. this region of the hippocampal formation may be a more generally excitable zone than other regions.

ENTORHINAL CORTEX

The entorhinal cortex (EC) is a principal source of efferents to the hippocampus, with major projection to the dentate gyrus (perforant pathway) and CA1 (temporo-ammonic pathway). Abnormal epileptiform activity has been recorded from this region. 136, 137 Histopathological studies of this region in TLE patients report neuronal loss and gliosis in layer III and, to a lesser extent in layer II of the anterior portion of the medial entorhinal cortex of the sclerotic hippocampus. 138 However, other studies have reported a more variable pattern of neuronal loss 139 or no detectable loss. 140 Nevertheless, gliosis is a finding common to the EC in both sclerotic and non-sclerotic hippocampus. 140 Primary cultures from the EC (parahippocampus) of patients with a sclerotic hippocampus had increased Ca2+ release and Ca2+ oscillations in astrocytes. 83 If Ca2+ release is associated with glutamate release, these cells may trigger hyperexcitation of EC neurons and thus spread activity through the hippocampus. More needs to be learned about the role of these glial cells in the EC region.

PROBABLE PATHOPHYSIOLOGICAL MECHANISMS OF SEIZURE GENERATION

The probable pathophysiological mechanism of seizure generation must take into account the complex histopathological picture sketched above. Prominent features of this histopathology – a reorganized dentate gyrus, a neuronally depleted and gliotic Ammon’s horn with an essentially intact and even spontaneously active subiculum -- constitute the epileptogenic hippocampus.

Broadly, two sets of processes that have distinct histological substrates may be identified. The first set of processes are based in the dentate gyrus and may be important at least in the early genesis of the epileptic focus, before the loss of dentate granule cells is extensive. In this presumed early phase of pathogenesis, several cellular changes may underlie the development of hyper-excitable granule cells. These processes involve morphological remodeling of surviving granule cells, selective loss of hilar interneurons, sprouting of new axonal projections, and adaptive changes in neurotransmitter receptor distribution, representing a reorganization of the neural circuits within the dentate gyrus. An attempt to integrate these changes and hypothesize how they may result in granule cells hyperexcitability is given in de Lanerolle & Lee (22, see Figure 1).

Despite the existence of such hyperexcitable granule cells, unanswered is the question of how such hyperexcitability passes from the granule cells to the subiculum to generate behavioral seizures, since the neural pathways connecting the two are interrupted due to extensive neuronal loss in the CA1 area early in pathogenesis. One hypothesis is that in the sclerotic region NG2 like cells, which are depolarizable and excitable, provide a substrate for the spread of waves of depolarization from the dentate gyrus to the subiculum. An alternative though less likely possibility is that the neural networks in the stratum oriens, which are very resistant to loss, provide a sufficient substrate for spread of neural excitation from dentate to subiculum.

It is notable that in several surgically removed hippocampi, the dentate gyrus is almost completely devoid of granule cells and Ammon’s horn is very gliotic; yet such hippocampi may take part in seizure generation in these patients. Seizures have been shown to originate from such hippocampi 14 and in vivo dialysis studies demonstrate elevated extracelluar glutamate ictally in such foci. 105, 141 In such hippocampi, the subiculum remains intact; however, the evidence, as discussed above, that the subiculum by itself is responsible for epileptic activity is weak. Alternative explanations must therefore be considered, such as astrocyte and astrocyte-related mechanisms. The unique properties of astrocytes in the hippocampal seizure focus are reviewed above, and an attempt to integrate these properties in the pathophysiology of epilepsy is provided in de Lanerolle et al. 80 Astrocytes may contribute to vascular permeability and increased extracellular glutamate and K+ levels in the sclerotic hippocampus, which in turn may contribute to the hyperexcitability of the subiculum resulting in epileptic activity.

SPECULATION ON FUTURE CHALLENGES IN THE AREA

Our knowledge of the histopathology of temporal lobe epilepsy has advanced significantly over the recent past. However, we are still engaged largely in descriptive pathology. One of the major challenges for the future is to understand the functional significance of the various changes observed. Of primary importance is an understanding of the functional biology of astrocytes and astrocyte-like cells in the sclerotic hippocampus. We have advanced several hypotheses regards their role, but these need experimental testing. What is the proportion of GluR- and GluT-like astrocytes? How do they contribute to the increased extracellular glutamate and K+ concentrations observed at the sclerotic seizure focus? Much of our information of their role in these mechanisms has come from the study of astrocytes from embryonic animal tissue, and very little from the study of adult astrocytes from human foci. Do these abnormal astrocytes provide an excitable substrate for the spread of neural excitation from the dentate to the subiculum? Understanding the biology of astrocytes in TLE sclerotic foci would be very useful in providing insight into the epileptogenic mechanisms of other seizure foci such as mass lesions, tubers in tuberous sclerosis complex, and many of the neocortical epilepsies in which astrocytes may be a major constituent.

Our understanding of the functional implications of the anatomical changes in the dentate gyrus in the human also remains sketchy. Though much attention has been given to certain aspects such as the functional significance of granule cell sprouting and glutamatergic and GABAergic mechanisms regulating granule cell function in the sclerotic hippocampus, there is far less experimental analysis of the significance of other changes such as the new neuropeptidergic inputs to granule cells and the changed distribution of the several receptors. Studies that attempt to correlate the time course of functional changes resulting from dentate gyrus reorganization and abnormal astrocytes in the sclerotic human hippocampus are essential for the better understanding of the causative mechanisms of human TLE.

IMPACT ON FINDING CURES AND REPAIRS FOR EPILEPSIES

The current antiepileptic drugs and other repair strategies have focused attention almost exclusively on neurons as the primary cell type involved in mechanisms underlying epilepsy. The existence of a large group of patients with drug refractory epilepsy is evidence that more effective therapeutic measures are required. The activity of the currently available antiepileptic drugs broadly target glutamatergic and GABAergic mechanisms, though they may have other effects as well. They are mediated through three main mechanisms – blockade of voltage-gated sodium and calcium channels that limit glutamate release; blockade of glutamate receptors, and the potentiation of GABAergic inhibition, through increasing GABA availability or modulation of GABA receptors. 142, 143 The histopathology of the epileptic hippocampus reviewed above, which shows that there is extensive loss of neurons in the hippocampus, suggests why current antiepileptic drugs, which focus on modulating functions of neurons, may be ineffective. New therapeutic agents should target astrocytes and astrocyte-like cells instead of neurons, as these cells are the primary constituents of the sclerotic hippocampus. Cell replacement strategies might consider transplantations of human stem cells that may develop into normal astrocytes that may restore a more normal extracellular glutamate concentration in the seizure focus, and compensate for the abnormal astrocytes.

ACKNOWLEDMENTS

We thank Ms. Ilona Kovacs our histology associate over the past 25 years, whose outstanding expertise has contributed greatly to all of the studies from our laboratory. We also thank our many collaborators who contributed to the studies from our laboratory reported in this paper, especially Drs. Michael Brines, Tore Eid, Jung Kim, Matthew Phillips, Sanjoy Sunderasen, John Tompkins, Subu Magge and Edward O’Connor.

BIBLIOGRAPHY

1.
Kim JH. Pathology of seizure disorders. Neuroimag Clinics of North Am. 1995;5:527–545. [PubMed: 8564282]
2.
Kim JH. Pathology of Epilepsy Exptl Mol Pathol. 2001;70:345–367. [PubMed: 11418013]
3.
Gloor P. Mesial temporal sclerosis: Historical background and an overview from a modern perspective. In: Luders H, editor. Epilepsy Surgery. New York: Raven Press; 1991. pp. 6889–6903.
4.
Sommer W. Erkrankung des Ammonshorns als aetiologisches Moment der Epilepsie. Arch Psychiatr Nervenkr. 1880;10:631–675.
5.
Bouchet C. De l’épilepsie considérée dans ses rapports avec l’aliénation mentale. Recherche sur la nature et le siége de ces deux maladies. Arch Gén Méd. 1825;9:510–542.
6.
Bratz E. Ammonshornbefunde bei Epileptikern. Arch Psychiatr Nervenkr. 1899;32:820–835.
7.
de Lanerolle NC, Kim JH, Williamson A, Spencer SS, Zaveri HP, Eid T, Spencer DD. A Retrospective Analysis of Hippocampal Pathology in Human Temporal Lobe Epilepsy: Evidence for Distinctive Patient Subcategories. Epilepsia. 2003;44:677–687. [PubMed: 12752467]
8.
Kim JH, Guimaraes PO, Shen M-Y, Masukawa LM, Spencer DD. Hippocampal neuronal density in temporal lobe epilepsy with and without gliomas. Acta Neuropathol. 1990;80:41–45. [PubMed: 2360416]
9.
Engel J, Brown WJ, Kuhl DE, Phelps ME, Mazziota JC, Crandall PH. Pathological findings underlying focal temporal lobe hypometabolism in partial epilepsy. Ann Neurol. 1982;12:518–528. [PubMed: 6984318]
10.
Dlugos DJ, Jaggi J, O’Connor WM, Ding XS, Reivich M, O’Connor MJ, Sperling MR. Hippocampal cell density and subcortical metabolism in temporal lobe epilepsy. Epilepsia. 1999;40:408–413. [PubMed: 10219265]
11.
Spencer SS. Neural networks in human epilepsy: evidence of and implications for treatment. Epilepsia. 2002;43:219–227. [PubMed: 11906505]
12.
Stauder KH. Epilepsie und schläfenlappen. Arch Psychiatr. 1935;104:181–212.
13.
Falconer MA. The significance of mesial temporal sclerosis (Ammon’s horn sclerosis) in epilepsy. Guy’s Hosp Reports. 1968;117:1–12. [PubMed: 5671984]
14.
Babb TL, Lieb JP, Brown WJ, Pretorius J, Crandall PH. Distribution of pyramidal cell density and hyperexcitability in the epileptic human hippocampal formation. Epilepsia. 1984;25:721–728. [PubMed: 6510380]
15.
Spencer SS, Berg AT, Vickrey BG, Sperling MR, Bazil CW, Shinnar S, Langfitt JT, Walczak TS, Pacia SV. Surgery TMSoE. Predicting long-term seizure outcome after resective epilepsy surgery: The Multicenter Study. Neurology. 2005;65:912–918. [PubMed: 16186534]
16.
Bower SPC, Kilpatrick CJ, Vogrin SJ, Morris K, Cook MJ. Degree of hippocampal atrophy is not related to a history of febrile seizures in patients with proved hippocampal sclerosis. J Neurol Neurosurg Psyciatr. 2000;69:733–738. [PMC free article: PMC1737179] [PubMed: 11080224]
17.
Kothare SV, Sotrel A, Duchowny M, Jayakar P, Marshall PC, Smith TW. Absence of hippocampal sclerosis in children with multiple daily seizures since infancy. J Child Neurol. 2001;16:562–564. [PubMed: 11510925]
18.
Thom M, Zhou J, Martinian L, Sisodiya SM. Quantitative post-mortem study of the hippocampus in chronic epilepsy: seizures do not invariably cause neuronal loss. Brain. 2005:1–14. [PubMed: 15758032] [CrossRef]
19.
Falconer MA. Surgery of temporal lobe epilepsy. Proc Royal Soc Med. 1952;51:613–616. [PubMed: 13578962]
20.
Spencer DD, Spencer SS. Surgery for epilepsy. Neurol Clin. 1985;3:313–330. [PubMed: 3927130]
21.
Blümcke I, Kistner I, Clusmann H, Schramm J, Becker AJ, Elger CF, Bien CG, Merschhemke M, Meencke H-J, Lehmann TN, Buchfelder M, Weigel D, Buslei R, Stefan H, Pauli E, Hildebrandt M. Towards a clinico-pathological classification of granule cell dispersion in human mesial temporal lobe epilepsies. Acta Neuropathol. 2009;117:535–544. [PubMed: 19277686]
22.
de Lanerolle NC, Lee TS. New facets of the neuropathology and molecular profile of human temporal lobe epilepsy. Epilepsy & Behavior. 2005;7:190–203. [PubMed: 16098816]
23.
Margerison JH, Corsellis JAN. Epilepsy and the temporal lobes. Brain. 1966;89:499–530. [PubMed: 5922048]
24.
Bruton CJ. The Neuropathology of Temporal Lobe Epilepsy. Oxford; Oxford University Press; 1988.
25.
de Lanerolle NC, Kim JH, Brines ML. Cellular and molecular alterations in partial epilepsy. Clin Neurosci. 1994;2:64–81.
26.
Zaveri HP, Duckrow RB, de Lanerolle NC, Spencer SS. Distinguishing subtypes of temporal lobe epilepsy with background hippocampal activity. Epilepsia. 2001;42:725–730. [PubMed: 11422326]
27.
Amaral DG, Scharfman HE, Lavenex P. The dentate gyrus: fundamental neuroanatomical organization (dentate gyrus for dummies). Progress in Brain Res. 2007;163:3–25. [PMC free article: PMC2492885] [PubMed: 17765709]
28.
Seress L, Mrzljak L. Basal dendrites of granule cells are normal features of the fetal and adult dentate gyrus of both monkey and human hippocampal formations. Brain Res. 1987;405:169–174. [PubMed: 3567591]
29.
de Lanerolle NC, Brines ML, Williamson A, Kim JH, Spencer DD. Neurotransmitters and their receptors in human temporal lobe epilepsy. In: Ribak CE, Gall CM, Mody I, editors. The Dentate Gyrus and Its Role in Seizures. Amsterdam: Elsevier Science Publishers B.V; 1992. pp. 235–250. [PubMed: 1361331]
30.
Haas CA, Dudeck O, Kirsch M, Huszka C, Kann G, Pollak S, Zentner J, Frotscher M. Role for reelin in the development of granule cells dispersion in temporal lobe epilepsy. J Neurosci. 2002;22:5797–5802. [PMC free article: PMC6757930] [PubMed: 12122039]
31.
Fahrner A, Kann G, Flubacher A, Heinrich C, Freiman TM, Zentner J, Frotscher M, Haas CA. Granule cell dispersion is not accompanied by enhanced neurogenesis in temporal lobe epilepsy patients. Exptl Neurol. 2007;203:320–332. [PubMed: 17049346]
32.
Harding B, Thom M. Bilateral hippocampal granule cell dispersion: autopsy study of 3 infants. Neuropathol Appl Neurobiol. 2001;27:245–251. [PubMed: 11489144]
33.
Robboli MG, Giulioni M. Role of dentate gyrus alterations in mesial temporal sclerosis. Clin Neuropathol. 2010;29:32–35. [PubMed: 20040331]
34.
von Campe G, Rai P, Spencer DD, de Lanerolle NC. Morphology of granular neurons in the human hippocampus. Soc Neurosci Abst. 1994;20:1450.
35.
Isokawa M, Levesque MF, Babb TL, Engel JJ. Single mossy fiber axonal systems of human dentate granule cells studied in hippocampal slices from patients with temporal lobe epilepsy. J Neurosci. 1993;13:1511–1522. [PMC free article: PMC6576742] [PubMed: 8463831]
36.
Isokawa M. Preservation of dendrites with the presence of reorganized mossy fiber collaterals in hippocampal dentate granule cells in patients with temporal lobe epilepsy. Brain Res. 1997;744:339–343. [PubMed: 9027394]
37.
Pirker S, Gasser E, Czech T, Baumgartner C, Schuh E, Feucht M, Novak K, Zimprich F, Sperk G. Dynamic up-regulation of prodynorphin transcription in temporal lobe epilepsy. Hippocampus. 2009;19:1051–1054. [PMC free article: PMC3073604] [PubMed: 19437412]
38.
Sloviter RS, Sollas AL, Barbaro NM, Laxer KD. Calcium-binding protein (Calbindin-D28K) and parvalbumin immunocytochemistry in the normal and epileptic human hippocampus. J Comp Neurol. 1991;308:381–396. [PubMed: 1865007]
39.
ender RA, Soleymani SV, Brewster AL, Nguyen ST, Beck H, Mathern GW, Baram TZ. Enhanced expression of a specific hyperpolarization-activated cyclic nucleotide-gated cation channel (HCN) in surviving dentate gyrus granule cells of human and experimental epileptic hippocampus. J Neurosci. 2003;30:6826–6826. [PMC free article: PMC3100807] [PubMed: 12890777]
40.
de Lanerolle NC, Kim JH, Robbins RJ, Spencer DD. Hippocampal interneuron loss and plasticity in human temporal lobe epilepsy. Brain Res. 1989;495:387–395. [PubMed: 2569920]
41.
Babb TL, Pretorius JK, Crandall PH. Glutamate decarboxylase-immunoreactive neurons are preserved in human epileptic hippocampus. J Neurosci. 1989;9:2562–2574. [PMC free article: PMC6569780] [PubMed: 2501460]
42.
Sloviter RS. Decreased hippocampal inhibition and a selective loss of interneurons in experimental epilepsy. Science. 1987;235:73–76. [PubMed: 2879352]
43.
Amaral DG. A Golgi study of cell types in the hilar region of the hippocampus in the rat. J Comp Neurol. 1978;182:851–914. [PubMed: 730852]
44.
Soriano E, Frotscher M. Mossy cells of the rat fascia dentata are glutamate-immunoreactive. Hippocampus. 1994;4:65–69. [PubMed: 7914798]
45.
Sloviter RS. Permanently altered hippocampal structure, excitability, and inhibition after experimental status epilepticus in the rat: The “dormant basket cell” hypothesis and its possible relevance to temporal lobe epilepsy. Hippocampus. 1991;1:41–66. [PubMed: 1688284]
46.
Seress L. Comparative anatomy of the hippocampal dentate gyrus in adult and developing rodents, non-human primates and humans. Progress in Brain Res. 2007;163:23–41. [PubMed: 17765710]
47.
de Lanerolle NC, Williamson A, Meredith C, Kim JH, Tabuteau H, Spencer DD, Brines ML. Dynorphin and the kappa 1 ligand [3H]-U69,539 binding in the human epileptogenic hippocampus. Epilepsy Res. 1997;28:189–205. [PubMed: 9332884]
48.
Eid T, Meredith C, Kovacs I, de Lanerolle NC. Ultrastructural localization of GluR1 in human temporal lobe epilepsy. Soc Neurosci Abstr. 1996;22:1657.
49.
Seress L, Abraham H, Horvath Z, Dóczi T, Janszky J, Klemm J, Byrne R, Bakay RAE. Survival of mossy cells of the hippocampal dentate gyrus in humans with mesial temporal lobe epilepsy. J Neurosurg. 2009;111:1237–1247. [PubMed: 19392605]
50.
Babb TL, Kupfer WR, Pretorius JK, Levesque MF. Light and electron microscopy of mossy fiber terminals in human “epileptic” fascia dentata. Soc Neurosci Abstr. 1989;14:881.
51.
Sutula T, Cascino G, Cavazos J, Pavada I, Ramirez L. Mossy fiber synaptic reorganization in the epileptic human temporal lobe. Ann Neurol. 1989;26:321–330. [PubMed: 2508534]
52.
Houser CR, Miyashiro JE, Swartz BE, Walsh GO, Rich JR, Delgado-Escueta AV. Altered patterns of dynorphin immunoreactivity suggest mossy fiber reorganization in human hippocampal epilepsy. J Neurosci. 1990;10:267–282. [PMC free article: PMC6570337] [PubMed: 1688934]
53.
Philips MW. Synaptic and Pathway Remodeling of the Human Hippocampus in Temporal Lobe Epilepsy [MD]. Yale University; 1993.
54.
Green R, Blume H, Kupperschmid S, Mesulam MM. Alterations of hippocampal acetylcholinesterase in human temporal lobe epilepsy. Ann Neurol. 1989;26:351–367. [PubMed: 2802534]
55.
de Lanerolle NC, Eid T, von Campe G, Kovacs I, Spencer DD, Brines ML. Glutamate receptor subunits GluR1 and GluR2/3 distribution shows reorganization in the human epileptogenic hippocampus. Eur J Neurosci. 1998;10:1687–1703. [PubMed: 9751141]
56.
Eid T, Kovacs I, Spencer DD, de Lanerolle NC. Novel expression of AMPA-receptor subunit GluR1 on mossy cells and CA3 pyramidal neurons in the human epileptogenic hippocampus. Eur J Neurosci. 2002;15:517–527. [PubMed: 11876779]
57.
Notenboom RGE, Hampson DR, Jansen GH, van Rijen PC, van Veelen CWM, Nieuwenhuizen O, de Grann PN. Up-regulation of hippocampal metabotropic glutamate receptor 5 in temporal lobe epilepsy patients. Brain. 2006;129:96–107. [PubMed: 16311265]
58.
Tang F-R, Lee WL. Expression of the group ii and III metabotropic glutamate receptors in the hippocampus of patients with mesial temporal lobe epilepsy. J Neuroscytol. 2001;30:137–143. [PubMed: 11577252]
59.
Tang F-R, Lee W-L, Yeo TT. Expression of group 1 metabotropic glutamate receptor in the hippocampus of patients with mesial temporal lobe epilepsy. J Neurocytol. 2001;30:403–411. [PubMed: 11951051]
60.
Moldrich RX, Chapman AG, De Sarro G, Meldrum BS. Glutamate metabotropic receptors as targets for drug therapy. Eur J Pharmacol. 2003;476:3–16. [PubMed: 12969743]
61.
Loup F, Weiser H-G, Yonekawa Y, Aguzzi A, Fritschy JM. Selective alterations in GABA-A receptor subtypes in human temporal lobe epilepsy. J Neurosci. 2000;20:5401–5419. [PMC free article: PMC6772330] [PubMed: 10884325]
62.
Billington A, Baird VH, Thom M, Duncan JS, Upton N, Bowery NG. GABA-B receptor autoradiography in hippocampal sclerosis associated with human temporal lobe epilepsy. Brit J Pharmacol. 2001;132:475–480. [PMC free article: PMC1572589] [PubMed: 11159697]
63.
rincivalle AP, Duncan JS, Thom M, Bowery NG. GABA-B1a, GABA-B1b and GABA-B2 m RNA variants expression in hippocampus resected from patients with temporal lobe epilepsy. Neurosci. 2003;122:975–984. [PubMed: 14643764]
64.
Csaba Z, Pirker S, Lelouvier B, Simon A, Videau C, Epelbaum J, Czech T, Baumgartner C, Sperk G, Dournaud P. Somatostatin receptor TYpe 2 undergoes plastic changes in the human epileptic dentate gyrus. J Neuropathol Exp Neurol. 2005;64:956–969. [PubMed: 16254490]
65.
Sundaresan S. Neurochemical changes in the hippocampus in human temporal lobe epilepsy [MD]. New Haven: Yale University; 1990.
66.
Furtinger S, Pirker S, Czech T, Baumgartner C, Ransmayr G, Sperk G. Plasticity of Y1 and Y2 receptors and neuropeptide Y fibers in patients with temporal lobe epilepsy. J Neurosci. 2001;21:5804–5812. [PMC free article: PMC6762672] [PubMed: 11466452]
67.
Tompkins J. Alterations in D2 receptor concentrations in human temporal lobe epilepsy [MD]. New Haven: Yale University; 1990.
68.
de Lanerolle NC, Gunel M, Sundaresan S, Shen MY, Brines ML, Spencer DD. Vasoactive intestinal polypeptide and its receptor changes in human temporal lobe epilepsy. 1995;686:182–193. [PubMed: 7583284]
69.
Patrylo PR, Van den Pol AN, Spencer DD, Williamson A. NPY inhibits glutamatergic excitation in the epileptic human dentate gyrus. J Neurophysiol. 1999;82:478–483. [PubMed: 10400974]
70.
Dudek FE, Sutula TP. Epileptogenesis in the dentate gyrus: a critical perspective. Progress in Brain Res. 2007;163:755–773. [PubMed: 17765749]
71.
Williamson A, Patrylo PR, Spencer DD. Decrease in inhibition in dentate granule cells from patients with medial temporal lobe epilepsy. Ann Neurol. 1999;45:92–99. [PubMed: 9894882]
72.
Corsellis JAN. The incidence of Ammon’s horn sclerosis. Brain. 1955;80:193–208. [PubMed: 13446308]
73.
Falconer M. Mesial temporal (Ammon’s horn) sclerosis as a common cause of epilepsy. Aetiology, treatment and prevention. Lancet. 1974;2:767–770. [PubMed: 4143026]
74.
de Lanerolle NC, Brines ML, Kim JH, Williamson A, Philips MF, Spencer DD. Neurochemical remodelling of the hippocampus in human temporal lobe epilepsy. In: Engel J Jr, Wasterlain C, Cavalheiro EA, Heinemann U, Avanzini G, editors. Molecular Neurobiology of Epilepsy. Amsterdam: Elsevier Science Publishers B.V.; 1992. pp. 205–220. [PubMed: 1363041]
75.
de Lanerolle NC, LaMotte CC. The morphological relationships between substance P immunoreactive processes and ventral horn neurons in the human and monkey spinal cord. J Comp Neurol. 1982;207:305–313. [PubMed: 6181100]
76.
Eid T. The use of hydrophobic adhesive tape to produce miniature wells on microscope slides. Biotech Histochem. 1993;68:189–192. [PubMed: 8218571]
77.
Ganter P, Szucs P, Paulsen O, Somogyi P. Properties of horizontal axo-axonic cells in the stratum oriens of the hippocampal CA1 areaof rats in vitro. Hippocampus. 2004;14:232–243. [PubMed: 15098728]
78.
Maccaferri G. Stratum oriens horizontal interneurone diversity and hippocampal network dynamics. J Physiol. 2005;562:73–80. [PMC free article: PMC1665470] [PubMed: 15498801]
79.
Aradi I, Maccaferri G. Cell-type specific synsptic dynamics of synchrinized bursting in the juvenile CA3 rat hippocampus. J Neurosci. 2004;24:9681–9682. [PMC free article: PMC6730161] [PubMed: 15509756]
80.
de Lanerolle NC, Lee TS, Spencer DD. Astrocytes and epilepsy. Neurotherapeutics. 2010;7:424–438. [PMC free article: PMC5084304] [PubMed: 20880506]
81.
Steinhäuser C, Haydon PG, de Lanerolle NC. Astroglial mechanisms in epilepsy. In: Engel JJ, Pedley TA, Aicardi J, Dichter M, Moshe S, editors. Epilepsy: A comprehensive textbook. Philadelphia: Lipincott, Williams & Wilkins; 2008. pp. 277–288.
82.
Steinhäuser C, Seifert G. Glial membrane channels and receptors in epilepsy: impact for generation and spread of seizure activity. Eur J Pharmacol. 2002;447:227–237. [PubMed: 12151014]
83.
Lee SH, Magge S, Spencer DD, Sontheimer H, Cornell-Bell A. Human epileptic astrocytes exhibit increased gap junction coupling. Glia. 1995;15:195–202. [PubMed: 8567071]
84.
Volterra A, Meldolesi J. Astrocytes, from brain glue to communication elements: the revolution continues. Nat. rev. Neurosci. 2005;6:626–640. [PubMed: 16025096]
85.
Seifert G, Hüttmann K, Schramm J, Steinhäuser C. Enhanced relative expression of glutamate receptor 1 flip AMPA receptor subunits in hippocampal astrocytes of epilepsy patients with Ammon’s horn sclerosis. J Neurosci. 2004;24:1996–2003. [PMC free article: PMC6730392] [PubMed: 14985442]
86.
Seifert G, Schroder W, Hinterkeuser S, Schumacher T, Schramm J, Steinhauser C. Changes in flip/flop splicing of astroglial AMPA receptors in human temporal lobe epilepsy. Epilepsia. 2002;43(Suppl 5):162–167. [PubMed: 12121314]
87.
Barres BA, Chun LLY, Corey DP. Ion channel expression by white matter glia: I. Type 2 astrocytes and oligodendrocytes. Glia. 1988;1:10–30. [PubMed: 2466789]
88.
Bordey A, Spencer DD. Distinct electrophysiological alterations in dentate gyrus versus CA1glial cells from epileptic humans with temporal lobe sclerosis. Epilepsy Res. 2004;59:107–122. [PubMed: 15246115]
89.
O’Connor ER, Sontheimer H, Spencer DD, de Lanerolle NC. Astrocytes from human hippocampal epileptogenic foci exhibit action potential-like responses. Epilepsia. 1998;39:347–354. [PubMed: 9578024]
90.
Djamshidian A, Grassl R, Seltenhammer M, Czech T, Baumgartner C, Schmidbauer M, Ulrich W, Zimprich F. Altered expression of voltage-dependent calcium channel a1 subunits in temporal lobe epilepsy with Ammon’s horn sclerosis. Neuroscience. 2002;111:57–69. [PubMed: 11955712]
91.
Hinterkeuser S, Schröder W, Hager G, Seifert G, Blümcke I, Elger CE, Schramm J, Steinhauser C. Astrocytes in the hippocampus of patients with temporal lobe epilepsy display changes in potassium conductances. Eur J Neurosci. 2000;12:2087–2096. [PubMed: 10886348]
92.
Schroder W, Hinterkeuser S, Seifert G, Schramm J, Jabs R, Wilkin GP, Steinhauser C. Functional and molecular properties of human astrocytes in acute hippocampal slices obtained from patients with temporal lobe epilepsy. Epilepsia. 2000;41:S181–184. [PubMed: 10999541]
93.
Eid T, Thomas MJ, Spencer DD, Rundén-Pran E, Kim JH, Ottersen OP, de Lanerolle NC. Loss of glutamine synthetase in the human epileptogenic hippocampus: a possible mechanism for elevated exttracellular glutamate in mesial temporal lobe epilepsy. Lancet. 2004;363:28–37. [PubMed: 14723991]
94.
van der Hel WS, Notenboom RG, Bos IW, van Rijen PC, van Veelen CW, de Grann PN. Reduced glutamine synthetase in hippocampal areas with neuron loss in temporal lobe epilepsy. Neurology. 2005;64:326–333. [PubMed: 15668432]
95.
Eid T, Williamson A, Lee T-S, Petroff OA, de Lanerolle NC. Glutamate and astrocytes -- Key players in human mesial temporal lobe epilepsy. Epilepsia. 2008;49:42–52. [PubMed: 18226171]
96.
Petroff OA, Errante LD, Kim JH, Spencer DD. N-acetyl-aspartate, total creatine, and myo-inositol in the epileptogenic human hippocampus. Neurology. 2003;60:1646–1651. [PubMed: 12771256]
97.
Petroff OA, Errante LD, Rothman DL, Kim JH, Spencer DD. Neuronal and glial metabolite content of the epileptogenic human hippocampus. Ann Neurol. 2002;52:635–642. [PubMed: 12402262]
98.
Malthankar-Phatak GH, de Lanerolle NC, Eid T, Spencer DD, Behar KL, Spencer SS, Kim JH, Lai JCK. Differential glutamate dehydrogenase (GDH) activity profile in patients with temporal lobe epilepsy. Epilepsia. 2006;47:1292–1299. [PubMed: 16922873]
99.
Eid T, Lee T-SW, Thomas MJ, Amiry-Moghaddam M, Bjornsen LP, Spencer DD, Agre P, Ottersen OP, de Lanerolle NC. Loss of perivascular aquaporin 4 inderlie deficient water and K+ homeostasis in the human epileptogenic hippocampus. Proc Natl Acad Sci U S A. 2005;102:1193–1198. [PMC free article: PMC545857] [PubMed: 15657133]
100.
Mathern GW, Mendoza D, Lozada A, Pretorius JK, Dehnes Y, Danbolt NC, Nelson N, Leite JP, Chimelli L, Born DE, Sakamoto AC, Assirati JA, Fried I, Peacock WJ, Ojemann GA, Adelson PD. Hippocampal GABA and glutamate transporter immunoreactivity in patients with temporal lobe epilepsy. Neurology. 1999;52:453–472. [PubMed: 10025773]
101.
Proper EA, Hioogland G, Kappen SM, Jansen GH, Rensen MGA, Schrama LH, van Veelen CWM, van Rijen PC, van Nieuwenhuizen O, Gispen WH, de Graan PNE. Distribution of glutamate transporters in the hippocampus of patients with pharmaco-resistant temporal lobe epilepsy. Brain. 2002;125:32–43. [PubMed: 11834591]
102.
Tessler S, Danbolt NC, Faull RLM, Storm-mathisen J, Emson P. Expression of the glutamate transporters in human temporal lobe epilepsy. Neuroscience. 1998;88:1083–1091. [PubMed: 10336123]
103.
Bjørnsen LP, Eid T, Holmseth S, Danbolt NC, Spencer DD, de Lanerolle NC. Changes in glial glutamate transporters in human epileptogenic hippocampus: Inadequate explanation for high extracellular glutamate during seizures. Neurobiol Dis. 2006;25:319–330. [PubMed: 17112731]
104.
Lee T-S, Bjornsen LP, Paz C, Kim JH, Spencer SS, Spencer DD, Eid T, de Lanerolle NC. GAT1 and GAT3 expression are differently localized in the human epileptogenic hippocampus. Acta Neuropathol. 2006;111:351–363. [PubMed: 16456667]
105.
During MJ, Spencer DD. Extracellular hippocampal glutamate and spontaneous seizure in the conscious human brain. Lancet. 1993;341:1607–1610. [PubMed: 8099987]
106.
Eid T, Brines M, Cerami A, Spencer DD, Kim JH, Schweitzer JS, Ottersen OP, de Lanerolle NC. Increased expression of erythropoietin receptor on blood vessels in the human epileptogenic hippocampus with sclerosis. J Neuropathol Exptl Neurol. 2004;63:73–83. [PubMed: 14748563]
107.
Rigau V, Morin M, Rousset M-C, de Bock F, Lebrun A, Coubes P, Picot M-C, Baldy-Moulinier M, Bockaert J, Crespel A, Lerner-Natoli M. Angiogenesis is associated with blood-brain barrier permeability in temporal lobe epilepsy. Brain. 2007;130:1942–1956. [PubMed: 17533168]
108.
Van Vliet EA, da Costa Araújo S, Redeker S, van Schaik R, Aronica E. Blood-brain barrier leakage may lead to progression of temporal lobe epilepsy. Brain. 2007;130:521–534. [PubMed: 17124188]
109.
Cacheaux LP, Ivens S, David Y, Lakhter AJ, Bar-Klein G, Shapira M, Heinemann U, Friedman A, Kaufer D. Transcriptome profiling reveals TGF-b signalling involvement in epileptogenesis. J Neurosci. 2009;29:8927–8935. [PMC free article: PMC2875073] [PubMed: 19605630]
110.
Ivens S, Kaufer D, Flores LP, Bechmann I, Zumsteg D, Tompkins O, Seiffert E, Heinemann U, Friedman A. TGF-b receptor-mediated albumin uptake into astrocytes is involved in neocortical epileptogenesis. Brain. 2007;130:535–547. [PubMed: 17121744]
111.
Tishler DM, Weinberg KI, Hinton DR, Barbaro N, Annett GM, Raffel C. MDR1 gene expression in brain of patients with medically intractable epilepsy. Epilepsia. 1995;36:1–6. [PubMed: 8001500]
112.
Lee T-S, Mane S, Eid T, Zhao H, Lin A, Guan Z, Kim JH, Schweitzer J, King-Stevens D, Weber P, Spencer SS, Spencer DD, de Lanerolle NC. Gene expression in temporal lobe epilepsy is consistent with increased release of glutamate by astrocytes. Mol Med. 2007;13:1–13. [PMC free article: PMC1869627] [PubMed: 17515952]
113.
Matthias K, Kirchhoff F, Seifert G, Hüttmann K, Matyash M, Kettenmann H, Steinhäuser C. Segregated expression of AMPA-type glutamate receptors and glutamate transporters defines distinct astrocyte populations in the mouse hippocampus. J Neurosci. 2003;23:1750–1758. [PMC free article: PMC6741945] [PubMed: 12629179]
114.
Wallraff A, Odermatt B, Willecke K, Steinhauser C. Distinct types of astroglial cells in the hippocampus differ in gap junction coupling. Glia. 2004;48:36–43. [PubMed: 15326613]
115.
Paukert M, Bergles DE. Synaptic communication between neurons and NG2 cells. Curr Opinion in Neurobiol. 2006;16:515–521. [PubMed: 16962768]
116.
Petroff OAC. Metabolic biopsy of the brain. In: Waxman SG, editor. Molecular neurology. New York: Elsevier; 2007. pp. 77–100.
117.
Malarkey EB, Parpura V. Mechanisms of glutamate release from astrocytes. Neurochem International. 2008;52:142–154. [PMC free article: PMC2267911] [PubMed: 17669556]
118.
Kimelberg HK, Mongin AA. Swelling-activated release of excitatory amino acids in the brain: relevance for pathophysiology. Contrib Neprhol. 1998;123:240–257. [PubMed: 9761972]
119.
Bordey A, Sontheimer H. Properties of human glial cells associated with epileptic tissue. Epilepsy Res. 1998;32:286–303. [PubMed: 9761328]
120.
Williamson A. Electrophysiology of epileptic human neocortical and hippocampal neurons maintained in vitro. Clin Neurosci. 1994;2:47–52.
121.
Özbas-Gerceker F, Redeker S, Boer K, Özgüc M, Saygi S, Dalkara T, Soylemezoglu F, Akalan N, Baayen JC, Gorter JA, Aronica E. Serial analysis of gene expression in the hippocampus of patients with mesial temporal lobe epilepsy. Neuroscience. 2006;138:457–474. [PubMed: 16413123]
122.
van Gassen KLI, de Wit M, Groot Koerkamp MJA, Rensen MGA, van Rijen PC, Holstege FCP, Lindhout D, de Graan PNE. Possible role of the innate immunity in temporal lobe epilepsy. Epilepsia. 2008;49:1055–1065. [PubMed: 18076643]
123.
Andjelkovic AV, Pachter JS. Characterization of binding sites for chemokines MCP-1 and MIP-1a on human brain microvessels. J Neurochem. 2000;75:1898–1906. [PubMed: 11032879]
124.
Dong Y, Benveniste EN. Immune function of astrocytes. Glia. 2001;36:180–190. [PubMed: 11596126]
125.
John GR, Lee SC, Song X, Rivieccio M, Brosnan CF. IL-1-Regulated responses in astrocytes: relevance to injury and recovery. Glia. 2005;49:161–176. [PubMed: 15472994]
126.
Ravizza T, Gagliardi B, Noé F, Boer K, Aronica E, Vezzani A. Innate and adaptive immunity during epileptogenesis and spontaneous seizures: Evidence from experimental models and human temporal lobe epilepsy. Neurobiol Dis. 2008;29:142–160. [PubMed: 17931873]
127.
Vezzani A, Moneta D, Richichi C, Aliprandi M, Burrows SJ, Ravizza T, Perego C, De Simoni MG. Functional role of inflammatory cytokines and antiinflammatory molecules in seizures and epileptogenesis. Epilepsia. 2002;43(suppl 5):30–35. [PubMed: 12121291]
128.
Amaral DG, Witter MP. Hippocampal formation. In: Paxinos G, editor. The rat nervous system. 2nd ed. San Diego: Academic Press; 1995. pp. 443–493.
129.
Fisher PD, Sperber EF, Moshe SL. Hippocampal sclerosis revisited. Brain Dev. 1998;20:563. [PubMed: 9865538]
130.
Mathern GW, Babb TL, Armstrong DL. Hippocampal sclerosis. In: Engel J, Pedley TA, editors. Epilepsy: A comprehensive Textbook. Philadelphia: Lippincott-Raven; 1997. pp. 133–155.
131.
Cohen I, Navarro V, Clemenceau S, Baulac M, Miles R. On the origin of interictal activity in human temporal lobe epilepsy in vitro. Science. 2002;298:1418–1421. [PubMed: 12434059]
132.
Hiuberfield G, Wittner L, Clemenceau S, Baulac M, Kaila K, Miles R, Rivera C. Perturbed chloride homeostasis and GABAergic signalling in human temporal lobe epilepsy. J Neurosci. 2007;27:9866–9873. [PMC free article: PMC6672644] [PubMed: 17855601]
133.
Munoz A, Mendez P, DeFelipe J, Alvarez-Leefmans FJ. Cation-chloride cotransporters and GABAergic innervation in the human epileptic hippocampus. Epilepsia. 2007;48:663–673. [PubMed: 17319917]
134.
Palma E, Amici M, Sobrero F, Spinelli G, Di Angelantonio S, Ragozzino D, Mascia A, Scoppetta C, Esposito V, Miledi R, Eusebi F. Anomolous levels of Cl- transporters in the hippocampal subiculum from temporal lobe epilepsy patients make GABA excitatory. PNAS. 2006;103:8465–8468. [PMC free article: PMC1482515] [PubMed: 16709666]
135.
Wozny C, Kivi A, Lehmann T-N, Dehnicke C, Heinemann U, Behr J. Comment on “On the origin of interictal activity in human temporal lobe epilepsy in vitro” Science. 2003;301:463c–464c. [PubMed: 12881553]
136.
Bernasconi N, Andermann F, Arnold DL, Bernasconi A. Entorhinal cortex MRI assesment in temporal, extratemporal and idiopathic generalized epilepsy. Epilepsia. 2003;44:1070–1074. [PubMed: 12887438]
137.
Bernasconi N, Bernasconi A, Carmanos Z, Antel SB, Andermann F, Arnold DL. Mesial temporal damage in temporal lobe epilepsy: A volumetric study of the hippocampus and parahippocampal region. Brain. 2003;126:462–469. [PubMed: 12538412]
138.
Du F, Whetsell WO Jr, Abou-Khalil B, Blumenkopf B, Lothman EW, Schwarcz R. Preferential neuronal loss in layer III of the entorhinal cortex in patients with temporal lobe epilepsy. Epilepsy Res. 1993;16:223–233. [PubMed: 8119273]
139.
Yilmazer-Hanke DM, Wolf HK, Schramm J, Elger CE, Wiestler OD, Blümcke I. Subregional pathology of the amygdala complex and entorhinal region in surgical specimens from patients with pharmacoresistant temporal lobe epilepsy. J Neuropath Exp Neurol. 2000;59:907–920. [PubMed: 11079781]
140.
Dawodu S, Thom M. Quantitative neuropathology of the entorhinal cortex region in patients with hippocampal sclerosis and temporal lobe epilepsy. Epilepsia. 2005;46:23–30. [PubMed: 15660765]
141.
Cavus I, Abi-Saab WM, Cassadey M, Jackob R, Sherwin RS, Krysal J, Spencer DD. Basal glutamate, gamma-aminobutyric acid, glucose, and lactate levels in the epileptogenic and non-epileptogenic brain site in neurosuergery patients. Epilepsia. 2002;43:247.
142.
Landmark CJ. Targets for antiepileptic drugs in the synapse. Med Sci Monit. 2007;13:RA1–7. [PubMed: 17179916]
143.
White HS, Smith MD, Wilcox KS. Mechanisms of action of antiepileptic drugs. Int Rev Neurobiol. 2007;81:85–110. [PubMed: 17433919]
144.
Brines ML, Sundaresan S, Spencer DD, de Lanerolle NC. Quantitative autoradiographic analysis of ionotropic glutamate receptor subtypes in human temporal lobe epilepsy: upregulation in reorganized epileptogenic hippocampus. Eur J Neurosci. 1997;9:2035–2044. [PubMed: 9421164]
Copyright © 2012, Michael A Rogawski, Antonio V Delgado-Escueta, Jeffrey L Noebels, Massimo Avoli and Richard W Olsen.

All Jasper's Basic Mechanisms of the Epilepsies content, except where otherwise noted, is licensed under a Creative Commons Attribution-NonCommercial-NoDerivs 3.0 Unported license, which permits copying, distribution and transmission of the work, provided the original work is properly cited, not used for commercial purposes, nor is altered or transformed.

Bookshelf ID: NBK98141PMID: 22787669

Views

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...