Specific imbalance of excitatory/inhibitory signaling establishes ...

4 downloads 0 Views 965KB Size Report
Apr 13, 2016 - experimental conditions (i.e., 4-aminopyridine application), the initiation of LVF or HYP onset. 52 seizures depends on the preponderant ...
Articles in PresS. J Neurophysiol (April 13, 2016). doi:10.1152/jn.01128.2015 Journal of Neurophysiology JN-01128-2015 (Review)

1 2

SPECIFIC IMBALANCE OF EXCITATORY/INHIBITORY SIGNALING

3

ESTABLISHES SEIZURE ONSET PATTERN IN TEMPORAL LOBE

4

EPILEPSY

5 6

Massimo Avoli1,2,*, Marco de Curtis3, Vadym Gnatkovsky3, Jean Gotman1, Rüdiger Köhling4,

7

Maxime Lévesque1, Frédéric Manseau5, Zahra Shiri1, Sylvain Williams5

8 9

1

Montreal Neurological Institute and Departments of Neurology & Neurosurgery and of

10

Physiology, McGill University, Montréal, H3A 2B4 Québec, Canada; 2Facoltà di Medicina e

11

Odontoiatria, Sapienza Università di Roma, 00185 Roma, Italy; 3Epilepsy Unit, Fondazione

12

Istituto Neurologico Carlo Besta, 20133 Milano, Italy; 4Oscar-Langendorff-Institute of

13

Physiology, Rostock University Medical Center, 18057 Rostock, Germany; 5Douglas Mental

14

Health University Institute, McGill University, Montréal, H4H 1R3 Québec, Canada.

15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

Running title: Mechanisms of seizure onset Number of characters in title and running head: 140 Number of words in abstract: 217 Number of words in body of manuscript: 3 677 Number of figures: 5 * Corresponding Author:

Massimo Avoli, M.D., Ph.D. Montreal Neurological Institute 3801 University Street, Montréal, QC Canada, H3A 2B4 Tel Canada: +1 514 998 6790 Tel Europe: +39 333 483 1060 Fax: +1 514 398 8106 E-mail: [email protected]

CONFLICTS OF INTEREST None of the authors have any conflict of interest to disclose.

37

1

Copyright © 2016 by the American Physiological Society.

Journal of Neurophysiology JN-01128-2015 (Review)

38

Abstract

39

Low-voltage fast (LVF) and hypersynchronous (HYP) patterns are the seizure onset patterns

40

most frequently observed in intracranial EEG recordings from mesial temporal lobe epilepsy

41

(MTLE) patients. Both patterns also occur in models of MTLE in vivo and in vitro, and these

42

studies have highlighted the predominant involvement of distinct neuronal network/neu-

43

rotransmitter receptor signaling in each of them. First, LVF onset seizures in epileptic rodents

44

can originate from several limbic structures, frequently spread, and are associated with high

45

frequency oscillations (HFOs) in the ripple band (80-200 Hz), while HYP onset seizures initiate

46

in the hippocampus, and tend to remain focal with predominant fast ripples (250-500 Hz).

47

Second, in vitro intracellular recordings from principal cells in limbic areas indicate that

48

pharmacologically induced, seizure-like discharges with LVF onset are initiated by a

49

synchronous inhibitory event or by a hyperpolarizing IPSPs barrage; in contrast, HYP onset is

50

associated with a progressive impairement of inhibition and concomitant unrestrained enhance-

51

ment of excitation. Finally, in vitro optogenetic experiments show that under comparable

52

experimental conditions (i.e., 4-aminopyridine application), the initiation of LVF or HYP onset

53

seizures depends on the preponderant involvement of interneuronal or principal cell networks,

54

respectively. Overall, these data may provide insight to delineate better therapeutic targets in

55

the treatment of patients presenting with MTLE and, perhaps, with other epileptic disorders as

56

well.

57 58 59 60 61

2

Journal of Neurophysiology JN-01128-2015 (Review)

62

New and noteworthy

63

We review here the mechanisms leading to two seizure onset EEG patterns that occur in mesial

64

temporal lobe epilepsy. Based on data obtained in epileptic patients and in animal models, we

65

propose that the initiation of low-voltage fast onset and hypersynchronous onset seizures

66

depend on the involvement of GABAergic interneuron and of principal (glutamatergic) networks,

67

respectively, which in both cases rest on functional GABAA receptor signaling.

68 69

3

Journal of Neurophysiology JN-01128-2015 (Review)

70

Background

71

The EEG activity recorded during focal seizures in patients with mesial temporal lobe epilepsy

72

(MTLE) is characterized by different levels of detail, depending on the recording approach used.

73

Standard scalp electrode recordings may reveal, at the onset of a seizure, flattening of the EEG

74

signals localized in the temporal lobe that is at times associated with the appearance of low

75

amplitude, fast rhythms (Gloor, 1975; Fisher et al., 1992; see also de Curtis and Gnatkovsky,

76

2009). However, this straightforward diagnostic tool is not ideal for identifying the exact seizure

77

initiation area that resides in mesial cortical structures such as the hippocampus or the

78

amygdala; in addition, many seizures in MTLE do not show this pattern: some show other

79

patterns, and many only show artefacts at onset. In contrast, intracranial EEG recordings

80

obtained from deep limbic regions - which are the gold standard in presurgical evaluation of

81

pharmacoresistant epileptic patients who are candidates for surgery - can lead to a more

82

precise identification of the brain structures involved in seizure initiation while revealing in detail

83

the features of discharge onset (Lieb et al., 1976; Pacia and Ebersole, 1997).

84

Seizures in MTLE patients studied with intracranial electrodes have variable EEG

85

features, but evidence published during the last two decades indicate that two specific seizure

86

onset patterns may predominate in temporal lobe (limbic) structures (Spencer and Pappas,

87

1992; Spanedda et al., 1997; Wendling et al., 2005; Ogren et al., 2009; Perucca et al., 2014;

88

Singh et al., 2015). The most frequent onset pattern is characterized by “low-voltage fast” (LVF)

89

activity in the gamma range, at times initiated by a single (also termed “sentinel’) spike (Fig. 1A,

90

asterisk in the left panel); the second onset pattern- referred to as “hypersynchronous” (HYP) -

91

is associated with an initial series of large amplitude spikes that occur at a frequeny of approx. 1

92

Hz (Fig. 1A, asterisks in the right panel) and can also reappear during the initial part of the

93

seizure (Fig. 1A, arrow heads in the right panel). As illustrated in Fig. 1A, LVF as well as HYP

4

Journal of Neurophysiology JN-01128-2015 (Review)

94

onset seizures evolve into a phase of sustained oscillatory activity (also defined as “tonic”) and

95

then into bursting, rhythmic (“clonic”) discharges; in both cases, termination of seizure discharge

96

is followed by a period of spontaneous electrical activity suppression that is termed “post-ictal

97

depression”.

98

The clinical evidence for two distinct seizure onset patterns has been confirmed in

99

animal models of MTLE and of epileptiform synchronization in both in vivo (Fig. 1B) (Bragin et

100

al., 2005; Lévesque et al., 2012; Grasse et al., 2013; Salami et al., 2015; Toyoda et al., 2015)

101

and in vitro preparations (Fig. 1C) (Avoli et al., 1996; Lopantsev and avoli, 1998a and b;

102

Derchansky et al., 2008; Zhang et al., 2012; Boido et al., 2014; Köhling et al., 2016; see for

103

review Avoli and de Curtis, 2011). Moreover, these fundamental studies have provided evidence

104

suggesting that specific neuronal networks may contribute to LVF and HYP seizure onset

105

patterns. In this review, we will briefly analyze evidence obtained in patients suffering from

106

MTLE; these clinical studies suggest that LVF and HYP seizure onset patterns may reflect

107

different epileptic conditions along with different degrees of brain damage, at least in MTLE.

108

Then, we will consider experimental data obtained in vivo from the pilocarpine model of MTLE

109

as well as in normal, control rodents following acute, diverse convulsive pharmacological

110

manipulations. Next, we will summarize in vitro experiments indicating the predominant

111

involvement of GABAA receptor signaling in initiating LVF seizure-like discharges and the

112

dynamic changes in inhibition that accompany the onset of HYP seizure-like activity. Finally, we

113

will analyze recent optogenetic in vitro findings; these results strongly suggest that the initiation

114

of LVF and HYP onset seizures depends on the preponderant involvement of (GABAergic)

115

interneuron and of principal (glutamatergic) cell networks, respectively, which in both cases rest

116

on an operational GABAA receptor signaling.

117

5

Journal of Neurophysiology JN-01128-2015 (Review)

118

Seizure onset patterns in the clinical context

119

Long-term, intracranial depth recordings obtained during pre-surgical stereo-EEG monitoring

120

have revealed that LVF seizure onset represents the most common pattern of ictal discharge

121

initiation. Indeed, it occurs across several neocortical focal epilepsies and is not restricted to

122

MTLE (Gotman et al., 1995; Gnatkovsky et al., 2011; Perucca et al., 2014; Wu et al., 2014;

123

Singh et al., 2015). In addition, MTLE patients presenting with LVF onset seizures show diffuse

124

neuronal loss that can comprise the CA2/CA4 regions (Velasco et al., 2000; Ogren et al., 2009).

125

This pattern is also more frequent in patients showing amygdala atrophy in addition to

126

hippocampal atrophy as well as in patients with normal mesial temporal structures (Spanedda et

127

al., 1997).

128

In contrast, the HYP onset pattern is most often, if not solely, observed in MTLE with

129

hippocampal sclerosis and it has never been reported in neocortical focal epilepsies (Spencer et

130

al., 1992; Velasco et al., 2000; Ogren et al., 2009; Perucca et al., 2014). Magnetic resonance

131

imaging studies and histopathological evaluation of post-surgical specimens have also

132

demonstrated that MTLE patients presenting with HYP onset seizures present with cell loss in

133

all hippocampal areas with the exception of the CA2 subfield and the presubiculum (Spencer et

134

al., 1992; Velasco et al., 2000). In addition, Ogren et al. (2009) have proposed that HYP onset

135

seizures occur only in MTLE patients presenting with pronounced but restricted hippocampal

136

atrophy. Overall, evidence obtained in epileptic patients indicate that these two seizure onset

137

patterns often reflect different histopathological conditions, which in turn suggests the

138

involvement of different types of neuronal networks and presumably a specific contribution of

139

ligand-gated mechanisms.

140

Figure 1 approx. here

141

6

Journal of Neurophysiology JN-01128-2015 (Review)

142

Seizure onset patterns in in vivo animal models of epilepsy and of epileptiform

143 144 145

synchronization Bragin et al. (2005) were first in reporting that in the “local” kainic acid model of MTLE, rats

146

generate spontaneous seizures characterized by both LVF and HYP onset patterns; in these

147

experiments, EEG recordings were obtained with depth electrodes positioned in several limbic

148

structures. These researchers also found that HYP onset seizures initiate most often from the

149

hippocampus ipsilateral to the original kainic acid injection and that they are rarely accompanied

150

by behavioral symptoms; therefore, these data suggest that these seizures remained focal. On

151

the contrary, LVF onset seizures occurring in these experiments originated from the

152

hippocampus and from the entorhinal cortex, and they frequently propagated to other limbic and

153

extralimbic areas (Bragin et al., 2005).

154

Similar results have been later obtained in the pilocarpine model of MTLE by Lévesque

155

et al. (2012) who found that the majority of HYP onset seizures originated from the hippocampal

156

CA3 region, while LVF onset seizures initiated in this hippocampal subfield as well as in the

157

entorhinal cortex or even outside the hippocampal formation (Fig. 1B). An additional crucial

158

difference between LVF and HYP seizures in the pilocarpine model rests on their association

159

with specific types of pathological high frequency oscillations (HFOs) at 80-500 Hz, which have

160

been arbitrarily categorized into ripples (80-200 Hz) and fast ripples (250-500 Hz) (Fig. 2A)

161

according to the original proposal made by Bragin et al. (1999). HFOs - which can only be

162

extracted by amplifying the appropriately filtered EEG signal - have been recorded from patients

163

presenting with MTLE and in animal models mimicking this neurological condition, and they

164

have been proposed to represent better biomarkers than interictal spikes for identifying seizure

165

onset zones in focal epileptic disorders (Jacobs et al., 2008, 2010; see also Jacobs et al. 2012;

166

Staba, 2012).

7

Journal of Neurophysiology JN-01128-2015 (Review)

167

Excitatory and inhibitory synaptic interactions along with intrinsic membrane

168

oscillations, gap junctions and ephaptic coupling have been considered to underlie physiological

169

(Buzsáki and Chrobak, 1995; Ylinen et al., 1995; Buzsáki, 2015) as well as pathological

170

(Jefferys et al., 2012a, 2012b) HFOs, and the contribution of these mechanisms to the

171

generation of HFOs is under active examination. In addition, it has been reported that HFOs

172

with similar frequency ranges can differ considerably in their physiological mechanisms (Jefferys

173

et al., 2012a, 2012b). However, a convenient – though reductionist and, perhaps, simplistic -

174

view is that pathologic HFOs in the ripple band mainly represent population IPSPs generated by

175

principal neurons that are entrained by synchronously active interneuron networks; in contrast,

176

HFOs in the fast ripple band could be supported by the synchronous firing of abnormally active

177

principal cells (Foffani et al., 2007), and could be independent of inhibitory neurotransmission

178

(cf., Jefferys et al., 2012a). However, data obtained in some studies (e.g., D’Antuono et al.,

179

2005; Bragin et al., 2011) have suggested that pathological ripples and fast ripples are both

180

produced by pyramidal cell action potentials.

181

As shown in Fig. 2B and C, HFO occurrence during pre-ictal and ictal periods in

182

pilocarpine-treated epileptic rats markedly differ between LVF and HYP onset seizures. During

183

both pre-ictal and ictal period, LVF seizures were associated with a preponderant increase in

184

ripple occurrence, while HYP seizures were mostly characterized by increased occurrence of

185

fast ripples (Lévesque et al., 2012). Similar results had also been previously reported in kainic

186

acid- treated epileptic rats (Bragin et al., 2005). Therefore, according to what was summarized

187

in the previous paragraph, evidence obtained to date suggest that distinct transmitter signaling

188

(and underlying neuronal network activity) predominate during LVF and HYP seizures recorded

189

from epileptic animals; specifically, LVF onset should be mirrored by increased interneuron

8

Journal of Neurophysiology JN-01128-2015 (Review)

190

(GABAergic cell) activity while excitatory (glutamatergic) cells should be the main actors in the

191

initiation of HYP onset seizures (Lévesque et al., 2012).

192

This hypothesis is supported by a recent study in which multiunit activity recordings

193

were obtained during spontaneous seizures in awake epileptic animals treated with kainic acid

194

(Grasse et al., 2013). These investigators discovered that LVF seizures are associated with

195

increased interneuron activity at onset followed by intense firing generated by principal cells. No

196

in vivo study has yet investigated the activity of interneurons and principal cells during HYP

197

onset seizures in animal models of MTLE, although this aspect has been addressed in the feline

198

neocortex in an in vivo acute preparation (Timofeev et al., 2002; Grenier et al., 2003).

199

Figure 2 approx. here

200

The hypothesis that the inititation of LVF onset seizures mainly rests on GABAergic

201

function has been further tested in vivo by Salami et al. (2015); they induced seizures in normal,

202

control rats - which were chronically implanted with depth electrodes - through the systemic

203

injection of the K+ channel blocker 4-aminopyridine, which is known to enhance both

204

glutamatergic and GABAergic transmission (Buckle and Haas, 1982; Rutecki et al., 1987;

205

Perreault and Avoli, 1991), or the GABAA receptor antagonist picrotoxin (De Groat et al., 1972).

206

These experiments have shown that 4-aminopyridine causes seizures with LVF onset in the

207

majority of cases, while seizures induced by picrotoxin are consistently characterized by HYP

208

onset. In addition, HFO analysis revealed that 4-aminopyridine-induced LVF seizures are

209

associated with higher ripple rates compared to fast ripples, whereas picrotoxin-induced

210

seizures contained significantly higher rates of fast ripples compared to ripples (Salami et al.,

211

2015).

212 213

Seizure onset patterns in in vitro models of epileptiform synchronization

9

Journal of Neurophysiology JN-01128-2015 (Review)

214

Electrograhic activity closely resembling the ictal discharges recorded in epileptic patients and in

215

in vivo animal models can also be reproduced in several in vitro preparations such as the brain

216

slice (Fig. 1C), the isolated hippocampus, and the guinea pig whole brain (see for review: Avoli

217

and Jefferys, 2016; de Curtis et al., 2015; de Curtis and Avoli, 2016). In these studies, both LVF

218

and HYP onset seizure-like discharges can be recorded from several limbic and extralimbic

219

areas, and to review in detail these data is beyond the goal of this paper. However, a few

220

remarks should be made. First, long-lasting periods of epileptiform synchronization, resembling

221

ictal discharges, are rarely observed during pharmacological manipulations that fully antagonize

222

GABAA receptor signaling, at least when employing adult brain tissue (see for review: Avoli and

223

de Curtis, 2011; Avoli and Jefferys, 2016). Second, HYP onset ictal discharges are commonly

224

seen when cortical tissue is perfused with medium containing low Mg2+ (Derchansky et al.,

225

2008; Zhang et al., 2012) but are less frequently recorded during bath application of 4-

226

aminopyridine (Lopantsev and Avoli, 1998b; Avoli et al., 2013) with the notable exception of

227

experiments carried out in the perirhinal cortex (Biagini et al., 2013; Köhling et al., 2016). In fact,

228

as reviewed in detail by Avoli and de Curtis (2011), this K+ channel blocker readily and most

229

often induces ictal discharges that closely resemble an LVF seizure onset in several limbic and

230

extralimbic structures (Figs. 1C and 3A & B).

231

LVF seizure-like discharges recorded in vitro during application of 4-aminopyridine

232

have been extensively studied in our laboratories, employing both rodent brain slices and the

233

guinea pig whole brain preparation. When intracellularly recorded from principal (glutamatergic)

234

cells of the entorhinal cortex in a brain slice preparation, the LVF onset coincides with a

235

depolarization that is associated with few if any action potentials and becomes hyperpolarizing

236

when the neuron membrane potential is brought to values less negative than -60 mV with

237

steady injection of depolarizing current (Lopantsev and Avoli, 1998a) (Fig. 3A). Therefore, the

10

Journal of Neurophysiology JN-01128-2015 (Review)

238

LVF onset of 4-aminopyridine-induced ictal discharges appears to be associated with a robust

239

synchronous inhibitory event; this view is supported by studies in which powerful interneuron

240

discharges could be recorded at the onset of these ictal discharges (Ziburkus et al., 2006; de

241

Curtis et al., 2015; Uva et al., 2015; Lévesque et al., 2016) as well as by the ability of GABAA

242

receptor antagonists to abolish this event along with ictal discharge occurrence (Avoli et al.,

243

1996; Lopantsev and Avoli, 1998a). It should be noticed that short-lasting depolarizations with

244

similar pharmacological and electrophysiological (e.g., few action potentials and reversal

245

potentials) characteristics are associated with the interictal-like discharges (Fig. 3A, arrow) that

246

occur in vitro in the hippocampus and entorhinal cortex during 4-aminopyridine treatement (Avoli

247

et al., 1996; Lopantsev and Avoli, 1998; Uva et al., 2009).

248

It has also been demonstrated that the negative shift of the local field potential

249

observed at the onset of an LVF seizure correlates with elevations in [K+]o (Fig. 3B) that rest on

250

excessive GABAergic signaling (Avoli et al., 1996; reviewed by Avoli and de Curtis, 2011; Avoli

251

et al., 2013; de Curtis and Avoli, 2016), leading to intracellular Cl- accumulation and subsequent

252

activity of the KCC2 co-transporter that extrudes both Cl- and K+ from the intraneuronal

253

compartment (Viitanen et al., 2010). Such increases in [K+]o should depolarize neighbouring

254

neurons thus causing hyperexcitability as suggested by the occurrence of small amplitude,

255

presumably ectopic spikes (Lopantsev and Avoli, 1998a; Trombin et al., 2011; Kaila et al., 2014;

256

but also see Avoli et al., 1998). In addition, an elevation in [K+]o supports the emergence of

257

neuronal network resonance, thus generating oscillatory patterns in the beta-gamma range

258

(Bartos et al., 2007) (see also next paragraph), and should cause a positive shift of the

259

membrane reversal of the GABAA receptor mediated inhibitory postsynaptic potential therefore

260

weakening inhibition (Jensen et al., 1993). The notion that elevations in [K+]o increase neuronal

261

excitability and cause seizure activity has been firmly established over the last few decades in

11

Journal of Neurophysiology JN-01128-2015 (Review)

262

both in vivo (Zuckermann and Glaser, 1968) and in vitro (Traynelis and Dingledine, 1988)

263

preparations. Moreover, the role played by KCC2 activity in LVF onset seizure initiation and

264

maintainance is supported by recent evidence showing that ictal discharges induced by 4-

265

aminopyridine are abolished or facilitated by inhibiting or enhancing the activity of this co-

266

transporter, respectively (Hamidi and Avoli, 2015).

267

Figure 3 approx. here

268

As illustrated in Fig. 3C, LVF ictal discharge in the entorhinal cortex of the isolated

269

guinea pig brain can also be induced by a short lasting arterial perfusion of bicuculline, a

270

pharmacological procedure that reduces inhibition efficacy only by approx. 30%, as also

271

suggested by the ability of piriform or entorhinal cortical cells to generate a robust inhibitory

272

post-synaptic potential following electrical activation of the lateral olfactory tract under this

273

experimental condition (Fig. 3D). In this specific in vitro model of ictogenesis, LVF seizures are

274

initiated by fast field oscillations in the beta-gamma range that are mirrored by intracellular

275

hyperpolarizing potentials becoming of smaller amplitude as seizure progresses (Gnatkovsky et

276

al., 2008) (Fig.3C). In addition, similar to what was observed in the experiments performed with

277

4-aminopyridine, these LVF onset events corresponded to elevation in [K+]o along with the

278

occurrence of ectopic action potentials (Gnatkovsky et al., 2008; Trombin et al., 2011).

279

As already mentioned, HYP onset ictal discharges have been analyzed in cortical

280

tissue bathed with medium containing low Mg2+ (Derchansky et al., 2008; Zhang et al., 2012),

281

an experimental manipulation that is known to weaken GABAA receptor signaling over time

282

(Whittington et al., 1995). Interestingly, maintained GABAA receptor-mediated activity was

283

initially reported by Derchansky et al. (2008) during the transition to HYP seizure activity in the

284

isolated immature mouse hippocampus; however, further experiments carried out in the same

285

laboratory have shown that HYP seizure onset is characterized by “exhaustion of presynaptic

12

Journal of Neurophysiology JN-01128-2015 (Review)

286

release of GABA, and unopposed increase in glutamatergic excitation” (Zhang et al., 2012).

287

More recently, we have reported that ictal discharges with HYP onset features can also occur in

288

the perirhinal cortex in brain slices superfused with 4-aminopyridine medium (Fig. 3E) (Biagini et

289

al., 2013; Köhling et al., 2015).

290

By employing intracellular recordings from principal cells of the perirhinal cortex we

291

have found that the recurrent field spikes typical of a HYP onset are characterized by

292

intracellular depolarizations associated with action potential bursting as well as that the post-

293

burst hyperpolarizations (presumably caused by activation of post-synaptic GABAA receptors)

294

gradually decrease in amplitude (Fig. 3E, arrowheads), a phenomenon that is characterized by

295

a gradual positive shift in their reversal potential (Köhling et al., 2015). In addition, these

296

changes were accompanied by a progressive increase in the associated transient elevations in

297

[K+]o (Fig. 3F, arrowheads). While these data are in line with the conclusions proposed by Zhang

298

et al. (2012), who identified a progressive impairment of inhibition at HYP onset with

299

concomitant unrestrained enhancement of excitation, they also reveal an additional mechanism

300

of inhibition weakening that may rest on progressively larger accumulations in [K+]o due to the

301

postsynaptic activation of GABAA receptors.

302 303

Optogenetic approaches reveal the involvement of specific neuronal networks

304

in different seizure onset patterns

305

The pivotal role played by interneurons in the initiation of ictal discharges characterized by an

306

LVF onset has recently been confirmed with optogenetic techniques in the entorhinal cortex

307

during 4-aminopyridine treatment (Shiri et al., 2015a; Yekhlef et al., 2015). It was shown in

308

these studies that optogenetic stimulation of parvalbumin- or somatostatin-positive interneurons

13

Journal of Neurophysiology JN-01128-2015 (Review)

309

can initiate ictal LVF onset events similar to those occurring spontaneously (Fig 4A). In addition,

310

as illustrated in the expanded traces of Fig. 4Aa (arrows), the onset of an ictal discharge

311

induced by optogenetic stimulation of parvalbumin-positive cells presented with the typical

312

pattern consisting of one or two interictal-like transients leading to fast, beta-gamma activity

313

marking the beginning of the tonic phase. Interestingly, Shiri et al. (2015a) found that during

314

both spontaneous and stimulated LVF discharges, ripples rates predominated at ictal onset (Fig.

315

4Ab).

316

Figure 4 approx. here

317

However, as shown in Fig. 4B, LVF ictal events that occurred spontaneously during 4-

318

aminopyridine application virtually switched to HYP onset events when the calcium/calmodulin-

319

dependent protein kinase II-positive principal cells were optogenetically stimulated in the

320

entorhinal cortex (Shiri et al., 2016). Specifically, the onset of ictal discharges triggered by

321

principal cell activation was characterized by repeated field spikes, thus closely resembling a

322

HYP onset pattern. In addition, these optogenetically induced HYP onset ictal events were

323

found to be associated with higher fast ripple rates at onset (Fig. 4Bb). Therefore, these

324

optogenetic experiments demonstrate that under identical conditions (i.e., during application of

325

4-aminopyridine), activation of each specific cell population can generate ictal discharges with a

326

different onset pattern: LVF onset events, similar to those occurring spontaneoulsly, depended

327

on the optogenetic activation of interneuronal networks while HYP onset discharges rest on the

328

optogenetic activation of glutamatergic principal cells.

329 330

Concluding remarks

331

The studies reviewed here underscore the concept that the “excessive” activity of interneurons

332

(and the resulting activation of postsynaptic GABAA receptors) can be sufficient to disrupt the

14

Journal of Neurophysiology JN-01128-2015 (Review)

333

excitation/inhibition balance within forebrain neuronal networks thus leading to ictal activity

334

(Avoli and de Curtis 2011; de Curtis and Avoli, 2016). This conclusion is in line with in vivo data

335

obtained from epileptic patients (Truccolo et al., 2011; Schevon et al., 2012) and animal models

336

(Grasse et al., 2013; Toyoda et al., 2015), in which seizure onset was shown to be associated

337

with sustained firing of interneurons or with depressed or consistent firing activity of principal

338

cells.

339

We also propose that different types of neuronal networks (i.e., interneurons and

340

principal, glutamatergic cells) and thus different neurotransmitter receptor signaling,

341

predominate in MTLE at the onset of LVF and HYP ictal discharges. As summarized in the

342

diagrams shown in Fig. 5, two different sets of cellular events/interactions are indeed likely to be

343

at work during these two focal seizure onset patterns. Moreover, in light of the evidence

344

obtained in human epileptic patients (Spencer et al., 1992; Velasco et al., 2000; Ogren et al.

345

2009) as well as of the structure-dependent (Köhling et al. 2015) and optogenetic (Shiri et al,

346

2016) findings identified in vitro, it is reasonable to hypothesize that the preponderant

347

involvement of interneuronal or principal cell networks in LVF and HYP onset seizures,

348

respectively, may reflect some distinctive features in network connectivity and/or in brain state

349

excitability.

350

The evidence reviewed here may provide insight to delineate better therapeutic targets

351

in the treatment of patients presenting with MTLE, and suggest that (theoretically) seizure onset

352

patterns and associated HFOs should be taken into consideration in order to implement optimal

353

pharmacological therapies. However, this last aspect may suffer a practical drawback since both

354

types of seizure onset can also coexist in the same experimental condition both in vivo

355

(Lévesque et al., 2012) and in vitro (Köhling et al., 2016) as well as in patients with MTLE (B.

356

Frauscher, F. Dubeau and J. Gotman, personal communication). Finally, it should be

15

Journal of Neurophysiology JN-01128-2015 (Review)

357

emphasized that the findings summarized in this review focus on determinants of seizure onset

358

types (i.e., on the mechanisms underlying ictogenesis), not on epileptogenesis. Therefore, the

359

impact of different seizure onset patterns on the development of MTLE remains to be defined.

360

16

Journal of Neurophysiology JN-01128-2015 (Review)

361

Acknowledgements

362

The original work reviewed here was supported by the CIHR (Grants 8109 and 74609 to MA,

363

143208 to JG, and 119340 to SW), CURE (MA), the Fondazione Banca del Monte di Lombardia

364

(2014-15 to MdC), the Italian Health Ministry (Ricerca Corrente 2014; grants RF-2010-2304417

365

and RF-2007-GR141 to MdC), the Savoy Foundation (MA), and the Telethon Foundation (grant

366

GGP12265 to MdC). We thank Drs. M. Barbarosie, R. Benini, G. Biagini, M. D’Antuono, P. de

367

Guzman, S. Hamidi, V. Lopantsev, J. Louvel, R. Pumain, P. Salami and L. Uva as well as Ms. I.

368

Kurcewicz for contributing to some of the original experiments that were reported in this review.

369

17

Journal of Neurophysiology JN-01128-2015 (Review)

370

References

371

Avoli M, Barbarosie M, Lücke A, Nagao T, Lopantsev V, Köhling R. Synchronous GABA-

372

mediated potentials and epileptiform discharges in the rat limbic system in vitro. J

373

Neurosci 16: 3912–3924, 1996.

374 375 376 377 378 379 380 381 382

Avoli M, de Curtis M. GABAergic synchronization in the limbic system and its role in the generation of epileptiform activity. Prog Neurobiol 95: 104–132, 2011. Avoli M, de Curtis M, Köhling R. Does interictal synchronization influence ictogenesis? Neuropharmacology 69: 37-44, 2013 Avoli M, Jefferys JGR. Models of drug-induced epileptiform synchronization in vitro. J. Neurosci. Methods 260: 26-32, 2016. Avoli M, Methot M, Kawasaki H. GABA-dependent generation of ectopic action potentials in the rat hippocampus. Eur J Neurosci 10: 2714–2722, 1998. Avoli M, Panuccio G, Herrington R, D’Antuono M, de Guzman P, Lévesque M. Two

383

different interictal spike patterns anticipate ictal activity in vitro. Neurobiol Dis 52: 168–176,

384

2013.

385

Barbarosie M, Louvel J, Kurcewicz I, Avoli M. CA3-released entorhinal seizures disclose

386

dentate gyrus epileptogenicity and unmask a temporoammonic pathway. J Neurophysiol

387

83: 1115-1124, 2000.

388 389 390 391

Bartos M, Vida I, Jonas P. Synaptic mechanisms of synchronized gamma oscillations in inhibitory interneuron networks. Nat Rev Neurosci 8: 45–56, 2007. Biagini G, D’Antuono M, Benini R, de Guzman P, Longo D, Avoli M. Perirhinal cortex and temporal lobe epilepsy. Front Cell Neurosci 7: 130, 2013.

18

Journal of Neurophysiology JN-01128-2015 (Review)

392

Boido D, Jesuthasan N, de Curtis M, Uva L. Network dynamics during the progression of

393

seizure-like events in the hippocampal-parahippocampal regions. Cereb Cortex N Y N

394

1991 24: 163–173, 2014.

395

Bragin A, Azizyan A, Almajano J, Wilson CL, Engel J Jr. Analysis of chronic seizure onsets

396

after intrahippocampal kainic acid injection in freely moving rats. Epilepsia 46: 1592–1598,

397

2005.

398

Bragin A, Benassi SK, Kheiri F, Engel J, Jr. Further evidence that pathologic high-frequency

399

oscillations are bursts of population spikes derived from recordings of identified cells in

400

dentate gyrus. Epilepsia: 52: 45-52, 2011.

401

Bragin A, Engel JJ, Wilson CL, Fried I, Mathern GW. Hippocampal and entorhinal cortex

402

high-frequency oscillations (100–500 Hz) in human epileptic brain and in kainic acid-

403

treated rats with chronic seizures. Epilepsia 40: 127-137, 1999.

404 405 406 407 408 409 410

Buckle PJ, Haas HL. Enhancement of synaptic transmission by 4-aminopyridine in hippocampal slices of the rat. J Physiol 326: 109-122, 1982. Buzsáki G. Hippocampal sharp wave-ripple: A cognitive biomarker for episodic memory and planning. Hippocampus 25:1073-1188, 2015. Buzsáki G, Chrobak JJ. Temporal structure in spatially organized neuronal ensembles: a role for interneuronal networks. Curr Opin Neurobiol 5: 504–510, 1995. D’Antuono M, de Guzman P, Kano T, Avoli M. Ripple activity in the dentate gyrus of

411

dishinibited hippocampus-entorhinal cortex slices. J Neurosci Res 80: 92-103, 2005.

412

de Curtis M, Avoli M. GABAergic networks jump-start focal seizures. Epilepsia 2016, in press.

413

de Curtis M, Gnatkovsky V. Reevaluating the mechanisms of focal ictogenesis: The role of

414

low-voltage fast activity. Epilepsia 50: 2514–2525, 2009.

19

Journal of Neurophysiology JN-01128-2015 (Review)

415 416 417

de Curtis M, Librizzi L, Uva L. The in vitro isolated whole guinea pig brain as a model to study epileptiform activity patterns. J. Neurosci. Methods 260: 83-90, 2016. De Groat WC, Lalley PM, Saum WR. Depolarization of dorsal root ganglia in the cat by GABA

418

and related amino acids: antagonism by picrotoxin and bicuculline. Brain Res 44: 273-277,

419

1972

420

Derchansky M, Jahromi SS, Mamani M, Shin DS, Sik A, Carlen PL. Transition to seizures in

421

the isolated immature mouse hippocampus: a switch from dominant phasic inhibition to

422

dominant phasic excitation. J Physiol 586: 477–494, 2008.

423

Fisher RS, Webber WR, Lesser RP, Arroyo S, Uematsu S. High-frequency EEG activity at

424

the start of seizures. J Clin Neurophysiol Off Publ Am Electroencephalogr Soc 9: 441–448,

425

1992.

426

Foffani G, Uzcategui YG, Gal B, Menendez de la Prida L. Reduced spike timing reliability

427

correlates with the emergence of fast ripples in the rat epileptic hippocampus. Neuron 55,

428

930–941, 2007.

429 430 431

Gloor P. Contributions of electroencephalography and electrocorticography to the neurosurgical treatment of the epilepsies. Neurosurg. Manag. Epilepsies. 1975. Gnatkovsky V, Francione S, Cardinale F, Mai R, Tassi L, Lo Russo G, de Curtis M.

432

Identification of reproducible ictal patterns based on quantified frequency analysis of

433

intracranial EEG signals. Epilepsia 52: 477–488, 2011.

434

Gnatkovsky V, Librizzi L, Trombin F, de Curtis M. Fast activity at seizure onset is mediated

435

by inhibitory circuits in the entorhinal cortex in vitro. Ann Neurol 64: 674–686, 2008.

436

Gotman J, Levtova V, Olivier A. Frequency of the electroencephalographic discharge in

437

seizures of focal and widespread onset in intracerebral recordings. Epilepsia 36: 697–703,

438

1995.

20

Journal of Neurophysiology JN-01128-2015 (Review)

439 440 441 442 443 444 445

Grasse DW, Karunakaran S, Moxon KA. Neuronal synchrony and the transition to spontaneous seizures. Exp Neurol 248: 72–84, 2013. Grenier F, Timofeev I, Steriade M. Neocortical very fast oscillations (ripples, 80-200 Hz) during seizures: intracellular correlates. J Neurophysiol 89: 841-852, 2003. Hamidi S, Avoli M. KCC2 function modulates in vitro ictogenesis. Neurobiol Dis 79: 51–58, 2015. Jacobs J, LeVan P, Chander R, Hall J, Dubeau F, Gotman J. Interictal high-frequency

446

oscillations (80-500 Hz) are an indicator of seizure onset areas independent of spikes in

447

the human epileptic brain. Epilepsia 49: 1893–1907, 2008.

448

Jacobs J, Staba R, Asano E, Otsubo H, Wu JY, Zijlmans M, Mohamed I, Kahane P,

449

Dubeau F, Navarro V, Gotman J. High-frequency oscillations (HFOs) in clinical epilepsy.

450

Prog Neurobiol 98: 302–315, 2012.

451

Jacobs J, Zijlmans M, Zelmann R, Chatillon C-E, Hall J, Olivier A, Dubeau F, Gotman J.

452

High-frequency electroencephalographic oscillations correlate with outcome of epilepsy

453

surgery. Ann Neurol 67: 209–220, 2010.

454

Jefferys JGR, Jiruska P, de Curtis M, Avoli M. Limbic Network Synchronization and Temporal

455

Lobe Epilepsy [Online]. In: Jasper’s Basic Mechanisms of the Epilepsies, edited by

456

Noebels JL, Avoli M, Rogawski MA, Olsen RW, Delgado-Escueta AV. National Center for

457

Biotechnology Information (US) http://www.ncbi.nlm.nih.gov/books/NBK98158/.

458

Jefferys JGR, Menendez de la Prida L, Wendling F, Bragin A, Avoli M, Timofeev I, Lopes

459

da Silva FH. Mechanisms of physiological and epileptic HFO generation. Prog Neurobiol

460

98: 250–264, 2012b.

461 462

Jensen MS, Cherubini E, Yaari Y. Opponent effects of potassium on GABAA-mediated postsynaptic inhibition in the rat hippocampus. J Neurophysiol 69: 764–771, 1993.

21

Journal of Neurophysiology JN-01128-2015 (Review)

463 464

Kaila K, Ruusuvuori E, Seja P, Voipio J, Puskarjov M. GABA actions and ionic plasticity in epilepsy. Curr Opin Neurobiol 26: 34–41, 2014.

465

Köhling R, D’Antuono M, Benini R, de Guzman P, Avoli M. Hypersynchronous ictal onset in

466

the perirhinal cortex results from dynamic weakening in inhibition. Neurobiol Dis 87: 1-10.

467

2016.

468

Lévesque M, Salami P, Gotman J, Avoli M. Two seizure-onset types reveal specific patterns

469

of high-frequency oscillations in a model of temporal lobe epilepsy. J Neurosci 32: 13264-

470

13272, 2012.

471 472

Lévesque M, Herrington R, Hamidi S, Avoli M. Interneurons spark seizure-like activity in the entorhinal cortex. Neurobiol Dis 87: 91-101, 2016.

473

Lieb JP, Walsh GO, Babb TL, Walter RD, Crandall PH. A comparison of EEG seizure

474

patterns recorded with surface and depth electrodes in patients with temporal lobe

475

epilepsy. Epilepsia 17: 137–160, 1976.

476 477 478 479 480

Lopantsev V, Avoli M. Participation of GABAA-mediated inhibition in ictal-like discharges in the rat entorhinal cortex. J Neurophysiol 79: 352–360, 1998a. Lopantsev V, Avoli M. Laminar organization of epileptiform discharges in the rat entorhinal cortex in vitro. J. Physiol 509: 785–796,1998b. Ogren JA, Bragin A, Wilson CL, Hoftman GD, Lin JJ, Dutton RA, Fields TA, Toga AW,

481

Thompson PM, Engel J, Staba RJ. Three-dimensional hippocampal atrophy maps

482

distinguish two common temporal lobe seizure-onset patterns. Epilepsia 50: 1361–1370,

483

2009.

484 485

Pacia SV, Ebersole JS. Intracranial EEG substrates of scalp ictal patterns from temporal lobe foci. Epilepsia 38: 642–654, 1997.

22

Journal of Neurophysiology JN-01128-2015 (Review)

486 487 488 489 490

Perreault P, Avoli M. Physiology and pharmacology of epileptiform activity induced by 4aminopyridine in rat hippocampal slices. J Neurophysiol 65, 771-785. 1991. Perucca P, Dubeau F, Gotman J. Intracranial electroencephalographic seizure-onset patterns: effect of underlying pathology. Brain 137: 183–196, 2014. Rutecki PA, Lebeda FJ, Johnston D. 4-Aminopyridine produces epileptiform activity in

491

hippocampus and enhances synaptic excitation and inhibition. J Neurophysiol 57: 1911-

492

1924, 1987.

493 494 495

Salami P, Lévesque M, Gotman J, Avoli M. Distinct EEG seizure patterns reflect different seizure generation mechanisms. J Neurophysiol 113: 2840–2844, 2015. Schevon CA, Weiss SA, McKhann G, Goodman RR, Yuste R, Emerson RG, Trevelyan AJ.

496

Evidence of an inhibitory restraint of seizure activity in humans. Nat Commun 3: 1060,

497

2012.

498 499

Shiri Z, Manseau F, Lévesque M, Williams S, Avoli M. Interneuron activity leads to initiation of low-voltage fast-onset seizures. Ann Neurol 77: 541–546, 2015.

500

Shiri Z, Manseau F, Lévesque M, Williams S, Avoli M. Activation of specific neuronal

501

networks leads to different seizure onset types. Ann. Neurol. 79(3): 354-65, 2016.

502

Singh S, Sandy S, Wiebe S. Ictal onset on intracranial EEG: Do we know it when we see it?

503 504

State of the evidence. Epilepsia 56: 1629–1638, 2015. Spanedda F, Cendes F, Gotman J. Relations between EEG seizure morphology,

505

interhemispheric spread, and mesial temporal atrophy in bitemporal epilepsy. Epilepsia 38:

506

1300–1314, 1997.

507 508

Spencer DD, Pappas CT. Surgical decisions regarding medically intractable epilepsy. Clin Neurosurg 38: 548–566, 1992.

23

Journal of Neurophysiology JN-01128-2015 (Review)

509 510 511

Spencer SS, Guimaraes P, Katz A, Kim J, Spencer D. Morphological patterns of seizures recorded intracranially. Epilepsia 33: 537–545, 1992. Staba RJ. Normal and Pathologic High-Frequency Oscillations [Online]. In: Jasper’s Basic

512

Mechanisms of the Epilepsies, edited by Noebels JL, Avoli M, Rogawski MA, Olsen RW,

513

Delgado-Escueta AV. National Center for Biotechnology Information (US)

514

http://www.ncbi.nlm.nih.gov/books/NBK98191/.

515

Timofeev I, Grenier F, Steriade M. The role of chloride-dependent inhibition and the activity of

516

fast-spiking neurons during cortical spike-wave electrographic seizures. Neuroscience

517

114: 1115-1132, 2002.

518

Toyoda I, Fujita S, Thamattoor AK, Buckmaster PS. Unit Activity of Hippocampal

519

Interneurons before Spontaneous Seizures in an Animal Model of Temporal Lobe

520

Epilepsy. J Neurosci 35: 6600–6618, 2015.

521 522 523

Traynelis SF, Dingledine R. Potassium-induced spontaneous electrographic seizures in the rat hippocampal slice. J Neurophysiol 59: 259-276, 1988. Trombin F, Gnatkovsky V, de Curtis M. Changes in action potential features during focal

524

seizure discharges in the entorhinal cortex of the in vitro isolated guinea pig brain. J

525

Neurophysiol 106: 1411–1423, 2011.

526

Truccolo W, Donoghue JA, Hochberg LR, Eskandar EN, Madsen JR, Anderson WS,

527

Brown EN, Halgren E, Cash SS. Single-neuron dynamics in human focal epilepsy. Nat

528

Neurosci 14: 635–641, 2011.

529 530

Uva L, Avoli M, de Curtis M. Synchronous GABA-receptor-dependent potentials in limbic areas of the in-vitro isolated adult guinea pig brain. Eur J Neurosci 29: 911–920, 2009.

24

Journal of Neurophysiology JN-01128-2015 (Review)

531

Uva L, Breschi GL, Gnatkovsky V, Taverna S, de Curtis M. Synchronous inhibitory potentials

532

precede seizure-like events in acute models of focal limbic seizures. J Neurosci 35: 3048–

533

3055, 2015.

534 535 536 537

Velasco AL, Wilson CL, Babb TL, Engel J Jr. Functional and anatomic correlates of two frequently observed temporal lobe seizure-onset patterns. Neural Plast 7: 49–63, 2000. Viitanen T, Ruusuvuori E, Kaila K, Voipio J. The K+-Cl cotransporter KCC2 promotes GABAergic excitation in the mature rat hippocampus. J Physiol 588: 1527–1540, 2010.

538

Wendling F, Hernandez A, Bellanger JJ, Chauvel P, Bartolomei F. Interictal to ictal transition

539

in human temporal lobe epilepsy: insights from a computational model of intracerebral

540

EEG. J Clin Neurophysiol. 22: 343-356, 2005.

541 542 543

Whittington MA, Traub RD, Jefferys JG. Erosion of inhibition contributes to the progression of low magnesium bursts in rat hippocampal slices. J Physiol 486 ( Pt 3): 723–734, 1995. Wu S, Kunhi Veedu HP, Lhatoo SD, Koubeissi MZ, Miller JP, Lüders HO. Role of ictal

544

baseline shifts and ictal high-frequency oscillations in stereo-electroencephalography

545

analysis of mesial temporal lobe seizures. Epilepsia 55: 690–698, 2014.

546

Yekhlef L, Breschi GL, Lagostena L, Russo G, Taverna S. Selective activation of

547

parvalbumin- or somatostatin-expressing interneurons triggers epileptic seizurelike activity

548

in mouse medial entorhinal cortex. J Neurophysiol 113: 1616–1630, 2015.

549

Ylinen A, Bragin A, Nadasdy Z, Jando G, Szabo I, Sik A, Buzsaki G. Sharp wave-associated

550

high-frequency oscillation (200 Hz) in the intact hippocampus: network and intracellular

551

mechanisms. J Neuirosci 15: 30–46, 1995.

552

Zhang ZJ, Koifman J, Shin DS, Ye H, Florez CM, Zhang L, Valiante TA, Carlen PL.

553

Transition to seizure: ictal discharge is preceded by exhausted presynaptic GABA release

554

in the hippocampal CA3 region. J Neurosci 32: 2499–2512, 2012.

25

Journal of Neurophysiology JN-01128-2015 (Review)

555 556 557 558

Ziburkus J, Cressman JR, Barreto E, Schiff SJ. Interneuron and pyramidal cell interplay during in vitro seizure-like events. J Neurophysiol 95: 3948–3954, 2006. Zuckermann EC, Glaser GH. Hippocampal epileptic activity induced by localized ventricular perfusion with high-potassium cerebrospinal fluid. Exp Neurol 20: 87-110, 1968.

26

Journal of Neurophysiology JN-01128-2015 (Review)

559

Legends to figures

560

Figure 1 - Electrographic features of LVF and HYP seizure onsets in humans (A) as well

561

as in in vivo (B) and in vitro (C) experimental models. Traces shown in A were obtained

562

from hippocampal recordings performed during pre-surgical intracranial monitoring with depth

563

macro-electrodes in a patient with a focal cortical dysplasia of the temporal lobe (left trace) and

564

in a patient with malformation (i.e., altered rotation) of the hippocampus that was associated

565

with hippocampal sclerosis verified on magnetic resonance and histological examinations (right

566

trace). The asterisk in the LVF sample identifies the single, “sentinel’ spike that can occur at

567

onset, while asterisks in the HYP sample highlight the initial series of large amplitude spikes at

568

approx. 1 Hz; note in this case that spikes at a similar frequency (indicated by arrow-heads)

569

reappear during the initial part of the seizure. Traces in B were recorded from the CA3 area of

570

the hippocampus in a pilocarpine-treated epileptic rat; note that both types of seizure onsets

571

could be recorded from the same animal; asterisks identify specific EEG events as described for

572

the recordings shown in panel A. Traces in C were obtained from experiments performed with

573

rat brain slices maintained in vitro and bathed in medium containing 4-aminopyridine; the LVF

574

seizure-like onset shown on the left was recorded from the medial entorhinal cortex while the

575

HYP seizure on the right was obtained from the perirhinal cortex; these in vitro recordings were

576

perfomed with DC amplifiers and thus seizure-like discharges are associated with robust

577

negative-going shifts. Traces shown in A were kindly provided by Dr. Stefano Francione and Dr.

578

Laura Tassi of the Claudio Munari Epilepsy Surgery Center, Milano, Italy.

579 580

Figure 2 - Analysis of HFOs occurring before and during LVF and HYP seizures in

581

pilocarpine-treated epileptic rats. A: The left panel shows recordings from the CA3 region of

582

a pilocarpine-treated rat during an LVF seizure. The signal is shown in the wideband, ripple (80-

27

Journal of Neurophysiology JN-01128-2015 (Review)

583

200 Hz) and fast ripple (250-500 Hz) frequency ranges. Note the occurrence of a ripple (boxed

584

dotted portion of the trace) with no co-occurrence of a fast ripple. The right panel shows

585

recording from CA3 during an HYP seizure in a pilocarpine-treated animal. Note in this case the

586

occurrence of a fast ripple (boxed dotted portion of the trace). B: Temporal distribution of HFOs

587

recorded during 18 LVF seizures in pilocarpine-treated animals. Note that ripples predominate

588

over fast ripples, especially after seizure onset. C: Temporal distribution of HFOs recorded

589

during 21 HYP seizures in pilocarpine-treated animals. Note the high occurrence of fast ripples

590

compared to ripples. Rates of ripples and fast ripples were compared using non-parametric

591

Wilcoxon signed rank tests followed by Bonferroni-Holm corrections for multiple comparisons (*

592

p < 0.05). Data shown in this figure were obtained during the experiments published by

593

Lévesque et al. (2012)

594 595

Figure 3 - Field, [K+]o, and intracellular recordings obtained in vitro during LVF onset

596

seizure-like dischargers. A: Simultaneous field (Field) and intracellular (-68 mV) recordings

597

obtained from the entorhinal cortex of a rat brain slice during bath application of 4-

598

aminopyridine; in this and following panels the dotted lines indicate the resting membrane

599

potential of the neuron recorded intracellularly. Traces shown in the insert were obtained from a

600

different neuron during active depolarization with intracellular current from a resting membrane

601

potential at approx. - 62 mV. B: Field and [K+]o activities recorded during application of 4-

602

aminopyridine from the deep layers of the rat entorhnal cortex; the dotted line indicates the [K+]o

603

base level. C: Simultaneous field (Field) and intracellular (-65 mV) recordings obtained from a

604

principal cell in the entorhinal cortex during the perfusion of 50 µM bicuculline for 3 minutes in

605

the in vitro isolated guinea pig brain. The spectrogram on the top illustrates the fast activity at

606

around 20 Hz at the onset of the seizure-like event. The fast activity occurring at the onset of the

28

Journal of Neurophysiology JN-01128-2015 (Review)

607

seizure-like discharge is expanded in the insert where the principal cell was depolarized by

608

intracellular injection of steady current. D: Field and intracellular recordings of the response

609

induced by lateral olfactory tract stimulation in an entorhinal cortex neuron with a resting

610

membrane potential of -60 mV. Note that the brief IPSPs that correlate to the fast activity

611

oscillations have reversal potentials similar to what is seen with the lateral olfactory tract-evoked

612

IPSP, E: Simultaneous field and intracellular recordings from a principal cell in the perirhinal

613

cortex of a rat brain slice during bath application of 4-aminopyridine; note that ictal discharge

614

onset is characterized by preictal spiking acceleration as well as that both interictal (arrow) and

615

“preictal” discharges consist of depolarizations with action potential bursts followed by a

616

hyperpolarizing potential that progressively decreases in amplitude (arrow heads) and coincides

617

with more intense action potential bursting. F: Simultaneous field and [K+]o recordings obtained

618

in the deep layers of the perirhinal cortex during 4-aminopyridine application. Note the

619

progressive increases in [K+]o that accompany the preictal spikes up to ictal discharge initiation

620

that coincides with values larger than 6.3 mM. Data shown in this figure were obtained during

621

experiments that have been published by Avoli et al. (1996, 2013), Biagini et al. (2013);

622

Gnatkovsky et al. (2008); Köhling et al. (2016); Lopantsev and Avoli (1998a and b); Trombin et

623

al. (2011); and Uva et al. (2015).

624 625

Figure 4 – Optogenetic activation of interneurons or principal cells leads to LVF or HYP

626

ictal discharges respectively. A: In a, LVF onset ictal discharges occurring spontaneously

627

during bath application of 4-aminopyridine (top) and during parvalbumin positive-interneuron

628

light stimulation (bottom); one event under each condition is further expanded to reveal the

629

onset pattern. Stimuli were 1 s long and delivered at 0.2 Hz for 30 s. Note that in both

630

spontaneous and stimulated events, the ictal discharge is preceded by one or two negative-

29

Journal of Neurophysiology JN-01128-2015 (Review)

631

going interictal field potentials (arrows). In b, plots of the average rate of ripples and fast ripples

632

occurring during spontaneous (top) and optogenetically stimulated (bottom) ictal discharges (n =

633

10 events were used for both plots). Note that ripple rates are significantly higher than fast ripple

634

rates at the onset of both LVF discharges (* p < 0.01). B: In a, spontaneous LVF ictal

635

discharges occurring during bath application of 4-aminopyridine are shown in the top panel

636

while those induced by optogenetic stimulation of calcium/calmodulin-dependent protein kinase

637

II-principal cells are illustrated in the bottom; light pulses were 20 ms long and were delivered at

638

2 Hz for 30 s; note that this procedure triggers ictal discharges preceded by repeated spiking

639

(arrows), a pattern that is characteristic of HYP onset events. In b, plots of the average rate of

640

ripples and fast ripples occurring during spontaneous (top) and optogenetically stimulated

641

(bottom) ictal discharges; note that the stimulated HYP events are characterized by higher fast

642

ripple rates at onset 10 events were used for both plots (* p < 0.01). Data shown in this figure

643

were obtained from the experiments published by Shiri et al. (2015, 2016).

644 645

Figure 5 – Flow diagram of the dysfunctional mechanisms leading to LVF and HYP

646

seizure onset. Experimental evidence taken into account for providing the mechanisms

647

involved in LVF seizure onset was provided in the following studies: Avoli et al., 1996;

648

Barbarosie et al., 2000; Gnatkovsky et al., 2008; Trombin et al., 2011; Grasse et al., 2013;

649

Lévesque et al., 2016; Lopantsev et al., 1998a; Uva et al., 2015; Ziburkus et al., 2006. Data

650

identifying the mechanisms leading to HYP seizure onset originate from the following studies:

651

Lopantsev and Avoli, 1998b; Derchanski et al., 2006; Zhang et al., 2012; Köhling et al., 2016.

30