Review | Disorders of the Nervous System The Kainic acid models of Temporal Lobe https://doi.org/10.1523/ENEURO.0337-20.2021

Cite as: eNeuro 2021; 10.1523/ENEURO.0337-20.2021 Received: 3 August 2020 Revised: 14 January 2021 Accepted: 24 January 2021

This Early Release article has been peer-reviewed and accepted, but has not been through the composition and copyediting processes. The final version may differ slightly in style or formatting and will contain links to any extended data.

Alerts: Sign up at www.eneuro.org/alerts to receive customized email alerts when the fully formatted version of this article is published.

Copyright © 2021 Rusina et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license, which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

1

2 Manuscript Title Page

3 1. Manuscript title: The Kainic acid models of Temporal Lobe Epilepsy

4 2. Abbreviated title: Kainic Acid Review in Epilepsy: Perspective

5 3. Author list:

6 • Evgeniia Rusina*1

7 • Christophe Bernard*1

8 • AdamWilliamson1,2

9 1Institute de Neurosciences des Systèmes (INS), INSERM, UMR_1106, Aix-Marseille Uni

10 versité, 27 Bd Jean Moulin 13005, Marseille, France

11 2Laboratory of Organic Electronics, Campus Norrköping, Linköping University, 581 83,

12 Norrköping, Sweden

13 4. Author contribution: AW conceived the idea; ER and CB wrote the paper; ER* and CB* are

14 equally first contributing authors.

15 5. Correspondence should be addressed to [email protected]

16 Faculté de Médecine, 27 Boulevard Jean Moulin, 13005, Marseille, France

17 6. Number of figures: 5

18 7. Number of tables: 6

19 8. Number of multimedia: 0

20 9. Number of words for abstract: 86

21 10. Number of words for significance statement: 90

22 11. Number of words for introduction: 377

23 12. Number of words for discussion: -

24 13. Acknowledgements: AW acknowledges funding from the European Research Council (ERC)

25 under the European Union’s Horizon 2020 research and innovation program (grant agreement No

26 716867). AW and ER acknowledge funding from the Excellence Initiative of Aix- Marseille Universi-

27 ty— A*MIDEX, a French “Investissements d’Avenir” program. AW acknowledges support from the

28 Knut and Alice Wallenberg Foundation.

29 14. Authors report no conflict of interests

30 15. Funding sources:

31 • Excellence Initiative of Aix- Marseille University—A*MIDEX

1

32 • European Research Council (ERC) under the European Union’s Horizon 2020 research and in-

33 novation program (grant agreement No 716867)

34 The Kainic Acid Models of Temporal Lobe Epilepsy

35

36 Evgeniia Rusina*, Christophe Bernard*, Adam Williamson

37

38

39 Institute de Neurosciences des Systèmes (INS), INSERM, UMR_1106, Aix-Marseille Université

40 Marseille, France

41 * equally contributing first authors

42 correspondance: [email protected]

43 Abstract

44

45 Experimental models of epilepsy are useful to identify potential mechanisms of epileptogenesis,

46 genesis, comorbidities, and treatment efficacy. The kainic acid (KA) model is one of the

47 most commonly used. Several modes of administration of KA exist, each producing different ef-

48 fects in a strain, species, gender, and age-dependent manner. In this review, we discuss the ad-

49 vantages and limitations of the various forms of KA administration (systemic, intrahippocampal,

50 and intranasal), as well as the histological, electrophysiological, and behavioral outcomes in differ-

51 ent strains and species. We attempt a personal perspective and discuss areas where work is

52 needed. The diversity of KA models and their outcomes offers researchers a rich palette of pheno-

53 types, which may be relevant to specific traits found in patients with temporal lobe epilepsy.

54

55 Keywords: Kainic acid, , Mice models of temporal lobe epilepsy, EEG

56 2

57

58

59

60 Significance statement

61

62 This review aims to help researchers use a knowledge-based approach to study specific aspects

63 of human epilepsy phenotypes. We focus on the Kainic Acid model of temporal lobe epilepsy in

64 rodents, presenting it as a set of sub-models, describing the various administration routes, and the

65 differences in outcome between species, strain, age, and sex. We have reviewed more than 200

66 research articles, summarizing the data with a ready-to-use structure.

67

68 Introduction

69

70 Mesial temporal lobe epilepsy (MTLE) is the most common type of partial epilepsy in adults

71 (J.Engel, 2001). It is characterized by recurrent spontaneous , often resistant to drug treat-

72 ments (Engel et al., 2012). Numerous alterations have been reported in patients with MTLE, includ-

73 ing hippocampal sclerosis (HS) and cell death (Berkovic et al., 1991), mossy fiber sprouting (Sutula

74 et al., 1989), reorganization of hippocampal interneuronal networks (Muñoz et al., 2004; Tóth et al.,

75 2010; Sloviter et al.,1991), alterations in neuropeptide signaling (Vezzani et al., 2004; Kharlamov et

76 al., 2007) and synaptic transmission regulation (Casillas-Espinosa et al., 2012), granular cell disper-

77 sion and gliosis (Blümcke et al., 2006), blood-brain-barrier dysfunction and angiogenesis (Crespel et

78 al., 2007).

79

80 A striking feature of MTLE in humans is its heterogeneity. There are various forms of HS in pa-

81 tients, which led to a consensus classification based on histopathological differences (Blümcke et

82 al., 2013). Likewise, there is a large diversity in terms of semiologic and electrophysiologic features

83 in temporal lobe seizures, including regional epileptogenicity and whether or not seizures second-

84 arily generalize (Maillard et al., 2004, Barba et al., 2007, Bartolomei et al., 2008). A large variability

85 also exists in hypometabolism in patients with TLE (Guedj et al., 2015). Finally, the diversity be- 3

86 tween patients is even more important when considering comorbidities (i.e. conditions co-occurring

87 with seizures), such as cognitive deficits, anxiety, and depression (Holmes, 2015, Krishnan, 2020,

88 de Barros Lourenço et al., 2020).

89

90 These considerations provide the context of this review: we need a wide array of phenotypes in

91 experimental models to ask specific questions that may be relevant to subsets of patients. It is rea-

92 sonable to propose that genetic and epigenetic differences strongly contribute to phenotypic diver-

93 sity in patients. In contrast, basic research uses inbred rodents living in a "stable" environment in

94 animal facilities. This facilitates the task of researchers, increasing the reproducibility of the results.

95 However, a given model may only be relevant to a specific phenotype in patients. Fortunately,

96 several experimental models of epilepsy have been developed, originally to mimic MTLE, including

97 the pilocarpine models, the electrode-based kindling models, and the kainic acid (KA) models. Alt-

98 hough homology between a rodent model and human epilepsy cannot be claimed, experimental

99 models must reproduce the main symptom: spontaneous recurrent seizures (SRSs). If other patho-

100 logical traits (e.g., mossy fiber sprouting or HS) or comorbidities (e.g., cognitive deficits and de-

101 pression) exist, it may be possible to achieve a degree of granularity compatible with the diversity

102 of phenotypes found in patients. This review will focus on the KA rodent model, its advantages and

103 limitations, methods of administration, histological, electrophysiological, and behavioral outcomes.

104 As we will see, this model alone is characterized by a wide range of phenotypes.

105

106 The ideal situation would be to map the diversity of KA phenotypes with that of MTLE phenotypes.

107 Unfortunately, most of the time, the existing literature does not provide enough information. For

108 example, the part of the hippocampus that is resected in patients corresponds to the ventral hippo-

109 campus in rodents. A minority of experimental studies (apart from intra/supra hippocampal injec-

110 tions) distinguish the ventral from the dorsal hippocampus, despite the fact that clear differences

111 appear in experimental epilepsy when specifically studied (do Canto et al., 2020, Canto et al.,

112 2020, Arnold et al., 2019, Evans & Dougherty, 2018, Isaeva et al., 2015, Ekstrand et al., 2011,

113 Bernard et al., 2007, Brancarti et al., 2020). Most electrophysiological recordings in rodent models

114 are performed in the dorsal hippocampus, or supradurally, essentially capturing the activity of the 4

115 dorsal hippocampus via volume conduction, not what happens in the ventral hippocampus. Like-

116 wise, morphological studies rarely distinguish the ventral and dorsal hippocampi in terms of cell

117 death, sprouting etc. The use of common data elements in future studies will be particularly useful

118 to try correlate experimental models and patient data.

119

120 The last cautionary note relates to the way rodent data are analyzed: results are averaged, which

121 assumes that the model will produce similar results with some variation. The clinic teaches us oth-

122 erwise. If we just consider the electrophysiological signature of seizures, patients with mesial tem-

123 poral atrophy/sclerosis can display low-voltage fast or hypersynchronous seizures (e.g. Table 1 in

124 Perucca et al., 2014). Individual patients can display several types of seizures (Saggio et al.,

125 2020). Rodents, even from the same litter, are not clones. They are also individuals(Gomez-Marin

126 & Ghazanfar, 2019). They can react very differently to a given insult, including epileptogenesis,

127 and produce different phenotypes (Medel-Matus et al., 2017, Becker et al., 2019, Becker et al.,

128 2015). Most studies presented hereafter did not report major differences within experimental

129 groups, but the data may be there. Looking for differences within a given batch of experimental an-

130 imals may constitute a very promising line of research, in keeping with a highly clinically relevant

131 question: the development of personalized medicine.

132

133 Summarizing the data over the past decades is not an easy task. There was no standardized way

134 to report experimental protocol and analysis methods, which prevents a real comparison between

135 labs. Although 24h video/EEG-recordings are now routinely used in experimental rodent models,

136 there were not as widespread in the past. Notwithstanding their importance, some studies used

137 short recording sessions (a few hours per day, during working days) and could potentially have

138 missed some epileptic activity. Some studies only used visual observation.

139 As mentioned previously, there is still no agreement between scientists on how to define a seizure

140 in a rodent. While the Racine scale provides an attempt to organize and quantify behavioral sei-

141 zures, subclinical seizures, frequent spikes, and bursts may be present on the EEG without overt

142 behavioral correlates. Considering the differences in seizure profile between rats and mice and the

5

143 variability between labs, it is really hard to draw clear conclusions from the huge pool of data col-

144 lected through so many years.

145

146 The problems of standardization also apply to neuroanatomical and functional morphological stud-

147 ies. Many confounding factors can affect the interpretation of the observations. The counting

148 method is important. Stereological techniques were not available in the past. Hence it is difficult to

149 compare studies even when the same experimental model of epilepsy is used. Furthermore, the

150 hippocampus is heterogeneous in terms of structure, cell distribution, and properties along both

151 septo-temporal and longitudinal axes, and even within the pyramidal cell layer (Malik et al., 2016,

152 Sotiriou et al., 2005, Pandis et al., 2006, Thompson et al., 2008, Dong et al., 2009, Mizuseki et al., 2011,

153 Sharif et al., 2021).

154 It is not always clear which regions/subfields are being investigated, e.g., to assess cell death.

155 Morphological alterations may be regionalized. Human data demonstrates even more complex

156 changes as basket innervation is patchy from one subfield to the next in a given patient (Arellano

157 et al., 2004). Such heterogeneity is also found in multidrug transporters' expression, which can be

158 very different from one slice to the next obtained after neurosurgery (Sandow et al., 2015). In animal

159 models, we tend to assume that neuroanatomical changes are homogeneous, i.e., that the meas-

160 urements performed in a few sections can be generalized to the structure. Many other brain re-

161 gions are also characterized by spatial gradients, which should be considered when assessing cir-

162 cuit alterations. Likewise, since it is cumbersome to do, few studies engage in a global assessment

163 of anatomical alterations in different regions in experimental models. Much work is needed to ad-

164 dress. It would be helpful to use common data elements to standardize protocols and facilitate

165 comparisons between studies (Lapinlampi et al., 2017).

166

167 We have attempted to be as neutral as possible in our evaluation, relying on the interpretation pro-

168 vided in each study and, whenever relevant, providing our own experience. In the following, we

169 describe the different phenotypes found in various KA models. Given the wide heterogeneity found

170 in patients, all models described hereafter reproduce some features found in some patients. Most

171 experimental models share similarities with properties found in a majority of MTLE patients, includ- 6

172 ing electrophysiological (seizures, interictal spikes, high frequency oscillations) and morphological

173 (granule cell dispersion, mossy fiber sprouting) signatures. We will not take a position regarding a

174 possible relevance to features found in patients, because there is no prototypical signature of

175 MTLE at the morpho-functional level. For example, hippocampal sclerosis is found in only 60% of

176 MTLE cases (Thom, 2014), and patients with MTLE can have low-voltage fast or hypersynchro-

177 nous seizures (Perucca et al., 2014). The choice of the model should be tailored to the scientific

178 question, which is laboratory-specific. We describe and provide tables listing the various observa-

179 tions made in different models. These tables can be used for knowledge-based approaches. For

180 example, if one is interested in the mechanisms of low-voltage fast seizures in patients with end

181 folium sclerosis, identify an experimental model reproducing such features (if available) and study

182 it. We also highlight issues that need to be investigated, which could constitute interesting oppor-

183 tunities for young investigators.

184

185 1. Models of TLE: a brief overview

186

187 The main goal of creating a reliable animal model of epilepsy is to develop a chronic condition,

188 which is consistent and efficient in generating SRSs. Several methods have been developed, all of

189 them are based on different approaches, but they generally use an initial brain insult. It is worth

190 noting that despite the fact that all these methods produce reliable epilepsy models, they do not

191 always show the same phenotypes, such as neuronal cell death, neuropathological alterations and

192 epileptogenesis. The reasons behind such heterogeneity are not clear. Every single model has its

193 nuances, which should be seen as a strength. We provide a brief information about other TLE

194 models to stress further the diversity that can be achieved in terms of phenotypes. We did not at-

195 tempt to compare them to the kainate models, which was beyond this review's scope. We are also

196 not discussing genetic conditions and transgenic strains that exhibit a TLE-like phenotype (Toth et

197 al., 1995; Zeng et al., 2011), but it is worth mentioning that TLE may have a genetic predisposition

198 (Andermann et al., 2005). We do not want to convey the wrong impression that we favor the KA

199 model. All models are interesting as they enable researchers to tap into a rich repertoire of pheno-

7

200 types. Although this brief description of other than KA models does not do justice to them, we start

201 this review with these models to highlight their diversity and richness.

202 The administration of various chemoconvulsants either systemically or directly into the brain is

203 commonly used to trigger epileptogenesis, the process that leads to the development of SRSs. Pi-

204 locarpine and kainic acid are broadly used chemoconvulsants. Pilocarpine, a muscarinic acetylcho-

205 line receptor agonist, was first proposed to generate experimental epilepsy in rats by W. Turski

206 (Turski et al., 1983). Intraperitoneal injection of 100-400mg/kg pilocarpine triggers seizures starting

207 with motor, olfactory, and gustatory automatisms, which evolve into status epilepticus (SE). Histo-

208 logical examination of the brains revealed specific alterations throughout the hippocampal for-

209 mation, amygdala, thalamus, neocortex, olfactory cortex, and substantia nigra (Turski et al., 1983).

210 The most remarkable difference between the pilocarpine and the KA models is the rapidity of the

211 pilocarpine one. Neuronal damage is visible within 3 hours after pilocarpine-induced SE, while in

212 the case of KA, the brain lesion in the same areas is visible 8 hours after SE induction (Covolan &

213 Mello, 2000). Regarding the extension of cellular damage, both KA and pilocarpine models, inject-

214 ed systemically, produce an extensive extrahippocampal cell loss (Levesque & Avoli, 2013). In

215 mice, the pattern of cell loss is strain-dependent (Schauwecker et al., 2012). It is important to stress

216 that each lab used/uses different evaluation methods for cell counting and general damage as-

217 sessment, making comparisons difficult.

218

219 Mortality in the pilocarpine model appears higher. Lithium treatment 24 h prior to the convulsant

220 injection can significantly reduce mortality, leading to a different phenotype (Curia et al., 2008). Mul-

221 tiple reports indicate a strain difference in the pilocarpine model (McKhann et al., 2003, Curia et al.,

222 2008, Inostroza et al., 2011, Levesque & Avoli, 2013). For instance, cell damage and mortality rates

223 are more significant in Long-Evans and Wistar rat strains than Sprague-Dawley rats (Curia et al.,

224 2008; Inostroza et al., 2011). The Sprague-Dawley strain also demonstrates less prominent neu-

225 ronal damage than Wistar rats (Inostroza et al., 2011). Overall, the Long-Evans strain is the most

226 sensitive strain to pilocarpine, followed by Wistar; while the Sprague-Dawley strain exhibits mini-

227 mal sensitivity. The Dose-response to pilocarpine is similar in mice and rats; however, mice appear

228 to be more sensitive and display higher mortality rates than rats (Curia et al., 2008). Also, mice 8

229 treated with pilocarpine are more likely to develop SRSs than KA-treated ones (Levesque & Avoli,

230 2013). Generally speaking, several parameters are involved in phenotype generation in both pilo-

231 carpine and KA models: administration route, dosage, duration and severity of the initial insult, en-

232 vironmental conditions etc. Evidently, the characteristics of experimental animals play crucial role

233 as well. Species, strain, age and sex create multiple variations in the models and will be thoroughly

234 discussed later. As it is hard to navigate between so many factors, we have made an attempt to

235 summarize existing data in the tables, demonstrating the full diversity of the KA models. The pilo-

236 carpine model, not any less complex, is not discussed in details in the current review, but the

237 reader is welcome to read further literature about the model (Curia et al., 2008; Levesque & Avoli,

238 2013).

239

240 It is important to note that the initial experiments performed with KA or pilocarpine showed high

241 death rates. Status epilepticus (SE) was left unchecked, and some drugs used to try to stop it had

242 unwanted effects. Thus, it is difficult to compare the results obtained with these studies, as they

243 may have generated phenotypes different from those described nowadays. More recent TLE mod-

244 els show lower levels of fatality in both mice and rats, in particular when SE is monitored with EEG

245 recordings. In our hands, we noted that SE might start electrographically 5 min before behavioral

246 manifestations, particularly in mice. We also noted that aberrant electrographic activity continues

247 after the injection of the most optimal combination of drugs used to stop SE (Brandt et al., 2015), up

248 to the next day following SE induction. Although the drugs have relaxant effects (no or few motor

249 events), the aberrant activity continues, hence the necessity to monitor continuously with EEG re-

250 cordings. Since SE triggers epileptogenesis, it is difficult to standardize its duration and severity,

251 which could be a variability source within the same group of injected animals. Finally, we observed

252 seasonal variability. In our hands, rodents seem to be more sensitive to pilocarpine and KA, with a

253 greater death rate during the summer, another possible variability source.

254 In terms of behavioral performance, spatial memory seems to be affected significantly in the pilo-

255 carpine model. However, pilocarpine-treated animals show reduced anxiety levels compared to the

256 KA-treated rats (Inostroza et al., 2011).

257 9

258 The injection of tetanus toxin into the hippocampus is also used to trigger epileptogenesis. The first

259 mention of this procedure dates as early as the nineteenth century, demonstrating that intracere-

260 bral injection of the toxin causes seizures in experimental animals (Roux and Borrel, 1889). Unlike

261 KA, tetanus toxin does not elicit SE after its injection but efficiently generates SRSs within 2-21

262 days post-injection, which usually ceases after a few weeks (Jefferys & Walker, 2006). A standard

263 dose causes nearly 30% pyramidal cell loss in CA1 in the unilateral hippocampus and 10% at dis-

264 tant sites. Moreover, loss of the dendritic spines in CA3 pyramidal neurons is also reported (Benke

265 & Swann, 2004).

266

267 The next common approach for modeling TLE in rodents includes various types of electrical stimu-

268 lation. The most widely used model is the kindling model, in which an electrode is implanted into

269 the brain, and a seizure focus is created by repeated electrical stimulation. The term “kindling” was

270 first proposed by G. Goddard and his colleagues (Goddard et al., 1967), where, after several exper-

271 imental trials, they demonstrated that daily electrical stimulation leads to the development of an

272 epileptic focus in the brain, creating generalized convulsions and permanent changes in brain tis-

273 sue (Goddard et al., 1969). The kindling model has been extensively used targeting various regions

274 such as the hippocampal formation (Kairiss et al., 1984; Palizvan et al., 2001), the piriform cortex

275 (Loscher & Ebert, 1996), and the perirhinal cortex (McIntyre et al., 1993). The electrical stimulation

276 model's primary hallmark is the gradual development of an epileptic activity, as compared to the

277 chemical administration, in which the initial brain insult leads to the development of chronic epilep-

278 sy after a latent period. The kindling model is sometimes described as a "functional" model due to

279 the absence of gross morphological damage, which is characteristic of chemical models (Morimoto

280 et al., 2004) and a general failure to induce chronic SRS (Goddard et al., 1983). Thus, it is rather a

281 model of epileptogenesis but not of epilepsy. A variation based on an optogenetic approach (rather

282 than electric stimulation) called “optokindling” has been introduced recently (Cela et al., 2019). Alt-

283 hough it was used to produce a model for neocortical epilepsy, the same technique may create a

284 TLE model. The main advantage of this kindling model is the ability to trigger seizures on demand.

285 At more advanced kindling stages, SRSs can occur (Wada et al., 1975).

10

286 Another way to obtain SRSs using the kindling approach is to trigger SE. Lothman and Bertram

287 developed a model of continuous hippocampal stimulation to induce chronic epilepsy (Lothman et

288 al., 1989). This model, named self-sustaining limbic status epilepticus (SSLSE), triggers SE and

289 thus epileptogenesis leading to the occurrence of SRSs. A similar model was proposed by

290 Mazarati et al., 1998, who suggested a brief 15-30 min perforant path stimulation (previously used

291 as a kindling model (Sloviter, 1987)), which led to continuous self-sustained SE and the develop-

292 ment of SRSs.

293

294 Most chemical and electrical models use SE to trigger epileptogenesis. It is reasonable to propose

295 that SE is the main determinant of epileptogenesis. However, causality has not been clearly estab-

296 lished. This issue could be addressed with optogenetics, which may be used to try to stop SE soon

297 after its onset.

298

299 One of TLE's most common causes in patients is traumatic brain injury (TBI) (Pitkänen et al.,

300 2006). Various TBI models have been developed, in particular, the fluid percussion injury model

301 (D’Ambrosio et al., 2004). A single severe injury produced by fluid percussion injury device is suffi-

302 cient to trigger epileptogenesis in rodents with some similarity with human TLE, including morpho-

303 logical alterations such as neuronal loss in the hippocampus and mossy fiber sprouting

304 (Kharatshvili et al., 2006; Pitkanen & Immonen, 2014).

305

306 The previous models involve brain insults triggered in adult animals. Insults occurring during de-

307 velopment can also lead to TLE later in life. For example, febrile seizures are a risk factor for TLE

308 development in humans (French et al., 1993). Hyperthermia-induced seizure models have been de-

309 veloped (Holtzman et al., 1981). Dubé et al., 2010 demonstrated that prolonged 70 min febrile sei-

310 zures in P11 rat pups could trigger epileptogenesis.

311

312 Neonatal hypoxia can also constitute an initial insult, resulting in TLE later in life (Jensen et al.,

313 1991). It is linked to the observation that hypoxic encephalopathy is the most common cause of ne-

314 onatal seizures (Aicardi et al., 1970). Exposing P10-P12 pups to graded global hypoxia (7%-4% 11

315 oxygen) for 15 minutes triggers epileptogenesis (Rakhade et al., 2011). Morphological changes in-

316 clude mossy fiber sprouting, and electrophysiological analysis shows increased excitability, facili-

317 tated long-term potentiation induction, and longer afterdischarges.

318 The developing brain is not a miniature adult brain. Hence, specific models need to be developed

319 to study that occur during development. The same issue applies to the aging brain, for

320 which mechanisms of epileptogenesis and seizure genesis may be specific to a given age. These

321 are key questions, but they were beyond the scope of the present review.

322

323 There are various ways to trigger epileptogenesis in the brain, each leading to slightly different

324 phenotypes. We will now focus on the kainic acid model.

325

326 2. The kainic acid model

327

328 Kainic acid, a cyclic analog of L-glutamate and an agonist of the ionotropic KA receptors, was first

329 reported to damage hippocampal pyramidal neurons by Nadler et al., 1978. However, the use of KA

330 as a model for epilepsy was first introduced by Ben-Ari and his colleagues (Ben-Ari & Lagowska,

331 1978; Ben-Ari et al., 1979), who performed a unilateral intra-amygdaloid injection of KA in unanaes-

332 thetized non-paralyzed rats and observed focal seizures evolving into SE as the dosage increased.

333 Moreover, the histological findings revealed neuronal degeneration and gliosis in the CA3 field of

334 the hippocampus. These, and other experiments, suggested using KA as a tool to model TLE in

335 rodents. The injection of KA will lead to the activation of its cognate receptors.

336

337 3.1. Kainic acid receptors

338

339 Extensive biomolecular research has provided us with information about the localization of kainic

340 acid receptors (KARs) in the mammalian brain. Experiments on KAR mapping showed that these

341 receptors could be found at different levels of expression throughout the brain, including the ento-

342 rhinal cortex (Patel et al., 1986), cerebellum (Wisden & Seeburg, 1993), amygdala (Rogawski et al.,

343 2003), basal ganglia (Jin & Smith, 2011) and the hippocampus in which they are particularly abun- 12

344 dant (Bloss & Hunter, 2010). Kainic acid receptors belong to a family of ionotropic glutamate recep-

345 tors, along with AMPA and NMDA receptors. They can be pre- or postsynaptic. Presynaptic KARs

346 act bidirectionally, performing an excitatory action through their ionotropic activity and inhibition via

347 a “non-canonical” metabotropic signaling (Lerma & Marques, 2013). Postsynaptic receptors con-

348 tribute to excitatory neurotransmission (Huettner, 2003). KA receptors can also control GABAergic

349 neurotransmission, both pre- and post-synaptically (Cossart et al., 2001; Cossart & Bernard, 2001).

350 There are five known types of KA receptor subunits: GluR5 (GluK1), GluR6 (GluK2), GluR7 (GluK3),

351 KA1 (GluK4), and KA2 (GluK5). Some subunits are highly expressed in the hippocampus. The

352 GluK4 subunit is almost exclusively found in the CA3 hippocampal field, whereas its expression in

353 CA1 is limited (Bahn et al., 1994; Darstein et al., 2003). GluK5 subunits are expressed in both CA1

354 and CA3 fields (Bahn et al., 1994) and in the cortex and striatum (Gallyas et al., 2003). The fact that

355 these two subunits have a high affinity for kainic acid (dissociation constant of 5-15 nM) and are

356 mainly concentrated in CA3 may explain the pattern of excitotoxic damage found in this region

357 (Fernandes et al., 2009) and suggests that they are likely responsible for neuronal cell death.

358 GluK1 are found in the hippocampus's CA3 field (Bahn et al., 1994), and GluK2 are highly ex-

359 pressed in both CA1 and CA3 (Bloss & Hunter, 2010). Although both subunits have low affinity for

360 KA (KD of 50-100 nM), GluK1 knock-out mice show increased susceptibility to the epileptogenic

361 effect of KA, while GluK2 ablation prevents the generation of epileptiform discharges (Fisahn et al.,

362 2004). GluK2-KO mice are also less sensitive to KA's epileptogenic effect (Mulle et al., 1998). Con-

363 versely, overexpression of GluK2 by HSVGluR6 viral vector injection leads to seizure induction and

364 hyperexcitabilty (Telfeian et al., 2000). Furthermore, Ullal et al., 2005 reported that the GluR7 subu-

365 nit, which has the lowest affinity to glutamate, is down-regulated by KA-induced seizures in the

366 long term. These findings suggest that the KA administration will target all types of kainate recep-

367 tors, with a unique effect in the hippocampus. Interestingly, the GluK1 subunit has been associated

368 with acute seizure induction (Fritsch et al., 2014), since the acute effect is believed to be mediated

369 by regulation of inhibition and GluK1 subunits are especially abundant in hippocampal interneurons

370 (Paternain et al., 2000). Chronic epilepsy, however, likely develops due to GluK2 and GluK5 subu-

371 nits, which are found in the newly formed KARs within the mossy fiber network (Artinian et al.,

372 2015, Pinheiro et al., 2013). Increased expression of GluK4 has been found in the brains of patients 13

373 with refractory TLE (Lowry et al., 2013). Additionally, GluK4 knockout mice demonstrated full neu-

374 roprotection in the CA3 area of the hippocampus following administration of KA (Das et al., 2012).

375 Thus, it is reasonable to assume that GluK4 modulates KA-induced neurodegeneration.

376 It could be interesting to determine which KARs in which cells are the main triggers of epileptogen-

377 esis. The recent development of photoswitchable regulators of ligand-gated channels

378 (Bregestovski et al., 2018) could enable a tight control of specific KARs in a cell type and brain re-

379 gion-dependent manner.

380

381 3.2. Mechanism of action

382

383 Excitotoxicity refers to a process in which neurons experience severe damage to the point of cell

384 death due to overstimulation by excitatory neurotransmitters such as glutamate (Dong et al., 2009).

385 The mechanism behind this includes a cascade of molecular interactions that lead to osmotic im-

386 balance, excessive depolarization, and, eventually, rupture of the postsynaptic membrane (Beck et

387 al., 2003). Several mechanisms are at play. A central one involves the intracellular accumulation of

388 Ca2+, following the excessive activation of glutamate receptors (Fig. 1D). The rise in Ca2+ can

389 strongly impact mitochondria and the endoplasmatic reticulum (Friedman et al., 2006). Elimination

390 of intracellular Ca2+ or blocking its influx into mitochondria can diminish cellular sensitivity to apop-

391 totic stimuli (Lee et al., 2009). The Na+ and Cl− ions are also involved since their removal from the

392 extracellular space stop neurodegeneration (Chen et al., 1998). Finally, extracellular K+ is also en-

393 gaged in KA-induced excitotoxicity (Ha et al., 2002).

394

395 Oxidative stress also plays a central role in cell death in the context of excitotoxic damage. The

396 excess of glutamate initiates reactive oxygen species (ROS) formation, which leads to mitochon-

397 drial dysfunction and molecular damage (Murphy et al., 1989; Nguyen et al., 2011). KA injection

398 leads to high levels of ROS (Chuang et al., 2004; Gano et al., 2018), as seen in brain tissue from

399 TLE patients (Rowley & Patel, 2018).

400 Even though oxidative stress is often considered as the main mechanism of cell death in TLE,

401 many authors argue that apoptotic cell death contributes to the SE-induced brain damage or even 14

402 prevails. Fujikawa et al., 1993 summarized and explained all the pathophysiological mechanisms

403 that are associated with SE. Apparently, gene changes induced by SE can lead to apoptotic neu-

404 ronal death, and it is possibly tied to the excitotoxic damage component. Mitochondrial mem-

405 branes, being a subject of oxidative stress, activate apoptotic-inducing factor (AIF), which gets

406 translocated into the nuclei and initiates DNA fragmentation (Fujikawa et al., 2005). Thus, it can be

407 proposed that excessive glutamate release triggered by KAR activation, leads to excitotoxic cellu-

408 lar damage, which, in turn, activates apoptotic factors, and results in both apoptotic and excitotoxic

409 cell death. However, this aspect is still debated.

410

411 The mechanisms of KA-induced SE remain incompletely understood. However, in different species

412 and strains, KA-induced SE may result in different phenotypic traits. More generally, we observe

413 different phenotypes when we use different independent variables such as the mechanisms of SE

414 induction (e.g. pilocarpine, TBI, KA etc.), species, strain, gender, age etc. Intuitively, the state of

415 the brain circuits at the moment of the induction of SE will be central to the future outcome. Since

416 we cannot have access to the brain's state, it is not yet possible to explain why different pheno-

417 types are obtained. At present, we can only observe and describe them. A full understanding

418 would allow a knowledge-based development of epilepsy models with the desired phenotype.

419 Much work is still needed to reach this stage. At present, our personal opinion is that the mecha-

420 nisms of induction of epileptogenesis (how and where KA acts) are less important than the pheno-

421 typic traits that one wishes to study.

422

423 3. Administration routes

424 Kainic acid, like many other drugs, can be administered in various ways, depending on the desired

425 outcome (Fig. 1A-C, Fig. 2). Each method has its advantages and drawbacks and should be cho-

426 sen accordingly. The key parameters one should consider selecting the injection route are mortali-

427 ty rate, labor-intensity, lesion control, and, finally, age, sex, and strain of the animal. The major

428 administration routes are systemic, intracerebral (which could be divided into intraventricular, in-

429 trahippocampal, suprahippocampal, and intra-amygdaloid), and intranasal. Each of them is de-

430 scribed further. The striking feature that will immediately appear is the diversity in terms of fea- 15

431 tures, including the duration of the latent period, seizure properties and morpho-functional altera-

432 tions, even within a given model. This should not be seen as blurring the picture. Rather it unravels

433 the richness and diversity of all these models. The tables listing the different features as a function

434 of model, sex, strain etc. are provided to help researchers choosing the model best fitting their

435 questions.

436

437

438 4.1. Systemic administration

439

440 The first experiments using systemic KA administration were performed in the late 1970s to deter-

441 mine whether the toxin could induce damage via axonal connections (Schwob et al., 1980, Fig. 1A).

442 The studies showed that rats injected by 12 mg/kg of KA intraperitoneally experienced the onset of

443 "wet dog shaking" seizures approximately 30-90 min after injection, which eventually evolved into

444 secondary generalized tonic-clonic convulsions in 88% of cases. Neuronal damage was present

445 already at 3 h post-injection and gradually increased for two weeks. This method has been widely

446 used with different modifications. For instance, it was shown that KA, injected subcutaneously,

447 causes similar behavioral and network alterations as an intraperitoneal injection (Sperk et al., 1983)

448 and can be used as a model for TLE (Bertoglio et al., 2017). The significant advantage of systemic

449 KA administration is its low labor-intensity, which allows the injection of numerous animals in a

450 comparatively shorter time. Moreover, the absence of a surgical procedure eliminates side effects

451 created by anesthesia, surgery invasiveness, and extra damage made by direct contact with brain

452 tissue during the intracerebral injection. However, this model's obvious disadvantages are 1) no

453 control over the bioavailability of KA in the brain and 2) high mortality rates. As it was already men-

454 tioned before, Long-Evans and Wistar rat strains have higher mortality (Curia et al., 2008). In mice,

455 C57 and CH3 strains display an increased mortality rate, while 129/SvJ and SvEms strains have a

456 higher ratio of survival (McKhann et al., 2003). The summarized data for systemic KA administra-

457 tion in mice and rats is presented in Tables 2 and 3. It is important to note that kainic acid is quite

458 expensive (e.g., as compared to pilocarpine) and that the purity of the molecule can vary from one

459 stock to another (personal experience). There are also different forms of the drug, e.g., pure acid 16

460 and its dehydrate, which can sometimes be an issue in terms of result variability. Periods of kainic

461 acid shortage were also experienced in the past.

462

463 Status epilepticus

464 The sequence of events, followed by KA administration, includes the "initial insult," e.g., status epi-

465 lepticus, a latent seizure-free period, and the eventual development of chronic epilepsy. SE is

466 characterized as a condition resulting either from the failure of the mechanisms responsible for

467 seizure termination or the initiation of events, leading to abnormally, prolonged seizures (after time

468 point t 1) (El Houssaini et al., 2015). It is a condition, which can have long‐ term consequences (af-

469 ter time point t 2), including neuronal death, neuronal injury, and alteration of neuronal networks,

470 depending on the type and duration of the status (Trinka et al., 2015). In adult rats, a single dose of

471 KA can trigger SE characterized by a catatonic posture followed by facial myoclonus — stage 1

472 (according to Racine scale, Racine, 1972); masticatory automatisms, wet-dog shakes and head

473 nodding (stage 2); rearing with facial automatisms and forelimb clonus (stages 3-4); and finally re-

474 peated rearing and falling (stage 5) (Raedt et al., 2009). Following systemic KA administration, SE

475 develops within two h and lasts for approximately 5-9 h (if not stopped with an ad hoc injection of a

476 drug), according to multiple experiments on rats (Stafstrom et al., 1992; Giorgi et al., 2005; Lado et

477 al., 2006; Williams et al., 2009; White et al., 2010; Drexel et al., 2012; Bertoglio et al., 2017). Mortality

478 in this model is relatively high (around 22%), which can be reduced with slight alterations of the

479 protocol. Repeated injections of 2.5 mg/kg i/p (Hellier & Dudek, 2005) at 30 min intervals reduce

480 mortality down to approximately 5-6%, as compared to a single full-dose injection. Likewise, to in-

481 crease survival rate, SE can be decreased in severity after a few hours by subcutaneous injection

482 of 10 mg/kg diazepam or zolazepam (Sharma et al., 2008; Gurbanova et al., 2008). Many groups try

483 to stop SE after 20 min. As mentioned above, control of SE duration requires instrumenting the an-

484 imals for EEG recordings. We found a decreased mortality and accelerated recovery when timing

485 SE with electrophysiological recordings (as opposed to timing SE duration with behavior). We are

486 using the protocol to instrument animals with a telemetry device for supradural recordings at least

487 two weeks before KA injection. Two-three days before KA injection, we start EEG recordings in the

488 home cage, including following KA injection. 17

489

490 In mice, the clinical picture of systemically induced SE is similar to those seen in rats with slight

491 differences and nuances. For instance, mice do not always fit in the Racine scale; "wet-dog

492 shakes", while being common for rats, are never observed in mice (Sharma et al., 2018). The be-

493 havioral manifestation of SE occurs within 15-30 min post-injection (Hu et al., 1998; Benkovic et al.,

494 2004). The duration of SE does not exceed 6 hours (McKhann et al., 2003; Puttachary et al., 2015).

495 The mortality of systemic KA administration for mice is relatively significant (around 27%) and can

496 be similarly reduced with repeated injections of 5 mg/kg i/p (Tse et al., 2014). Anticonvulsant injec-

497 tions are likewise efficient in reducing the severity of SE (Araki et al., 2002).

498

499 One important parameter to consider when inducing SE is the time of the day. Human studies

500 show that SE occurs at specific times during the night and day cycle (Sánchez Fernández et al.,

501 2019, Goldenholz et al., 2018). Since the molecular architecture of the hippocampus (and other brain

502 regions) oscillates in a circadian manner, the sensitivity to SE may also depend upon the hour of

503 the day (Debski et al., 2020, Bernard, 2020). Performing injections spread over several hours in a

504 given day may lead to variable outcomes, not only in terms of survival but also in terms of pheno-

505 type.

506

507 The KA-induced SE correlates with epileptogenesis. Although it is intuitively straightforward to pro-

508 pose that SE triggers epileptogenesis, it is still possible that the same cause triggers SE and epi-

509 leptogenesis, i.e. SE does not cause epileptogenesis per se. There is some variability in terms of

510 SE severity and recovery between individual rodents, which may be due to basal biological differ-

511 ences (Manouze et al., 2019), including anxiety levels, social hierarchy, etc. This comment is also

512 valid for all other KA models described hereafter.

513

514 Latent period and spontaneous recurrent seizures

515 The latent period is the time between the initial brain insult (SE in the case of the KA rodent model

516 of TLE), and the development of chronic epilepsy manifested as the first SRSs. The latent phase is

517 considered one of the hallmarks of human epilepsy, and it can last for many years (French et al., 18

518 1993; Mathern et al., 1995). In some cases of epilepsy induced by TBI, stroke, or meningitis, the

519 latent period can be as short as one week (Löscher et al., 2015). The variability of the latent period

520 in humans constitutes a major puzzling question. It would be interesting to determine whether the

521 phenotype depends upon the duration of the latent period. Alternatively, since seizures are physio-

522 logical or built-in activities in any normal brain (Jirsa et al., 2014), the problem can be rephrased as

523 a probability one. The brain insult may trigger reorganizations that will evolve on slow and fast time

524 scales to progressively increase the probability for spontaneous seizures. The path taken to SRSs

525 may be traveled at different speeds, but it could be the same. This is difficult to model in rodents.

526 As previously mentioned, several studies report that some animals never developed SRSs after

527 SE. However, animals were never recorded 24/7 for months. It is possible that some animals de-

528 velop SRSs after a very long-lasting latent period. Such models would be particularly useful to ad-

529 dress the question of the variability of the latent period in humans.

530 In the rat model, the usual duration of the seizure-free period is approximately 5-30 days (White et

531 al., 2010; Chauvière et al., 2012; Drexel et al., 2012), although in some cases it may take up to 5

532 months (Williams et al., 2006). Experimental animals do not exhibit any behavioral seizures during

533 this phase, but electrophysiological recordings may show some epileptic activity, e.g., non-

534 convulsive seizures, which will be described next. Whether a "true" latent period exists is debated,

535 as the EEGs display evident anomalies in the few hours following SE, particularly interictal-like

536 spikes. In fact, the latent period's definition can be challenging as variability between recording

537 methods, disagreement on the definition of convulsive seizures, and even the differences in elec-

538 trode placement create certain obstacles. Some studies considered only behavioral seizures to be

539 the hallmark of SRSs and the end of the latent period, while others claimed that the presence of

540 electrographic seizures is the beginning of the chronic phase. Thus, the data on the latent period

541 differs greatly. In our hands, pathological activities, as manifested by spikes and bursts of spikes,

542 appear soon after SE induction and never stop, even after the first spontaneous seizure (Chauvière

543 et al., 2012).

544

545 The appearance of SRSs can be considered as the onset of the chronic phase of epilepsy. The

546 average daily frequency of seizures is 8 per day (±5.4) (Stafstrom et al., 1992; Giorgi et al., 2005; 19

547 Lado et al., 2006; Williams et al., 2009; White et al., 2010; Drexel et al., 2012; Bertoglio et al., 2017).

548 SRSs developing following KA injection in rats at different ages last for approximately 40 s. They

549 resemble stage 5 limbic seizures induced by electric kindling: bilateral forelimb clonus, masticatory

550 movements, rearing, and falling (Stafstrom et al., 1992). Subsequent studies, using continuous

551 EEG/video-recording, demonstrated similar results, with the seizure ratio varying from 4 to 21 per

552 day (Lado et al., 2006; Williams et al., 2009). The F344 rat strain has been demonstrated to be sen-

553 sitive to KA with a much lower dose (10.5 mg/kg) needed to induce SRSs (Rao et al., 2006). Promi-

554 nent neurodegeneration in the hippocampus, extensive mossy fiber sprouting, and the consistent

555 SRS ratio, with the same frequency over time, makes this strain interesting to obtain a stable and

556 reproducible model.

557

558 In mice, the latent phase usually is significantly shorter than in rats and could be estimated as 2-3

559 days (Puttachary et al., 2015). Unlike the rat model, the development of SRSs is more difficult to

560 obtain in mice following a systemic administration, and those detected are infrequent and highly

561 variable (McKhann et al., 2003; Umpierre et al., 2017).

562

563 It is important to note that the high rate of SRSs reported in previous studies may be because ani-

564 mals are singly housed, which increases seizure frequency by a factor of 16, at least in the pilo-

565 carpine model (Manouze et al., 2019; Bernard, 2019). Single housing adds stress as an independent

566 variable and a confounding factor to the epilepsy phenotype. We recommend maintaining social

567 interaction (Manouze et al., 2019; Bernard, 2019). This comment applies to all models.

568

569 Take home message #4: it would be particularly informative to perform a classification of patholog-

570 ical events in each model, particularly spikes and seizures, using unsupervised analysis. As spikes

571 and seizures may evolve in time, it is important to describe their dynamics during epileptogenesis

572 and later. This comment is also valid for all other KA models described hereafter.

573

574 Histological evaluation

20

575 Systemic administration of KA induces extensive bilateral neuronal damage throughout the brain.

576 The first noticeable changes appear within 48 hours post-treatment and are present primarily in

577 CA1, CA3, and CA4 hippocampal subregions (Sperk et al., 1983; Sommer et al., 2000; Haas et al.,

578 2001; Drexel et al., 2012.) Subsequently, the entire hippocampus is affected in the rat brain (Kar et

579 al., 1997; Suárez et al., 2012). The typical pattern of KA-induced damage includes necrosis of py-

580 ramidal cells, gliosis, and mossy fiber sprouting within the dentate gyrus's inner molecular layer

581 (Tauck & Nadler, 1985). Additionally, there have been multiple reports indicating damage of various

582 extrahippocampal areas, such as the amygdala (Fritsch et al., 2009), subiculum, entorhinal cortex,

583 thalamus, caudoputamen (Drexel et al., 2012), substantia nigra, hypothalamus (Covolan & Mello,

584 2000), olfactory bulb, anterior olfactory nucleus (Altar & Boudry, 1990) and other areas.

585

586 T2-weighted MRI images of Wistar and Sprague-Dawley rat strains, systemically treated with KA,

587 showed an interesting phenomenon, contradicting previous mortality findings. The extent of neu-

588 ronal damage is higher in the Sprague-Dawley strain, while in Wistar rats, the relative volume of

589 the hippocampus is not different from the control animals (Inostroza et al., 2011). Confirming the

590 MRI data, post-mortem NeuN-staining revealed a significant pyramidal loss in CA1 and CA3 areas

591 of the hippocampus in Sprague-Dawley rats but not in Wistar (Table 1).

592

593 In mice, the predominantly affected areas are CA3 and CA1 (Hu et al., 1998, McKhann et al., 2003,

594 Kasugai et al.; 2007, Mouri et al., 2008). Similarly to rats, damage induced by systemic KA admin-

595 istration includes pyramidal cell loss, mossy fiber sprouting, and reactive gliosis (McKhann et al.,

596 2003, Benkovic et al., 2004). There is also evidence of extra-temporal damage, particularly in the

597 cortical areas, lateral amygdala, and dorsal thalamus (Hu et al., 1998, McKhann et al., 2003, Ben-

598 kovic et al., 2004, Umpierre et al., 2017).

599

600 Take home message #5: As noted above, a better understanding of all morphological changes in

601 different brain regions is needed to obtain a global picture of the model. This comment is also valid

602 for all other KA models described hereafter.

603 21

604 4.2. Intraventricular injection

605

606 Nadler et al., 1978 have demonstrated that the intracerebroventricular (icv) injection of 0.5 nmol KA

607 into the rat brain leads to CA3 neurodegeneration within 1-3 days, whereas higher doses (0.8mg

608 and more) cause damage to both CA1 and CA2. Similar findings were later observed by Lancaster

609 & Wheal, 1982, who confirmed the same neuronal damage pattern and helped establish the icv

610 paradigm, which is still used (Luo et al., 2013; Song et al., 2016). The data of icv KA administration in

611 rats is presented in Table 2.

612

613 Status epilepticus

614 The induction of SE using the icv route is consistent with other methods. It starts within 10-30 min

615 post-injection and manifests as bradypnea, circling behavior, and wet-dog shakes and subsequent-

616 ly progresses into continuous motor seizures (Miyamoto et al., 1997). The duration of SE varies be-

617 tween 2 and 6 h, which can be stopped by a diazepam injection to reduce the mortality rate

618 (Miyamoto et al., 1997; Jing et al., 2009; Song et al., 2016). Typically, the mortality rate does not ex-

619 ceed 10% and is considered relatively low compared to the systemic route of administration

620 (Miyamoto et al., 1997; Gao et al., 2019).

621

622 Latent period and spontaneous recurrent seizures

623 There are multiple reports regarding SRS development following icv KA administration. Having

624 performed a 24/7 continuous video/EEG-recording, few authors reported similar findings, with the

625 latent period duration varying between 1 and 2 weeks and SRSs appearing consistently, several

626 times per day (Song et al., 2016, Gao et al., 2019).

627

628 Histological evaluation

629 Morphological damage induced by an icv KA administration might seem paradoxical; even though

630 the toxin is injected in the ventricular system and thus potentially distributes throughout the brain,

631 the actual lesion appears to be restricted to CA3/C4 hippocampal subfields ipsilaterally to the injec-

632 tion site (Nadler et al., 1978; Miyamoto et al., 1997). The contralateral hippocampus and the ex- 22

633 trahippocampal areas are relatively spared, and in some cases, almost entirely intact (Song et al.,

634 2016). Mossy fiber sprouting is also present in most cases (Jing et al., 2009; Song et al., 2016; Gao et

635 al., 2019). The pathology is mostly unilateral, with occasional signs of damage on the contralateral

636 side (Lancaster & Wheal, 1982; Miyamoto et al., 1997). The model is closer to what is found in pa-

637 tients with MTLE, as the epileptogenic network is usually limited to one hemisphere (Table 1).

638

639

640 4.3. Intrahippocampal and intra-amygdaloid injections

641

642 The first experiment with intra-amygdaloid KA administration in 1978 demonstrated characteristic

643 behavioral, electrophysiological, and histological hallmarks in baboons (Ben-Ari & Lagowska, 1978;

644 Ben-Ari et al., 1979). Other studies used intrahippocampal KA injections in rats to observe the

645 same outcome (Fonnum & Walaas, 1978). Both methods of intracerebral administration are still

646 widely used. The principal advantage is a direct delivery into the brain tissue, bypassing the blood-

647 brain barrier. The idea is to produce focal damage. The mortality rate appears to be lower than

648 systemic administration (Sharma et al., 2007). Moreover, the standardized protocol for the intrahip-

649 pocampal KA administration, combined with the electrode implantation, which has been proposed

650 recently (Bielefeld et al., 2017), enables a good reproducibility and allows the combination of both

651 electrophysiological and behavioral approaches. A brief summary of both methods is presented in

652 Tables 2 and 3.

653

654 Status epilepticus

655 SE, following an intracerebral KA injection, develops rapidly. Experiments showed that in rats,

656 which underwent intrahippocampal KA administration in the dose of 0.4-2.0 μg, SE emerges within

657 5-30 min after the injection (Bragin et al., 1999; Raedt et al., 2009). Similar effects were observed

658 after the intra-amygdaloid injection of 0.75 μg of KA (Gurbanova et al., 2008). SE, elicited by in-

659 trahippocampal administration appears to be longer than systemic administration and can last for

660 more than 17 h (Rattka et al., 2013). Intra-amygdaloid KA administration in rats is efficient in trigger-

23

661 ing SE lasting 4-6 h on average (Ben-Ari et al., 1979; Tanaka et al., 1992). Generally, intracerebral

662 administration's lethality does not exceed 10% (Rattka et al., 2013).

663

664 Mice demonstrate the same pattern of SE starting within 30 min after intrahippocampal (Lee et al.,

665 2012) or within 20 min after intra-amygdaloid (Mouri et al., 2008) administration. The duration is ap-

666 proximately the same as for rats and can last for up to 12 h (Gröticke et al., 2008). However, there

667 is a common approach to termination seizures via an anticonvulsant injection after 40 min of sei-

668 zure activity, thus reducing mortality rate to a minimum (Shinoda et al., 2004). FVB/N mouse

669 strain exhibits prolonged seizure activity as compared to C57BL/6 strain after intra-amygdaloid KA

670 administration, although the overall SE pattern is similar in both strains (Kasugai et al., 2007). The

671 mortality rate is relatively low compared to the systemic mode of drug administration (Bouilleret et

672 al., 1999; Mouri et al., 2008).

673

674 Latent period and spontaneous recurrent seizures

675 One of the first successful experiments on intrahippocampal KA administration was performed by

676 Cavalheiro et al., 1982, when he and his colleagues conducted a series of tests in rats, using sev-

677 eral different doses of KA in the range of 0.1-3.0 μg, thus creating three distinct cohorts: animals

678 which were injected with either 0.1-0.4 μg or 0.8-2.0 μg experienced SE, but only the latter group

679 developed SRSs approximately 5-21 days after treatment. The rats injected with 3.0 μg of KA died

680 due to severe SE. The SRSs resembled those happening during the acute phase: lasting for 40-60

681 seconds, involving salivation, masticatory movements, bilateral forelimb clonus, rearing, and fall-

682 ing. The chronic period continued for 22-46 days, after which behavioral seizures ceased, and an-

683 imals entered a post-seizure period. Generally, the occurrence of SRSs after the intrahippocampal

684 administration of KA is consistent in rats and resembles the pattern of systemic injection (Rattka et

685 al., 2013).

686

687 Intra-amygdaloid administration is efficient for inducing chronic epilepsy in rats. The average

688 amount of intrahippocampal KA-induced SRSs is 12 per day (Gurbanova et al., 2008). Latency to

24

689 the first SRS is reported to be 1-4 weeks (Cavalheiro et al., 1982, Klee et al., 2017) in both intrahip-

690 pocampal and intra-amygdaloid administration.

691

692 The first experiment on intrahippocampal KA injection in mice was performed in 1999, and SRSs

693 were observed with the frequency of 1-2 seizures per week (Bouilleret et al., 1999). Unlike rats, the

694 EEG results in mice are highly variable, with some authors reporting the frequency of electrograph-

695 ic SRSs to reach dozens per hour (Riban et al., 2002, Gouder et al., 2003), which could be due to

696 social isolation-induced stress (Manouze et al., 2019; Bernard, 2019). According to a recent study,

697 this phenomenon of frequent electrographic seizures is specific to mice but not to rats (Klee et al.,

698 2017).

699 It is important to highlight the study of Sheybani and coll (Sheybani et al., 2018). To the best of our

700 knowledge, with the study performed in the pilocarpine model (Toyoda et al., 2015), this is the only

701 other study in which multisite recordings have been performed to assess an experimental model of

702 TLE at the large-scale network level. The study clearly shows that epileptiform activity progressive-

703 ly expands beyond and independently from the seizure focus. This result is particularly important

704 as it clearly shows that time-dependent alterations in neuronal circuits need to be considered when

705 studying epilepsy in chronic experimental models, including during epileptogenesis (Słowiński et

706 al., 2019).

707

708 Following the intra-amygdaloid KA administration, SRSs developing in mice have a frequency of 1-

709 5 per day (Mouri et al., 2008; Tanaka et al., 2010; Li et al., 2012; Liu et al., 2013). The latent period

710 does not exceed two weeks for both types of intracerebral administration (Lee et al., 2012; Mouri et

711 al., 2008).

712

713 An interesting phenomenon was noticed following an intrahippocampal KA administration in mice.

714 The dentate gyrus is the major zone of adult neurogenesis (Hochgerner et al., 2018). Normally, sys-

715 temic administration of KA induces bilateral neurogenesis in the hippocampus, which, combined

716 with the data collected from human patients, led to an assumption that increased neurogenesis

717 might contribute to epileptogenesis in TLE (Parent, 2007). However, findings in the intrahippocam- 25

718 pal KA mouse model of TLE suggest the opposite. Fritschy et al., 2005 demonstrated that KA, uni-

719 laterally injected into the hippocampus, causes a massive reduction of the neurogenesis on the

720 injection site and a compensatory neurogenic activity in the contralateral dentate gyrus. It could be

721 speculated that, according to this data, neurogenesis does not contribute to the genesis of the epi-

722 leptogenic zone and that, perhaps, it gets disrupted by changes induced in the hippocampus, e.g.

723 granule cells dispersion. This topic is interesting to pursue. A useful review on this topic was pub-

724 lished in 2012, and it partly covers the KA model (Parent & Kron, 2012).

725

726 Histological evaluation

727 Histological studies reveal similarities between the two intracerebral KA administration routes,

728 which are fairly similar in rats and mice (Table 1). The intrahippocampal injection creates a focal

729 lesion in the hippocampus, damaging predominantly the ipsilateral CA3 subfield and, to some ex-

730 tent, CA4 and CA1, regardless of the site of injection (French et al., 1982; Bouilleret et al., 1999;

731 Raedt et al., 2009; Twele et al., 2015). Interestingly, intra-amygdaloid administration induces massive

732 neuron destruction in the same hippocampal regions, along with a lesion in the amygdala (Tanaka

733 et al., 1992; Araki et al., 2002). Some authors reported extensive damage on the contralateral side

734 (Ben-Ari et al., 1980; Araki et al., 2002; Mouri et al., 2008) and the resistance of the CA2 hippocam-

735 pal subfield to the KA-induced toxicity (Shinoda et al., 2004). Mossy fiber sprouting was also ob-

736 served in both rats and mice (Raedt et al., 2009; Zeidler et al., 2018; Gurbanova et al., 2008; Mouri et

737 al., 2008). Granule cell dispersion in the dentate gyrus, however, is only reported with the intrahip-

738 pocampal administration (Rattka et al., 2013; Bielefeld et al., 2017). Additionally, there is multiple

739 evidence of damage in extra-temporal regions induced by intra-amygdaloid injection, such as thal-

740 amus (Ben-Ari et al., 1980), entorhinal, and perirhinal cortices (Mouri et al., 2008; Tanaka et al.,

741 2010). There is a significant strain difference in mice. The C57BL/6 mouse strain is resistant to cel-

742 lular damage induced by KA (Schauwecker & Steward, 1997; Kasugai et al., 2007). In contrast, the

743 FVB/N strain shows increased sensitivity and extensive neurodegeneration in the CA1 and CA3

744 areas of the hippocampus following intra-amygdaloid KA administration (Kasugai et al., 2007). A

745 representative image of the FluoroJade B (FjB) staining of murine hippocampus and cortex after

746 intra-amygdaloid KA injection is presented in Figure 3. It has to be stated that the figure does not 26

747 represent all the possible variations in histological findings but only a particular case. The variabil-

748 ity of KA-induced damage is high, and it should be assessed with a stereological method, if possi-

749 ble. Since it is difficult to summarize the data pool regarding histological studies, we chose only

750 one image.

751

752

753 4.4. Suprahippocampal

754

755 One of the variations of the intrahippocampal administration route is the so-called suprahippocam-

756 pal injection, where KA is injected in the cortical area just above the hippocampus. This technique

757 was introduced by Bedner et al., 2015, who performed an intracortical injection to reproduce the fea-

758 tures of the intrahippocampal administration and avoid damaging the CA1 subfield of the hippo-

759 campus. All animals experienced SE, triggering epileptogenesis. Subsequently, other researchers

760 tested this method several times, proving to be efficient in mice (Table 3).

761

762 Status epilepticus

763 Starting soon after recovery from anesthesia, SE induced by the intracortical KA application lasts

764 for up to 12 hours and consists of forelimb clonus and repetitive rearing and falling (Jefferys et al.,

765 2016). The reported mortality is relatively low and ranges from 0% (Pitsch et al., 2019) to 9.7%. As

766 mentioned by the authors, death occurred mostly due to surgical complications and not seizures

767 (Bedner et al., 2015).

768

769 Latent period and spontaneous recurrent seizures

770 SRSs in this model appear after a brief latent period of 4-13 days and are consistent and frequent.

771 Bedner et al., 2015 reported a frequency of 7.5 ± 6.2 seizures per day 7-9 months post-SE and pro-

772 gressive nature of chronic epilepsy development, consistent with previous data on KA-induced sei-

773 zure progression (Williams et al., 2009). Another study demonstrated success inducing SRSs in

774 100% of treated animals with the seizure frequency of 1-1.5 per day (Pitsch et al., 2019).

775 27

776 Histological evaluation

777 The lesion created by suprahippocampal injection of KA is unilateral and predominantly situated in

778 the CA1/CA3 hippocampal subfields. Three months post-treatment, there is a complete neuronal

779 loss in CA1, significant damage in CA3, and massive dentate granule cell dispersion (Bedner et al.,

780 2015, Jefferys et al., 2016, Pitsch et al., 2019). Mossy fiber sprouting is not characteristic of this mod-

781 el (Table 1).

782 It would be particularly interesting to determine whether the epileptogenic network develops as in

783 the intrahippocampal model (Sheybani et al., 2018).

784

785

786 4.5. Intranasal injection

787

788 The idea of delivering KA via nasal epithelium absorption first emerged as an attempt to trigger SE

789 in the most common strain of transgenic laboratory mice, C57BL/6, known for its resistance KA-

790 induced neuronal death (Schauwecker & Steward, 1997). The first experiments showed that KA,

791 dissolved in water and delivered intranasally in the dose of 40-60 mg/kg, causes focal seizures,

792 which then generalize into SE, in 100% of treated C57BL/6 mice (Chen et al., 2002). Behavioral and

793 anatomical findings were consistent with the previous results, obtained via standard drug delivery

794 protocols. Likewise, another study demonstrated that a dose of 30 mg/kg KA intranasally is asso-

795 ciated with low mortality and >90% success in developing SE (Sabilallah et al., 2016). Even though

796 the mechanism is not fully understood, it is assumed that the drug is absorbed by olfactory epithe-

797 lium and reaches the hippocampus (as well as other brain areas) via olfactory pathways (Frey et

798 al., 1997, Fig.1C), as the olfactory bulb has widespread connections with different regions of the

799 brain (Chen et al., 2002). Although this method has not been studied thoroughly and so far has not

800 found widespread success in the neuroscience community, it has its advantages, such as repro-

801 ducibility, low labor-intensity and low mortality (Table 3).

802 Status epilepticus

803 The induction of SE appears to be consistent in all existing studies; it starts within 15-30 min after

804 administration (Chen et al., 2002; Duan et al., 2006) with the first symptoms being immobility and 28

805 staring, followed by generalized tonic-clonic seizures, lasting for 1-5 h (Chen et al., 2002). The mor-

806 tality rate is relatively low and can be compared to intrahippocampal routes, ranging from 0% to

807 12% (Chen et al., 2002; Duan et al., 2006; Zhang et al., 2008; Lu et al., 2008; Sabilallah et al., 2016).

808 Generally, there are no specific SE features associated with the intranasal administration route

809 compared to the conventional methods.

810

811 Latent period and spontaneous recurrent seizures

812 Seizure progression was reported by at least one research group; Sabilallah et al., 2016 reported

813 latency of 15-30 days before electrographic SRSs are observed in most animals in the form of

814 spontaneous spike activity and occasional seizures. However, the study lacked continuous 24/7

815 video/EEG-recording, so the actual outcome might slightly differ. Other authors reported increased

816 locomotor activity (Chen et al., 2002; Zhang et al., 2008; Lu et al., 2008), perhaps in relationship with

817 the development of chronic epilepsy.

818

819 Histological evaluation

820 Morphological damage induced by intranasal KA administration has been studied. The brain re-

821 mains mostly intact, except for the hippocampus and the olfactory bulbs (Chen et al., 2002; Zhang

822 et al., 2008). The CA3 area of the hippocampus shows the most prominent damage (Duan et al.,

823 2006; Zhang et al., 2008). Mossy fiber sprouting and granule cell dispersion have not been reported.

824 There is evidence of massive astrogliosis and microglial response in the hippocampus (Chen et al.,

825 2002; Duan et al., 2006; Zhang et al., 2008; Lu et al., 2008; Sabilallah et al., 2016). Overall, there is

826 evidence for characteristic pathological features of “classic” KA-induced epilepsy. However, the

827 damage seems to be localized and restricted to the CA3 area of the hippocampus (Table 1).

828

829

830 4. Electrophysiology

831

832 SE elicited by KA administration is characterized by typically isolated spikes, polyspikes, spike and

833 wave complexes (Stafstrom et al., 1992; Riban et al., 2002; Jagirdar et al., 2015). With the advantage 29

834 of simultaneous EEG and video recording, it is possible to distinguish electrographic SE (ESE)

835 from the convulsive SE (CSE), since numerous studies reported ESE to last much longer than

836 CSE (Fritsch et al., 2010; Drexel et al., 2012; Sabilallah et al., 2016; Umpierre et al., 2016). Consider-

837 ing 20 min of ESE rather CSE results in better survival and more rapid recovery. However, this re-

838 quires instrumenting the animals. It is important to remember that any invasive procedure (like im-

839 planting electrodes/transmitters under anesthesia) may change an unknown number of biological

840 parameters in each experimental animal. In other words, we do not know whether instrumenting

841 animals produces a different model/phenotype as compared to non-instrumented animals.

842

843 In our hands, after a drug injection to stop SE, we observe the occurrence of spikes, which fre-

844 quency decreases during the first hours after SE, while the EEG slowly returns to the pre-SE level

845 (between spikes). Then, the spikes start to organize themselves in bursts. During this phase, we

846 see epileptiform discharges, which look like very short seizures (2 sec long). A few days after SE,

847 SRSs start to occur with a typical >10 s duration. In the pilocarpine model, two types of interictal

848 spikes can be distinguished during the latent period: type 1 is a spike followed by a long-lasting

849 wave, and type 2 is a spike without a wave (Chauvière et al., 2012). The authors suggest that type 1

850 spikes correspond to the activity of both excitatory and inhibitory neurons, while type 2 spikes re-

851 flect the activity of a small pool of excitatory cells. While the number, amplitude, and duration of

852 type 1 spikes decreases while type 2 spikes become more frequent before the first spontaneous

853 seizure. A similar study should be performed in the various KA models.

854

855 Spikes can evolve into hippocampal paroxysmal discharges (HPDs) (Riban et al., 2002; Jagirdar et

856 al., 2015), which are essentially short periods of epileptiform activity in the absence of any behav-

857 ioral symptoms (Pallud et al., 2011). Latency for electrographic seizures is shorter than for convul-

858 sive seizures (Lado, 2006, Williams et al., 2009). During this time, occasional interictal spikes and

859 HPDs can be observed (Riban et al., 2002; White et al., 2010).

860 The detection of electrographic SRSs precedes motor symptoms observation and does not always

861 correlate to it (Bragin et al., 2005; Li et al., 2012; Tse et al., 2014).

862 30

863 Various electrographic patterns were reported in KA studies. For example, Riban et al., 2002 classi-

864 fied seizures emerging during the chronic period as low voltage spikes (600–900 μV, 100–150 ms),

865 which persisted for 2 weeks and were never observed again, bursts of high-frequency, low voltage

866 spikes (300–500 μV, 18–26 Hz), appearing just for 1-2 days, high voltage sharp waves (1500–

867 4500 μV, 150–200 ms), persisting until termination of the experiment, and HPDs, which, being the

868 hallmark of the latent period, were still present during the chronic phase. Other authors (Bragin et

869 al., 2005; Lévesque et al., 2012) reported two distinct seizure onset patterns for SRSs: hypersyn-

870 chronous (HYP) and low-voltage fast (LVF) onset patterns, which are also found in patients with

871 epilepsy (Velasco et al., 2000, Ogren et al., 2009). HYP seizures, which represent 50% to 70% of

872 SRSs, are essentially multiple periodic spikes with a frequency of approximately 2 Hz, restricted to

873 a small portion in the hippocampus. LVF onset pattern consists of a single spike followed by high-

874 frequency activity (> 25Hz) originating from hippocampal or extrahippocampal networks. The typi-

875 cal electrographic recordings of SRSs are presented in Figure 4. A taxonomy of 16 types of sei-

876 zures has also been proposed and validated in patients (Jirsa et al., 2014; Saggio et al., 2020) and in

877 the tetanus toxin model of epilepsy (Crisp et al., 2020), in which individual animals switch between

878 different types of seizures during epilepsy. Future work is needed in the KA models to determine

879 the type of seizures they express and how they evolve. This requires Direct Current recordings, as

880 opposed to the most commonly used Alternative Current recordings.

881

882 The arguments developed previously clearly show complex dynamical phenomena that occur dur-

883 ing the latent period. As soon as SRSs start to occur, it is essential to remember that a steady

884 state is never reached. Seizures tend to progress over time (Table 2). Williams et al., 2009 pro-

885 posed that the evolution of seizures consists of 4 stages, where stage 1 represents a latent period,

886 stage 2 is characterized by a slow increase in seizure frequency, e.g., “slow growth phase," stage

887 3 is marked as the “exponential growth phase” and stage 4 is a final steady-state plateau phase,

888 which, however, was not observed in all the animals (Fig.5). This result agrees with other findings,

889 most of which described a progressive increase of SRSs during the chronic period of KA-induced

890 epilepsy (Gorter et al., 2001; White et al., 2010; Rattka et al., 2013; Van Nieuwenhuyse et al., 2015).

891 Thus, the evolution of seizures can perhaps be best represented by a sigmoid Boltzmann function, 31

892 showing the exponential growth phase. However, recent works depict a more complex picture as

893 circadian and multidien rhythms need to be considered (Bernard, 2020).

894

895 Human studies clearly show that seizure activity follows various rhythms: circadian, ultradian, mul-

896 tidien or multi-day, seasonal, etc. (Quigg, 2000; Manfredini et al., 2004; Spencer et al., 2016; Karoly

897 et al., 2018; Baud et al., 2018). The circadian rhythmicity has been recognized for millennia in hu-

898 mans, but multidien rhythms have been identified recently in patients (Baud et al., 2018). The au-

899 thors found that interictal activity, in addition to its circadian rhythmicity, has a slow envelope of

900 several days period, which is patient-specific. Interestingly, seizures tend to occur during the rising

901 phase of this low rhythm. Besides, clusters of seizures are also found during the rising phase.

902 Similar features have been found in the KA model with a robust periodicity of 5-7 days, each ani-

903 mal following its own cycle (Baud et al., 2019). Although rendering data analysis and interpretation

904 more complex, it is important to take circadian and multidien cycles into account, particularly when

905 investigating the molecular mechanisms underpinning seizure genesis (Debski et al., 2020). The

906 fact that the molecular architecture of the hippocampus oscillates in a circadian manner (Debski et

907 al., 2020), provides an entry point to understand the circadian rhythmicity of seizures (Bernard,

908 2020). It is important to stress again that KA's sensitivity (e.g. the threshold to induce SE) may

909 change during the night and day cycle (Debski et al., 2020).

910

911 Finally, as mentioned above, the way animals are housed appears as a key determinant of SRS

912 frequency. When animals are singly housed, they are highly stressed, and their seizure frequency

913 is 16 times greater than in animals maintaining social interaction (Manouze et al., 2019; Bernard,

914 2019). Solutions have been developed to maintain social interaction when animals are instrument-

915 ed (Manouze et al., 2019; Bernard, 2019).

916

917 Take home message #7. Studying seizure frequency requires several weeks of 24/7 continuous

918 recordings to account for circadian and multidien rhythms. Large discrepancies found in animals

919 recorded during one week may stem from the fact that these animals were going through a high or

920 low SRS frequency part of their multidien cycle. 32

921

922 5. Age, sex and strain specificity

923

924 The outcome of KA administration depends not only on the used species and strains but also on

925 the sex and age. This diversity is an advantage as it may allow us to reproduce the diversity of

926 phenotypes found in TLE patients. For instance, C57BL/6 mice, the most widely used transgenic

927 strain, are resistant to KA-induced damage, which can be, however, compromised by intranasal

928 drug administration. Other strains with known resistance are BALB/c (Schauwecker & Steward,

929 1997), C3HeB/FeJ, 129/SvEms, 129/SvJ (McKhann et al., 2003), ICR, and FVB/N (McLin et al.,

930 2006). It has also been proposed to differentiate between behavioral resistance (129/SvEms), lack

931 of anatomical alterations (C57, C3H), and combined resistance (129/SvJ) (McKhann et al., 2003).

932 In contrast, other mouse strains have increased sensitivity to KA, such as C57BL/10, DBA/2J, and

933 F1 C57BL/6*CBA/J (McLin et al., 2006). Also, DBA/2J and FVB/N mouse strains exhibit higher sei-

934 zure-induced mortality as compared to C57BL/6J (Ferraro et al., 1995; Royle et al., 1999). Overall,

935 the difference between rodent strains is crucial and should always be taken into account. Investi-

936 gating different outcomes in different strains could also help to understand the model better. In

937 fact, not all researchers agree that KA, injected into a mouse brain, leads to chronic epilepsy de-

938 velopment. Considering failures in inducing seizures followed systemic KA administration, incon-

939 sistent histological data, and great variability between strains and even within one strain, it is nec-

940 essary to point out that any result should be analyzed with care. However, current advances in sci-

941 ence allow us to investigate the issue in more detail. Perhaps, over the next few years, there will

942 be an agreement on whether mice indeed develop KA-induced epilepsy and how to interpret any

943 related strain differences.

944

945 It is important to note that we have found a large variability between animal vendors in Europe for

946 a given strain because the genetic background, the housing conditions etc. may differ. For exam-

947 ple, we experienced a major issue with several batches of animals from a given vendor. Our proto-

948 col was not working anymore. After probing, the vendor told us that they had changed the feeding

949 conditions to accelerate weight gain. Since some injections are per kg, we were injecting a "much 33

950 younger" animal. Besides, whatever they added to the food may have changed the biology of ani-

951 mals. This is also a source of variability and discrepancy between laboratories.

952

953 Another important factor one should consider when working with a rodent model of TLE induced by

954 KA administration is the animals' age. The first indications of different seizure susceptibility in

955 younger and older rats appeared in the late 1980s when it was demonstrated that an i/p injection of

956 KA in P12 rat pups causes more severe SE as compared to P27 adult rats (Holmes & Thompson,

957 1988). Subsequently, other authors reported similar results: shorter latency before developing SE

958 in younger (P5-10) rats compared to P20-60 (Stafstrom et al., 1992), and higher mortality in P15

959 rats as compared to P53 after an i/p KA injection (Mikati et al., 2003).

960

961 Aged rodents are more prone to seizure activity than adults (Wozniak et al., 1991, McCord et al.,

962 2008), and that this tendency does not depend on a strain (McCord et al., 2008). Aged rats have a

963 shorter latency to SE onset and exacerbation of pre-seizure behavioral manifestations than middle-

964 aged animals (Darbin et al., 2004). The reasons behind this are debated and include multiple theo-

965 ries including age-related changes in synaptic connectivity (Rapp et al., 1999), electrotonic coupling

966 (Barnes et al., 1997), number and type of neurons (Stanley & Shetty, 2004) and diminished density of

967 glutamate receptors in the aged brain (Wenk & Barnes, 2000; Lerma et al., 2001). Overall, this data

968 seems to describe a parabolic pattern of KA-induced seizure susceptibility, where rodents of young

969 and old age are more vulnerable than middle-aged animals. This correlates well with findings in

970 human epilepsy, where age-specific incidence peaks in childhood and then, after a decline in mid-

971 dle age, raises again at the age of 60 years and older (Kotsopoulos et al., 2004).

972

973 An important point regarding age differences was already mentioned before and includes multiple

974 models of TLE in immature rodents (Holtzman et al., 1981, Jensen et al., 1991, Dubé et al., 2010,

975 Rakhade et al., 2011). Given the high prevalence of epilepsy during development and the fact that a

976 developing brain is not a small adult brain, there is a dire need for specific developmental models

977 of epilepsy, as information gained in adults does not translate automatically to the developing

34

978 brain. Developmental epilepsy is a field in itself, and we refer the reader to the relevant literature,

979 as we mostly focused here on adult TLE.

980

981 It is remarkable to notice that the vast majority of epilepsy studies have been and are being per-

982 formed in male rodents. We do not know whether the wealth of data obtained in males applies to

983 females. There is a dire need for research on female subjects of various strains. The currently

984 available data demonstrates some clear differences between male and female rodents regarding

985 the phenotype, epileptogenic processes in the brain, and anatomical alterations. Limiting research

986 to male animals creates a possible misunderstanding of the model and epilepsy in general. This

987 key issue remains to be addressed. Differences are expected, if only when considering structural

988 arguments. The rodent hippocampus, particularly the CA3 area, contains various receptors for

989 gonadal steroids. For instance, the concentration of androgen receptors in stratum lucidum of the

990 CA3 subfield appears to be higher than in any other hippocampal region (Tabori et al., 2005), and

991 mossy fibers express ERα and ERβ estrogen receptors (Torres-Reveron et al., 2009). Several stud-

992 ies have shown that mossy fiber pathway stimulation evokes different levels of BDNF protein ex-

993 pression response in males, females in various stages of oestrus cycle and ovariectomized female

994 rats (Skucas et al., 2013; Scharfman et al., 2003, Scharfman et al., 2007). Only females in proestrus

995 and estrous stages of menstrual cycles exhibit epileptiform activity after a 10 s 1 Hz train of paired

996 pulses and exhibit a strong BDNF expression (Scharfman et al., 2003; Scharfman et al., 2007).

997 Furthermore, in C57BL/6 mice, KA injected at 20mg/kg or 30mg/kg causes SE in ~100% of aged

998 female mice compared to the young females and males of both groups and the highest level of

999 BDNF expression (Zhang et al., 2008). Altogether, this data provides a valuable insight into the in-

1000 fluence of sex on KA administration and susceptibility to seizures, showing that the oestrus cycle

1001 should be considered while modeling epilepsy in female mice. However, a tremendous amount of

1002 research effort is required to study epilepsy in experimental female TLE models.

1003

1004

1005 6. Concluding remarks

35

1006 In this review, we have described the different versions of the kainic acid model. This model has

1007 been extensively used for decades and is proven to be a reliable tool to mimic numerous behav-

1008 ioral, electrophysiological, and anatomical features of epilepsy. In our opinion, whether or not the

1009 KA models are mimicking human TLE is not a relevant question if only because a rodent brain is

1010 not a human brain. What is important are the questions addressed in the rodent models and the

1011 hypotheses being tested. We argue that all models are interesting as long as they are character-

1012 ized by spontaneous seizures.

1013

1014 Diversity is a hallmark of human epilepsy, even for some specific types such as TLE. Interestingly,

1015 different kainic acid models are also very diverse. The results vary as a function of how kainic acid

1016 is administered, species, strain, sex, and age. We argue that this is a strength as the different

1017 models may cover some of the patients' diversity.

1018 Because diversity is the hallmark of human epilepsy, clinicians consider patients as individuals. We

1019 argue that a similar approach should be used with rodents. We have started to consider each rat or

1020 mouse as an individual (Manouze et al., 2019). Given the existing individual variability, averaging

1021 values for a given observable may blur the reality. Although more time-consuming, we argue that

1022 an individual approach may allow the extraction of categories, which may bear translational values

1023 (e.g., responders vs. non-responders to a treatment).

1024

1025 Each model only revealed the tip of the iceberg. Many important issues remain unaddressed. Per-

1026 haps the most important one is determining the seizure onset zone(s) and the propagation pattern

1027 in the brain. This requires multi-site recordings, which are regularly done in patients during presur-

1028 gical evaluation. Such a study has been performed in the pilocarpine model (Toyoda et al., 2015)

1029 and the intrahippocampal kainate model (Sheybani et al., 2018). The same approach should now be

1030 used in the various KA models. More generally, we argue that the phenotype of each animal model

1031 is far from being completely described. Considerable work is needed regarding the type of seizures

1032 expressed, their time evolution, their site of origin, their zone of propagation, the morpho-functional

1033 reorganizations in different brain regions, and comorbidities (depression, anxiety, cognitive deficits,

1034 etc.). Another pressing question is sex. As mentioned, the number of studies performed in males 36

1035 far outweigh those done in females. It is like half of the field is missing. Although more work is

1036 done during development and aging, more work is also clearly needed as the mechanisms are

1037 likely to be age-dependent. Finally, recent studies highlight the necessity to consider circadian and

1038 multidien rhythms, the existence of different classes of seizures, and the impact of housing (single

1039 or colony), which can add the confounding factor of stress. Perhaps accepting the diversity of the

1040 different models and experimental conditions has to offer is the best way to understand human

1041 phenotypes' diversity.

1042

37

1043 References

1044

1045 AICARDI, J., & CHEVRIE, J. J. (1970). Convulsive Status Epilepticus in Infants and Children. Epi-

1046 lepsia, 11(2), 187–197. JOUR. http://doi.org/10.1111/j.1528-1157.1970.tb03880.x

1047

1048 Andermann, F., Kobayashi, E. and Andermann, E. (2005), Genetic Focal Epilepsies: State of the

1049 Art and Paths to the Future. Epilepsia, 46: 61-67. doi:10.1111/j.1528-1167.2005.00361.x

1050

1051 Anthony Altar, C., & Baudry, M. (1990). Systemic injection of kainic acid: Gliosis in olfactory and

1052 limbic brain regions quantified with [3H]PK 11195 binding autoradiography. Experimental Neurolo-

1053 gy, 109(3), 333–341. JOUR. https://doi.org/10.1016/S0014-4886(05)80024-X

1054

1055 Araki, T., Simon, R. P., Taki, W., Lan, J.-Q., & Henshall, D. C. (2002). Characterization of neuronal

1056 death induced by focally evoked limbic seizures in the C57BL/6 mouse. Journal of Neuroscience

1057 Research, 69(5), 614–621. JOUR. http://doi.org/10.1002/jnr.10356

1058

1059 Arellano, J. I., Muñoz, A., Ballesteros-Yáñez, I., Sola, R. G., & DeFelipe, J. (2004). Histopathology

1060 and reorganization of chandelier cells in the human epileptic sclerotic hippocampus. Brain : a jour-

1061 nal of neurology, 127(Pt 1), 45–64. https://doi.org/10.1093/brain/awh004

1062

1063 Arnold, E. C., McMurray, C., Gray, R., & Johnston, D. (2019). Epilepsy-Induced Reduction in HCN

1064 Channel Expression Contributes to an Increased Excitability in Dorsal, But Not Ventral, Hippocam-

1065 pal CA1 Neurons. eNeuro, 6(2), ENEURO.0036-19.2019. https://doi.org/10.1523/ENEURO.0036-

1066 19.2019

1067

1068 Artinian, J., Peret, A., Mircheva, Y., Marti, G. and Crépel, V. (2015), Impaired neuronal operation

1069 through aberrant intrinsic plasticity in epilepsy. Ann Neurol., 77: 592-606. doi:10.1002/ana.24348

1070

38

1071 Bahn, S., Volk, B., & Wisden, W. (1994). Kainate receptor gene expression in the developing rat

1072 brain. The Journal of Neuroscience, 14(9), 5525 LP-5547. JOUR.

1073 http://doi.org/10.1523/JNEUROSCI.14-09-05525.1994

1074

1075 Barba, C., Barbati, G., Minotti, L., Hoffmann, D., & Kahane, P. (2007). Ictal clinical and scalp-EEG

1076 findings differentiating temporal lobe epilepsies from temporal 'plus' epilepsies. Brain : a journal of

1077 neurology, 130(Pt 7), 1957–1967. https://doi.org/10.1093/brain/awm108

1078

1079 Barnes, C. A., Rao, G., & McNaughton, B. L. (1987). Increased electrotonic coupling in aged rat

1080 hippocampus: A possible mechanism for cellular excitability changes. Journal of Comparative Neu-

1081 rology, 259(4), 549–558. JOUR. http://doi.org/10.1002/cne.902590405

1082

1083 de Barros Lourenço, F. H., Marques, L., & de Araujo Filho, G. M. (2020). Electroencephalogram

1084 alterations associated with psychiatric disorders in temporal lobe epilepsy with mesial sclerosis: A

1085 systematic review. Epilepsy & behavior : E&B, 108, 107100.

1086 https://doi.org/10.1016/j.yebeh.2020.107100

1087

1088 Bartolomei, F., Chauvel, P., & Wendling, F. (2008). Epileptogenicity of brain structures in human

1089 temporal lobe epilepsy: a quantified study from intracerebral EEG. Brain : a journal of neurology,

1090 131(Pt 7), 1818–1830. https://doi.org/10.1093/brain/awn111

1091

1092 Baud, M. O., Ghestem, A., Benoliel, J.-J., Becker, C., & Bernard, C. (2019). Endogenous multidien

1093 rhythm of epilepsy in rats. Experimental Neurology, 315, 82–87. JOUR.

1094 https://doi.org/10.1016/j.expneurol.2019.02.006

1095

1096 Baud, M. O., Kleen, J. K., Mirro, E. A., Andrechak, J. C., King-Stephens, D., Chang, E. F., & Rao,

1097 V. R. (2018). Multi-day rhythms modulate seizure risk in epilepsy. Nature Communications, 9(1),

1098 88. JOUR. http://doi.org/10.1038/s41467-017-02577-y

1099 39

1100 Beck, J., Lenart, B., Kintner, D. B., & Sun, D. (2003). Na-K-Cl Cotransporter Contributes to Gluta-

1101 mate-Mediated Excitotoxicity. The Journal of Neuroscience, 23(12), 5061 LP-5068. JOUR.

1102 http://doi.org/10.1523/JNEUROSCI.23-12-05061.2003

1103

1104 Becker, C., Bouvier, E., Ghestem, A., Siyoucef, S., Claverie, D., Camus, F., Bartolomei, F., Be-

1105 noliel, J. J., & Bernard, C. (2015). Predicting and treating stress-induced vulnerability to epilepsy

1106 and depression. Annals of neurology, 78(1), 128–136. https://doi.org/10.1002/ana.24414

1107

1108 Becker, C., Mancic, A., Ghestem, A., Poillerat, V., Claverie, D., Bartolomei, F., Brouillard, F., Be-

1109 noliel, J. J., & Bernard, C. (2019). Antioxidant treatment after epileptogenesis onset prevents

1110 comorbidities in rats sensitized by a past stressful event. Epilepsia, 60(4), 648–655.

1111 https://doi.org/10.1111/epi.14692

1112

1113 Bedner, P., Dupper, A., Hüttmann, K., Müller, J., Herde, M. K., Dublin, P., … Steinhäuser, C.

1114 (2015). Astrocyte uncoupling as a cause of human temporal lobe epilepsy. Brain, 138(5), 1208–

1115 1222. JOUR. http://doi.org/10.1093/brain/awv067

1116

1117 Ben-Ari, Y., & Lagowska, J. (1978). [Epileptogenic action of intra-amygdaloid injection of kainic ac-

1118 id]. Comptes rendus hebdomadaires des seances de l’Academie des sciences. Serie D: Sciences

1119 naturelles, 287(8), 813–816. JOUR. http://europepmc.org/abstract/MED/103652

1120

1121 Ben-Ari, Y., Lagowska, J., Tremblay, E., & Le Gal La Salle, G. (1979). A new model of focal status

1122 epilepticus: intra-amygdaloid application of kainic acid elicits repetitive secondarily generalized

1123 convulsive seizures. Brain Research, 163(1), 176–179. JOUR. https://doi.org/10.1016/0006-

1124 8993(79)90163-X

1125

1126 Ben-Ari, Y., Tremblay, E., Ottersen, O. P., & Meldrum, B. S. (1980). The role of epileptic activity in

1127 hippocampal and “remote” cerebral lesions induced by kainic acid. Brain Research, 191(1), 79–97.

1128 JOUR. https://doi.org/10.1016/0006-8993(80)90316-9 40

1129

1130 Benke, T. A., & Swann, J. (2004). The Tetanus Toxin Model of Chronic Epilepsy BT - Recent Ad-

1131 vances in Epilepsy Research. In D. K. Binder & H. E. Scharfman (Eds.) (pp. 226–238). CHAP,

1132 Boston, MA: Springer US. http://doi.org/10.1007/978-1-4757-6376-8_16

1133

1134 Benkovic, S. A., O’Callaghan, J. P., & Miller, D. B. (2004). Sensitive indicators of injury reveal hip-

1135 pocampal damage in C57BL/6J mice treated with kainic acid in the absence of tonic–clonic sei-

1136 zures. Brain Research, 1024(1), 59–76. JOUR. http://doi.org/10.1016/j.brainres.2004.07.021

1137

1138 Berkovic, S. F., Andermann, F., Olivier, A., Ethier, R., Melanson, D., Robitaille, Y., … Feindel, W.

1139 (1991). Hippocampal sclerosis in temporal lobe epilepsy demonstrated by magnetic resonance im-

1140 aging. Annals of Neurology, 29(2), 175–182. JOUR. http://doi.org/10.1002/ana.410290210

1141

1142 Bernard, P. B., Macdonald, D. S., Gill, D. A., Ryan, C. L., & Tasker, R. A. (2007). Hippocampal

1143 mossy fiber sprouting and elevated trkB receptor expression following systemic administration of

1144 low dose domoic acid during neonatal development. Hippocampus, 17(11), 1121–1133.

1145 https://doi.org/10.1002/hipo.20342

1146

1147 Bernard, C. (2019). On the interpretation of results obtained in singly housed animals. Epilepsia,

1148 60(10), 2013–2015. JOUR. http://doi.org/10.1111/epi.16347

1149

1150 Bernard C. (2020). Circadian/multidien Molecular Oscillations and Rhythmicity of Epilepsy

1151 (MORE). Epilepsia, 10.1111/epi.16716. Advance online publication.

1152 https://doi.org/10.1111/epi.16716

1153

1154 Bertoglio, D., Amhaoul, H., Van Eetveldt, A., Houbrechts, R., Van De Vijver, S., Ali, I., &

1155 Dedeurwaerdere, S. (2017). Kainic Acid-Induced Post-Status Epilepticus Models of Temporal Lobe

1156 Epilepsy with Diverging Seizure Phenotype and Neuropathology. Frontiers in Neurology, 8, 588.

1157 JOUR. http://doi.org/10.3389/fneur.2017.00588 41

1158

1159 Bielefeld, P., Sierra, A., Encinas, J. M., Maletic-Savatic, M., Anderson, A., & Fitzsimons, C. P.

1160 (2017). A Standardized Protocol for Stereotaxic Intrahippocampal Administration of Kainic Acid

1161 Combined with Electroencephalographic Seizure Monitoring in Mice. Frontiers in Neuroscience,

1162 11, 160. JOUR. http://doi.org/10.3389/fnins.2017.00160

1163

1164 Bloss, E. B., & Hunter, R. G. (2010). Chapter 9 - Hippocampal Kainate Receptors. In G. B. T.-V. &

1165 H. Litwack (Ed.), Hormones of the Limbic System (Vol. 82, pp. 167–184). CHAP, Academic Press.

1166 https://doi.org/10.1016/S0083-6729(10)82009-6

1167

1168 MD, I. B., MRCPath, M. T., & Wiestler, O. D. (2002). Ammon’s Horn Sclerosis: A Maldevelopmen-

1169 tal Disorder Associated with Temporal Lobe Epilepsy. Brain Pathology, 12(2), 199–211. JOUR.

1170 http://doi.org/10.1111/j.1750-3639.2002.tb00436.x

1171

1172 Blümcke, I., Thom, M., Aronica, E., Armstrong, D. D., Bartolomei, F., Bernasconi, A., Bernasconi,

1173 N., Bien, C. G., Cendes, F., Coras, R., Cross, J. H., Jacques, T. S., Kahane, P., Mathern, G. W.,

1174 Miyata, H., Moshé, S. L., Oz, B., Özkara, Ç., Perucca, E., Sisodiya, S., … Spreafico, R. (2013).

1175 International consensus classification of hippocampal sclerosis in temporal lobe epilepsy: a Task

1176 Force report from the ILAE Commission on Diagnostic Methods. Epilepsia, 54(7), 1315–1329.

1177 https://doi.org/10.1111/epi.12220

1178

1179 Bouilleret, V., Ridoux, V., Depaulis, A., Marescaux, C., Nehlig, A., & Le Gal La Salle, G. (1999).

1180 Recurrent seizures and hippocampal sclerosis following intrahippocampal kainate injection in adult

1181 mice: electroencephalography, histopathology and synaptic reorganization similar to mesial tem-

1182 poral lobe epilepsy. Neuroscience, 89(3), 717–729. JOUR. https://doi.org/10.1016/S0306-

1183 4522(98)00401-1

1184

42

1185 Bragin, A., Engel Jr., J., Wilson, C. L., Vizentin, E., & Mathern, G. W. (1999). Electrophysiologic

1186 Analysis of a Chronic Seizure Model After Unilateral Hippocampal KA Injection. Epilepsia, 40(9),

1187 1210–1221. JOUR. http://doi.org/10.1111/j.1528-1157.1999.tb00849.x

1188

1189 Bragin, A., Azizyan, A., Almajano, J., Wilson, C. L., & Engel Jr., J. (2005). Analysis of Chronic Sei-

1190 zure Onsets after Intrahippocampal Kainic Acid Injection in Freely Moving Rats. Epilepsia, 46(10),

1191 1592–1598. JOUR. http://doi.org/10.1111/j.1528-1167.2005.00268.x

1192

1193 Brancati, G. E., Rawas, C., Ghestem, A., Bernard, C., & Ivanov, A. I. (2020). Spatio-temporal het-

1194 erogeneity in hippocampal metabolism in control and epilepsy conditions. BioRxiv,

1195 2020.06.17.155044. https://doi.org/10.1101/2020.06.17.155044

1196

1197 Brandt, C., Töllner, K., Klee, R., Bröer, S., & Löscher, W. (2015). Effective termination of status

1198 epilepticus by rational polypharmacy in the lithium-pilocarpine model in rats: Window of opportunity

1199 to prevent epilepsy and prediction of epilepsy by biomarkers. Neurobiology of disease, 75, 78–90.

1200 https://doi.org/10.1016/j.nbd.2014.12.015

1201

1202 Bregestovski, P., Maleeva, G., & Gorostiza, P. (2018). Light-induced regulation of ligand-gated

1203 channel activity. British journal of pharmacology, 175(11), 1892–1902.

1204 https://doi.org/10.1111/bph.14022

1205

1206 Canto, A. M., Matos, A., Godoi, A. B., Vieira, A. S., Aoyama, B. B., Rocha, C. S., Henning, B., Car-

1207 valho, B. S., Pascoal, V., Veiga, D., Gilioli, R., Cendes, F., & Lopes-Cendes, I. (2020). Multi-omics

1208 analysis suggests enhanced epileptogenesis in the Cornu Ammonis 3 of the pilocarpine model of

1209 mesial temporal lobe epilepsy. Hippocampus, 10.1002/hipo.23268. Advance online publication.

1210 https://doi.org/10.1002/hipo.23268

1211

1212 do Canto, A. M., Vieira, A. S., H B Matos, A., Carvalho, B. S., Henning, B., Norwood, B. A., Bauer,

1213 S., Rosenow, F., Gilioli, R., Cendes, F., & Lopes-Cendes, I. (2020). Laser microdissection-based 43

1214 microproteomics of the hippocampus of a rat epilepsy model reveals regional differences in protein

1215 abundances. Scientific reports, 10(1), 4412. https://doi.org/10.1038/s41598-020-61401-8

1216

1217 Casillas-Espinosa, P. M., Powell, K. L., & O’Brien, T. J. (2012). Regulators of synaptic transmis-

1218 sion: Roles in the pathogenesis and treatment of epilepsy. Epilepsia, 53(s9), 41–58. JOUR.

1219 http://doi.org/10.1111/epi.12034

1220

1221 Cavalheiro, E. A., Riche, D. A., & Le Gal La Salle, G. (1982). Long-term effects of intrahippocam-

1222 pal kainic acid injection in rats: A method for inducing spontaneous recurrent seizures. Electroen-

1223 cephalography and Clinical Neurophysiology, 53(6), 581–589. JOUR. https://doi.org/10.1016/0013-

1224 4694(82)90134-1

1225

1226 Cela, E., McFarlan, A. R., Chung, A. J., Wang, T., Chierzi, S., Murai, K. K., & Sjöström, P. J.

1227 (2019). An Optogenetic Kindling Model of Neocortical Epilepsy. Scientific Reports, 9(1), 5236.

1228 JOUR. http://doi.org/10.1038/s41598-019-41533-2

1229

1230 Chauvière, L., Doublet, T., Ghestem, A., Siyoucef, S. S., Wendling, F., Huys, R., … Bernard, C.

1231 (2012). Changes in interictal spike features precede the onset of temporal lobe epilepsy. Annals of

1232 Neurology, 71(6), 805–814. JOUR. http://doi.org/10.1002/ana.23549

1233

1234 Chen, Q., Olney, J. W., Lukasiewicz, P. D., Almli, T., & Romano, C. (1998). Ca+-Independent Exci-

1235 totoxic Neurodegeneration in Isolated Retina, an Intact Neural Net: A Role for Cl− and Inhibitory

1236 Transmitters. Molecular Pharmacology, 53(3), 564 LP-572. JOUR.

1237 http://doi.org/10.1124/mol.53.3.564

1238

1239 Chen, Z., Ljunggren, H.-G., Bogdanovic, N., Nennesmo, I., Winblad, B., & Zhu, J. (2002). Exci-

1240 totoxic neurodegeneration induced by intranasal administration of kainic acid in C57BL/6 mice.

1241 Brain Research, 931(2), 135–145.JOUR.https://doi.org/10.1016/S0006-8993(02)02268-0

1242 44

1243 Cho, J., Lee, S.-H., Seo, J. H., Kim, H.-S., Ahn, J.-G., Kim, S.-S., Yim, S.-V., Song, D.-K., & Cho,

1244 S. S. (2002). Increased expression of phosphatase and tensin homolog in reactive astrogliosis fol-

1245 lowing intracerebroventricular kainic acid injection in mouse hippocampus. Neuroscience Letters,

1246 334(2), 131–134. https://doi.org/10.1016/S0304-3940(02)01122-9

1247

1248 Chuang, Y.-C., Chang, A. Y. W., Lin, J.-W., Hsu, S.-P., & Chan, S. H. H. (2004). Mitochondrial

1249 Dysfunction and Ultrastructural Damage in the Hippocampus during Kainic Acid–induced Status

1250 Epilepticus in the Rat. Epilepsia, 45(10), 1202–1209. JOUR. http://doi.org/10.1111/j.0013-

1251 9580.2004.18204.x

1252

1253 Cossart, R., Dinocourt, C., Hirsch, J. C., Merchan-Perez, A., De Felipe, J., Ben-Ari, Y., … Bernard,

1254 C. (2001). Dendritic but not somatic GABAergic inhibition is decreased in experimental epilepsy.

1255 Nature Neuroscience, 4(1), 52–62. JOUR. http://doi.org/10.1038/82900

1256

1257 Cossart, R., Tyzio, R., Dinocourt, C., Esclapez, M., Hirsch, J. C., Ben-Ari, Y., & Bernard, C. (2001).

1258 Presynaptic Kainate Receptors that Enhance the Release of GABA on CA1 Hippocampal Interneu-

1259 rons. Neuron, 29(2), 497–508. JOUR. http://doi.org/https://doi.org/10.1016/S0896-6273(01)00221-

1260 5

1261

1262 Covolan, L., & Mello, L. E. A. . (2000). Temporal profile of neuronal injury following pilocarpine or

1263 kainic acid-induced status epilepticus. Epilepsy Research, 39(2), 133–152. JOUR.

1264 https://doi.org/10.1016/S0920-1211(99)00119-9

1265

1266 Crespel, A., Lebrun, A., de Bock, F., Bockaert, J., Picot, M.-C., Rousset, M.-C., … Rigau, V.

1267 (2007). Angiogenesis is associated with blood–brain barrier permeability in temporal lobe epilepsy.

1268 Brain, 130(7), 1942–1956. JOUR. http://doi.org/10.1093/brain/awm118

1269

45

1270 Crisp, D. N., Cheung, W., Gliske, S. V, Lai, A., Freestone, D. R., Grayden, D. B., … Stacey, W. C.

1271 (2020). Quantifying epileptogenesis in rats with spontaneous and responsive brain state dynamics.

1272 Brain Communications, 2(1). JOUR. http://doi.org/10.1093/braincomms/fcaa048

1273

1274 Curia, G., Longo, D., Biagini, G., Jones, R. S. G., & Avoli, M. (2008). The pilocarpine model of

1275 temporal lobe epilepsy. Journal of Neuroscience Methods, 172(2), 143–157. JOUR.

1276 http://doi.org/https://doi.org/10.1016/j.jneumeth.2008.04.019

1277

1278 D’Ambrosio, R., Fender, J. S., Fairbanks, J. P., Simon, E. A., Born, D. E., Doyle, D. L., & Miller, J.

1279 W. (2004). Progression from frontal–parietal to mesial–temporal epilepsy after fluid percussion in-

1280 jury in the rat. Brain, 128(1), 174–188. JOUR. http://doi.org/10.1093/brain/awh337

1281

1282 Darbin, O., Naritoku, D., & Patrylo, P. R. (2004). Aging Alters Electroencephalographic and Clinical

1283 Manifestations of Kainate-induced Status Epilepticus. Epilepsia, 45(10), 1219–1227. JOUR.

1284 http://doi.org/10.1111/j.0013-9580.2004.66103.x

1285

1286 Darstein, M., Petralia, R. S., Swanson, G. T., Wenthold, R. J., & Heinemann, S. F. (2003). Distribu-

1287 tion of Kainate Receptor Subunits at Hippocampal Mossy Fiber Synapses. The Journal of Neuro-

1288 science, 23(22), 8013 LP-8019. JOUR. http://doi.org/10.1523/JNEUROSCI.23-22-08013.2003

1289

1290 Das, A., Wallace, G. C., Holmes, C., McDowell, M. L., Smith, J. A., Marshall, J. D., Bonilha, L.,

1291 Edwards, J. C., Glazier, S. S., Ray, S. K., & Banik, N. L. (2012). Hippocampal tissue of patients

1292 with refractory temporal lobe epilepsy is associated with astrocyte activation, inflammation, and

1293 altered expression of channels and receptors. Neuroscience, 220, 237–246.

1294 https://doi.org/https://doi.org/10.1016/j.neuroscience.2012.06.002

1295

1296 Debski, K. J., Ceglia, N., Ghestem, A., Ivanov, A. I., Brancati, G. E., Bröer, S., Bot, A. M., Müller, J.

1297 A., Schoch, S., Becker, A., Löscher, W., Guye, M., Sassone-Corsi, P., Lukasiuk, K., Baldi, P., &

1298 Bernard, C. (2020). The circadian dynamics of the hippocampal transcriptome and proteome is al- 46

1299 tered in experimental temporal lobe epilepsy. Science advances, 6(41), eaat5979.

1300 https://doi.org/10.1126/sciadv.aat5979

1301

1302 Dong, H. W., Swanson, L. W., Chen, L., Fanselow, M. S., & Toga, A. W. (2009). Genomic-

1303 anatomic evidence for distinct functional domains in hippocampal field CA1. Proceedings of the

1304 National Academy of Sciences of the United States of America, 106(28), 11794–11799.

1305 https://doi.org/10.1073/pnas.0812608106

1306

1307 Dong, X., Wang, Y., & Qin, Z. (2009). Molecular mechanisms of excitotoxicity and their relevance

1308 to pathogenesis of neurodegenerative diseases. Acta Pharmacologica Sinica, 30(4), 379–387.

1309 JOUR. http://doi.org/10.1038/aps.2009.24

1310

1311 Drexel, M., Preidt, A. P., & Sperk, G. (2012). Sequel of spontaneous seizures after kainic acid-

1312 induced status epilepticus and associated neuropathological changes in the subiculum and ento-

1313 rhinal cortex. Neuropharmacology, 63(5), 806–817. JOUR.

1314 http://doi.org/10.1016/j.neuropharm.2012.06.009

1315

1316 Duan, R.-S., Chen, Z., Dou, Y.-C., Concha Quezada, H., Nennesmo, I., Adem, A., … Zhu, J.

1317 (2006). Apolipoprotein E deficiency increased microglial activation/CCR3 expression and hippo-

1318 campal damage in kainic acid exposed mice. Experimental Neurology, 202(2), 373–380. JOUR.

1319 https://doi.org/10.1016/j.expneurol.2006.06.013

1320

1321 Dubé, C. M., Ravizza, T., Hamamura, M., Zha, Q., Keebaugh, A., Fok, K., … Baram, T. Z. (2010).

1322 Epileptogenesis Provoked by Prolonged Experimental Febrile Seizures: Mechanisms and Bi-

1323 omarkers. The Journal of Neuroscience, 30(22), 7484 LP-7494. JOUR.

1324 http://doi.org/10.1523/JNEUROSCI.0551-10.2010

1325

47

1326 Ekstrand, J. J., Pouliot, W., Scheerlinck, P., & Dudek, F. E. (2011). Lithium pilocarpine-induced

1327 status epilepticus in postnatal day 20 rats results in greater neuronal injury in ventral versus dorsal

1328 hippocampus. Neuroscience, 192, 699–707. https://doi.org/10.1016/j.neuroscience.2011.05.022

1329

1330 El Houssaini, K., Ivanov, A. I., Bernard, C., & Jirsa, V. K. (2015). Seizures, refractory status epilep-

1331 ticus, and depolarization block as endogenous brain activities. Physical review. E, Statistical, non-

1332 linear, and soft matter physics, 91(1), 010701. https://doi.org/10.1103/PhysRevE.91.010701

1333

1334 Engel, J. (2001). Mesial Temporal Lobe Epilepsy: What Have We Learned? The Neuroscientist,

1335 7(4), 340–352. JOUR. http://doi.org/10.1177/107385840100700410

1336

1337 Engel, J., McDermott, M. P., Wiebe, S., Langfitt, J. T., Stern, J. M., Dewar, S., … Early Random-

1338 ized Surgical Epilepsy Trial (ERSET) Study Group, for the. (2012). Early Surgical Therapy for

1339 Drug-Resistant Temporal Lobe Epilepsy: A Randomized Trial. JAMA, 307(9), 922–930. JOUR.

1340 http://doi.org/10.1001/jama.2012.220

1341

1342 Evans, M. C., & Dougherty, K. A. (2018). Carbamazepine-induced suppression of repetitive firing

1343 in CA1 pyramidal neurons is greater in the dorsal hippocampus than the ventral hippocampus. Epi-

1344 lepsy research, 145, 63–72. https://doi.org/10.1016/j.eplepsyres.2018.05.014

1345

1346 Ferraro, T. N., Golden, G. T., Smith, G. G., & Berrettini, W. H. (1995). Differential Susceptibility to

1347 Seizures Induced by Systemic Kainic Acid Treatment in Mature DBA/2J and C57BLl6J Mice. Epi-

1348 lepsia, 36(3), 301–307. JOUR. http://doi.org/10.1111/j.1528-1157.1995.tb00999.x

1349

1350 Fernandes, H. B., Catches, J. S., Petralia, R. S., Copits, B. A., Xu, J., Russell, T. A., … Contractor,

1351 A. (2009). High-Affinity Kainate Receptor Subunits Are Necessary for Ionotropic but Not Metabo-

1352 tropic Signaling. Neuron, 63(6), 818–829. JOUR.

1353 http://doi.org/https://doi.org/10.1016/j.neuron.2009.08.010

1354 48

1355 Fisahn, A., Contractor, A., Traub, R. D., Buhl, E. H., Heinemann, S. F., & McBain, C. J. (2004).

1356 Distinct Roles for the Kainate Receptor Subunits GluR5 and GluR6 in Kainate-Induced Hippocam-

1357 pal Gamma Oscillations. The Journal of Neuroscience, 24(43), 9658 LP-9668. JOUR.

1358 http://doi.org/10.1523/JNEUROSCI.2973-04.2004

1359

1360 Fonnum, F., & Walaas, I. (1978). THE EFFECT OF INTRAHIPPOCAMPAL KAINIC ACID INJEC-

1361 TIONS AND SURGICAL LESIONS ON NEUROTRANSMITTERS IN HIPPOCAMPUS AND SEP-

1362 TUM. Journal of Neurochemistry, 31(5), 1173–1181. JOUR. http://doi.org/10.1111/j.1471-

1363 4159.1978.tb06241.x

1364

1365 French, E. D., Aldinio, C., & Schwarcz, R. (1982). Intrahippocampal kainic acid, seizures and local

1366 neuronal degeneration: Relationships assessed in unanesthetized rats. Neuroscience, 7(10),

1367 2525–2536. JOUR. https://doi.org/10.1016/0306-4522(82)90212-3

1368

1369 French, J. A., Williamson, P. D., Thadani, V. M., Darcey, T. M., Mattson, R. H., Spencer, S. S., &

1370 Spencer, D. D. (1993). Characteristics of medial temporal lobe epilepsy: I. Results of history and

1371 physical examination. Annals of Neurology, 34(6), 774–780. JOUR.

1372 http://doi.org/10.1002/ana.410340604

1373

1374 Frey, W. H., Liu, J., Chen, X., Thorne, R. G., Fawcett, J. R., Ala, T. A., & Rahman, Y.-E. (1997).

1375 Delivery of 125I-NGF to the Brain via the Olfactory Route. Drug Delivery, 4(2), 87–92. JOUR.

1376 http://doi.org/10.3109/10717549709051878

1377

1378 Friedman, L. K. (2006). Calcium: a role for neuroprotection and sustained adaptation. Molecular

1379 Interventions, 6(6), 315–329. JOUR. http://doi.org/10.1124/mi.6.6.5

1380

1381 Fritsch, B., Qashu, F., Figueiredo, T. H., Aroniadou-Anderjaska, V., Rogawski, M. A., & Braga, M.

1382 F. M. (2009). Pathological alterations in GABAergic interneurons and reduced tonic inhibition in the

49

1383 basolateral amygdala during epileptogenesis. Neuroscience, 163(1), 415–429. JOUR.

1384 http://doi.org/10.1016/j.neuroscience.2009.06.034

1385

1386 Fritsch, B., Reis, J., Gasior, M., Kaminski, R. M., & Rogawski, M. A. (2014). Role of GluK1 Kainate

1387 Receptors in Seizures, Epileptic Discharges, and Epileptogenesis. The Journal of Neuroscience,

1388 34(17), 5765 LP – 5775. https://doi.org/10.1523/JNEUROSCI.5307-13.2014

1389

1390 Fritsch, B., Stott, J. J., Joelle Donofrio, J., & Rogawski, M. A. (2010). Treatment of early and late

1391 kainic acid-induced status epilepticus with the noncompetitive AMPA receptor antagonist GYKI

1392 52466. Epilepsia, 51(1), 108–117. JOUR. http://doi.org/10.1111/j.1528-1167.2009.02205.x

1393

1394 Fujikawa, D. G. (2005). Prolonged seizures and cellular injury: Understanding the connection. Epi-

1395 lepsy & Behavior, 7, 3–11. https://doi.org/10.1016/j.yebeh.2005.08.003

1396

1397

1398 Gallyas Jr, F., Ball, S. M., & Molnar, E. (2003). Assembly and cell surface expression of KA-2 sub-

1399 unit-containing kainate receptors. Journal of Neurochemistry, 86(6), 1414–1427. JOUR.

1400 http://doi.org/10.1046/j.1471-4159.2003.01945.x

1401

1402 Gano, L. B., Liang, L.-P., Ryan, K., Michel, C. R., Gomez, J., Vassilopoulos, A., … Patel, M.

1403 (2018). Altered mitochondrial acetylation profiles in a kainic acid model of temporal lobe epilepsy.

1404 Free Radical Biology and Medicine, 123, 116–124. JOUR.

1405 https://doi.org/10.1016/j.freeradbiomed.2018.05.063

1406

1407 Gao, X., Guo, M., Meng, D., Sun, F., Guan, L., Cui, Y., … Qi, S. (2019). Silencing MicroRNA-134

1408 Alleviates Hippocampal Damage and Occurrence of Spontaneous Seizures After Intraventricular

1409 Kainic Acid-Induced Status Epilepticus in Rats. Frontiers in Cellular Neuroscience, 13, 145. JOUR.

1410 http://doi.org/10.3389/fncel.2019.00145

1411 50

1412 Giorgi, F. S., Malhotra, S., Hasson, H., Velíšková, J., Rosenbaum, D. M., & Moshé, S. L. (2005).

1413 Effects of Status Epilepticus Early in Life on Susceptibility to Ischemic Injury in Adulthood. Epilep-

1414 sia, 46(4), 490–498. JOUR. http://doi.org/10.1111/j.0013-9580.2005.42304.x

1415

1416 GODDARD, G. V. (1967). Development of Epileptic Seizures through Brain Stimulation at Low In-

1417 tensity. Nature, 214(5092), 1020–1021. JOUR. http://doi.org/10.1038/2141020a0

1418

1419 Goddard, G. V, McIntyre, D. C., & Leech, C. K. (1969). A permanent change in brain function re-

1420 sulting from daily electrical stimulation. Experimental Neurology, 25(3), 295–330. JOUR.

1421 https://doi.org/10.1016/0014-4886(69)90128-9

1422

1423 Goddard, G. V. (1983). The kindling model of epilepsy. Trends in Neurosciences, 6, 275–279.

1424 https://doi.org/10.1016/0166-2236(83)90118-2

1425

1426 Goldenholz, D. M., Rakesh, K., Kapur, K., Gaínza-Lein, M., Hodgeman, R., Moss, R., Theodore,

1427 W. H., & Loddenkemper, T. (2018). Different as night and day: Patterns of isolated seizures, clus-

1428 ters, and status epilepticus. Epilepsia, 59(5), e73–e77. https://doi.org/10.1111/epi.14076

1429

1430 Gomez-Marin, A., & Ghazanfar, A. A. (2019). The Life of Behavior. Neuron, 104(1), 25–36.

1431 https://doi.org/10.1016/j.neuron.2019.09.017

1432

1433 Gorter, J. A., Van Vliet, E. A., Aronica, E., & Da Silva, F. H. L. (2001). Progression of spontaneous

1434 seizures after status epilepticus is associated with mossy fibre sprouting and extensive bilateral

1435 loss of hilar parvalbumin and somatostatin-immunoreactive neurons. European Journal of Neuro-

1436 science, 13(4), 657–669. JOUR. http://doi.org/10.1046/j.1460-9568.2001.01428.x

1437

1438 Gouder, N., Fritschy, J.-M., & Boison, D. (2003). Seizure Suppression by Adenosine A1 Receptor

1439 Activation in a Mouse Model of Pharmacoresistant Epilepsy. Epilepsia, 44(7), 877–885. JOUR.

1440 http://doi.org/10.1046/j.1528-1157.2003.03603.x 51

1441

1442 Gröticke, I., Hoffmann, K., & Löscher, W. (2008). Behavioral alterations in a mouse model of tem-

1443 poral lobe epilepsy induced by intrahippocampal injection of kainate. Experimental Neurology,

1444 213(1), 71–83. JOUR. https://doi.org/10.1016/j.expneurol.2008.04.036

1445

1446 Guedj, E., Bonini, F., Gavaret, M., Trébuchon, A., Aubert, S., Boucekine, M., Boyer, L., Carron, R.,

1447 McGonigal, A., & Bartolomei, F. (2015). 18FDG-PET in different subtypes of temporal lobe epilep-

1448 sy: SEEG validation and predictive value. Epilepsia, 56(3), 414–421.

1449 https://doi.org/10.1111/epi.12917

1450

1451 Gurbanova, A. A., Aker, R. G., Sirvanci, S., Demiralp, T., & Onat, F. Y. (2008). Intra-Amygdaloid

1452 Injection of Kainic Acid in Rats with Genetic Absence Epilepsy: The Relationship of Typical Ab-

1453 sence Epilepsy and Temporal Lobe Epilepsy. The Journal of Neuroscience, 28(31), 7828 LP-7836.

1454 JOUR. http://doi.org/10.1523/JNEUROSCI.1097-08.2008

1455

1456 Ha, B. K., Vicini, S., Rogers, R. C., Bresnahan, J. C., Burry, R. W., & Beattie, M. S. (2002).

1457 Kainate-induced excitotoxicity is dependent upon extracellular potassium concentrations that regu-

1458 late the activity of AMPA/KA type glutamate receptors. Journal of Neurochemistry, 83(4), 934–945.

1459 JOUR. http://doi.org/10.1046/j.1471-4159.2002.01203.x

1460

1461 Haas, K. Z., Sperber, E. F., Opanashuk, L. A., Stanton, P. K., & Moshé, S. L. (2001). Resistance of

1462 immature hippocampus to morphologic and physiologic alterations following status epilepticus or

1463 kindling. Hippocampus, 11(6), 615–625. JOUR. http://doi.org/10.1002/hipo.1076

1464

1465 Hellier, J. L., & Dudek, F. E. (2005). Chemoconvulsant Model of Chronic Spontaneous Seizures.

1466 Current Protocols in Neuroscience, 31(1), 9.19.1-9.19.12. JOUR.

1467 http://doi.org/10.1002/0471142301.ns0919s31

1468

52

1469 Hochgerner, H., Zeisel, A., Lönnerberg, P., & Linnarsson, S. (2018). Conserved properties of den-

1470 tate gyrus neurogenesis across postnatal development revealed by single-cell RNA sequencing.

1471 Nature Neuroscience, 21(2), 290–299. https://doi.org/10.1038/s41593-017-0056-2

1472

1473 Holmes, G. L., & Thompson, J. L. (1988). Effects of kainic acid on seizure susceptibility in the de-

1474 veloping brain. Developmental Brain Research, 39(1), 51–59. JOUR. https://doi.org/10.1016/0165-

1475 3806(88)90066-1

1476

1477 Holmes G. L. (2015). Cognitive impairment in epilepsy: the role of network abnormalities. Epileptic

1478 disorders : international epilepsy journal with videotape, 17(2), 101–116.

1479 https://doi.org/10.1684/epd.2015.0739

1480

1481 Holtzman, D., Obana, K., & Olson, J. (1981). Hyperthermia-induced seizures in the rat pup: a

1482 model for febrile convulsions in children. Science, 213(4511), 1034 LP-1036. JOUR.

1483 http://doi.org/10.1126/science.7268407

1484

1485 Hu, R. Q., Koh, S., Torgerson, T., & Cole, A. J. (1998). Neuronal stress and injury in C57/BL mice

1486 after systemic kainic acid administration. Brain Research, 810(1), 229–240. JOUR.

1487 http://doi.org/10.1016/S0006-8993(98)00863-4

1488

1489 Huettner, J. E. (2003). Kainate receptors and synaptic transmission. Progress in Neurobiology,

1490 70(5), 387–407. JOUR. http://doi.org/https://doi.org/10.1016/S0301-0082(03)00122-9

1491

1492 Inostroza, M., Cid, E., Brotons-Mas, J., Gal, B., Aivar, P., Uzcategui, Y. G., … Menendez de la

1493 Prida, L. (2011). Hippocampal-Dependent Spatial Memory in the Water Maze is Preserved in an

1494 Experimental Model of Temporal Lobe Epilepsy in Rats. PLOS ONE, 6(7), e22372. JOUR. Re-

1495 trieved from https://doi.org/10.1371/journal.pone.0022372

1496

53

1497 Isaeva, E., Romanov, A., Holmes, G. L., & Isaev, D. (2015). Status epilepticus results in region-

1498 specific alterations in seizure susceptibility along the hippocampal longitudinal axis. Epilepsy re-

1499 search, 110, 166–170. https://doi.org/10.1016/j.eplepsyres.2014.12.009

1500

1501 Jagirdar, R., Drexel, M., Kirchmair, E., Tasan, R. O., & Sperk, G. (2015). Rapid changes in expres-

1502 sion of class I and IV histone deacetylases during epileptogenesis in mouse models of temporal

1503 lobe epilepsy. Experimental Neurology, 273, 92–104. JOUR.

1504 https://doi.org/10.1016/j.expneurol.2015.07.026

1505

1506 JEFFERYS, J. G. R., & WALKER, M. C. (2006). Tetanus toxin model of focal epilepsy. In Models

1507 of seizures and epilepsy (pp. 407–414). CHAP, Elsevier.

1508

1509 Jefferys, J., Steinhäuser, C., & Bedner, P. (2016). Chemically-induced TLE models: Topical appli-

1510 cation. Journal of Neuroscience Methods, 260, 53–61. JOUR.

1511 https://doi.org/10.1016/j.jneumeth.2015.04.011

1512

1513 Jensen, F. E., Applegate, C. D., Holtzman, D., Belin, T. R., & Burchfiel, J. L. (1991). Epileptogenic

1514 effect of hypoxia in the immature rodent brain. Annals of Neurology, 29(6), 629–637. JOUR.

1515 http://doi.org/10.1002/ana.410290610

1516

1517 Jeon, G. S., Shin, D. H., & Cho, S. S. (2004). Induction of transcription factor c-myb expression in

1518 reactive astrocytes following intracerebroventricular kainic acid injection in mouse hippocampus.

1519 Neuroscience Letters, 360(1), 13–16. https://doi.org/10.1016/j.neulet.2004.01.017

1520

1521 Jin, X.-T., & Smith, Y. (2011). Localization and functions of kainate receptors in the basal ganglia.

1522 Advances in Experimental Medicine and Biology, 717, 27–37. JOUR. http://doi.org/10.1007/978-1-

1523 4419-9557-5_3

1524

54

1525 Jin, Y., Lim, C.-M., Kim, S.-W., Park, J.-Y., Seo, J.-S., Han, P.-L., Yoon, S. H., & Lee, J.-K. (2009).

1526 Fluoxetine attenuates kainic acid-induced neuronal cell death in the mouse hippocampus. Brain

1527 Research, 1281, 108–116. https://doi.org/10.1016/j.brainres.2009.04.053

1528

1529 Jing, M., Shingo, T., Yasuhara, T., Kondo, A., Morimoto, T., Wang, F., … Date, I. (2009). The

1530 combined therapy of intrahippocampal transplantation of adult neural stem cells and intraventricu-

1531 lar erythropoietin-infusion ameliorates spontaneous recurrent seizures by suppression of abnormal

1532 mossy fiber sprouting. Brain Research, 1295, 203–217. JOUR.

1533 https://doi.org/10.1016/j.brainres.2009.07.079

1534

1535 Jirsa, V. K., Stacey, W. C., Quilichini, P. P., Ivanov, A. I., & Bernard, C. (2014). On the nature of

1536 seizure dynamics. Brain : a journal of neurology, 137(Pt 8), 2210–2230.

1537 https://doi.org/10.1093/brain/awu133

1538

1539 Kairiss, E. W., Racine, R. J., & Smith, G. K. (1984). The development of the interictal spike during

1540 kindling in the rat. Brain Research, 322(1), 101–110. JOUR. https://doi.org/10.1016/0006-

1541 8993(84)91185-5

1542

1543 Kar, S., Seto, D., Doré, S., Chabot, J.-G., & Quirion, R. (1997). Systemic administration of kainic

1544 acid induces selective time dependent decrease in [125I]insulin-like growth factor I, [125I]insulin-

1545 like growth factor II and [125I]insulin receptor binding sites in adult rat hippocampal formation.

1546 Neuroscience, 80(4), 1041–1055. JOUR. https://doi.org/10.1016/S0306-4522(97)00185-1

1547

1548 Karoly, P. J., Goldenholz, D. M., Freestone, D. R., Moss, R. E., Grayden, D. B., Theodore, W. H.,

1549 & Cook, M. J. (2018). Circadian and circaseptan rhythms in human epilepsy: a retrospective cohort

1550 study. The Lancet Neurology, 17(11), 977–985. JOUR. https://doi.org/10.1016/S1474-

1551 4422(18)30274-6

1552

55

1553 Kasugai, M., Akaike, K., Imamura, S., Matsukubo, H., Tojo, H., Nakamura, M., … Sano, A. (2007).

1554 Differences in two mice strains on kainic acid-induced amygdalar seizures. Biochemical and Bio-

1555 physical Research Communications, 357(4), 1078–1083. JOUR.

1556 https://doi.org/10.1016/j.bbrc.2007.04.067

1557

1558 Kharatishvili, I., Nissinen, J. P., McIntosh, T. K., & Pitkänen, A. (2006). A model of posttraumatic

1559 epilepsy induced by lateral fluid-percussion brain injury in rats. Neuroscience, 140(2), 685–697.

1560 JOUR. http://doi.org/https://doi.org/10.1016/j.neuroscience.2006.03.012

1561

1562 Kharlamov, E. A., Kharlamov, A., & Kelly, K. M. (2007). Changes in neuropeptide Y protein ex-

1563 pression following photothrombotic brain infarction and epileptogenesis. Brain Research, 1127(1),

1564 151–162. JOUR. http://doi.org/10.1016/j.brainres.2006.09.107

1565

1566 Klee, R., Brandt, C., Töllner, K., & Löscher, W. (2017). Various modifications of the intrahippocam-

1567 pal kainate model of mesial temporal lobe epilepsy in rats fail to resolve the marked rat-to-mouse

1568 differences in type and frequency of spontaneous seizures in this model. Epilepsy & Behavior, 68,

1569 129–140. JOUR. https://doi.org/10.1016/j.yebeh.2016.11.035

1570

1571 Kobayashi, M., & Buckmaster, P. S. (2003). Reduced Inhibition of Dentate Granule Cells in a Mod-

1572 el of Temporal Lobe Epilepsy. The Journal of Neuroscience, 23(6), 2440 LP – 2452.

1573 https://doi.org/10.1523/JNEUROSCI.23-06-02440.2003

1574

1575 Kotsopoulos, I. A. W., Van Merode, T., Kessels, F. G. H., De Krom, M. C. T. F. M., & Knottnerus, J.

1576 A. (2002). Systematic Review and Meta-analysis of Incidence Studies of Epilepsy and Unprovoked

1577 Seizures. Epilepsia, 43(11), 1402–1409. JOUR. http://doi.org/10.1046/j.1528-1157.2002.t01-1-

1578 26901.x

1579

56

1580 Kovács, R., Raue, C., Gabriel, S., & Heinemann, U. (2011). Functional test of multidrug transporter

1581 activity in hippocampal-neocortical brain slices from epileptic patients. Journal of neuroscience

1582 methods, 200(2), 164–172. https://doi.org/10.1016/j.jneumeth.2011.06.032

1583

1584 Kralic, J.E., Ledergerber, D.A. and Fritschy, J.‐ M. (2005), Disruption of the neurogenic potential of

1585 the dentate gyrus in a mouse model of temporal lobe epilepsy with focal seizures. European Jour-

1586 nal of Neuroscience, 22: 1916-1927. doi:10.1111/j.1460-9568.2005.04386.x

1587

1588 Krishnan V. (2020). Depression and Anxiety in the Epilepsies: from Bench to Bedside. Current

1589 neurology and neuroscience reports, 20(9), 41. https://doi.org/10.1007/s11910-020-01065-z

1590

1591 Lado, F. A. (2006). Chronic Bilateral Stimulation of the Anterior Thalamus of Kainate-treated Rats

1592 Increases Seizure Frequency. Epilepsia, 47(1), 27–32. JOUR. http://doi.org/10.1111/j.1528-

1593 1167.2006.00366.x

1594

1595 Lancaster, B., & Wheal, H. V. (1982). A comparative histological and electrophysiological study of

1596 some neurotoxins in the rat hippocampus. Journal of Comparative Neurology, 211(2), 105–114.

1597 JOUR. http://doi.org/10.1002/cne.902110202

1598

1599 Lapinlampi, N., Melin, E., Aronica, E., Bankstahl, J. P., Becker, A., Bernard, C., Gorter, J. A.,

1600 Gröhn, O., Lipsanen, A., Lukasiuk, K., Löscher, W., Paananen, J., Ravizza, T., Roncon, P., Simo-

1601 nato, M., Vezzani, A., Kokaia, M., & Pitkänen, A. (2017). Common data elements and data man-

1602 agement: Remedy to cure underpowered preclinical studies. Epilepsy research, 129, 87–90.

1603 https://doi.org/10.1016/j.eplepsyres.2016.11.010

1604

1605 Lee, B. K., Lee, D. H., Park, S., Park, S. L., Yoon, J.-S., Lee, M. G., … Jung, Y.-S. (2009). Effects

1606 of KR-33028, a novel Na+/H+ exchanger-1 inhibitor, on glutamate-induced neuronal cell death and

1607 ischemia-induced cerebral infarct. Brain Research, 1248, 22–30. JOUR.

1608 https://doi.org/10.1016/j.brainres.2008.10.061 57

1609

1610 Lee, D. J., Hsu, M. S., Seldin, M. M., Arellano, J. L., & Binder, D. K. (2012). Decreased expression

1611 of the glial water channel aquaporin-4 in the intrahippocampal kainic acid model of epileptogene-

1612 sis. Experimental Neurology, 235(1), 246–255. JOUR.

1613 https://doi.org/10.1016/j.expneurol.2012.02.002

1614

1615 Lerma, J., Paternain, A. V, Rodríguez-Moreno, A., & López-García, J. C. (2001). Molecular Physi-

1616 ology of Kainate Receptors. Physiological Reviews, 81(3), 971–998. JOUR.

1617 http://doi.org/10.1152/physrev.2001.81.3.971

1618

1619 Lerma, J., & Marques, J. M. (2013). Kainate Receptors in Health and Disease. Neuron, 80(2),

1620 292–311. JOUR. https://doi.org/10.1016/j.neuron.2013.09.045

1621

1622 Lévesque, M., Langlois, J. M. P., Lema, P., Courtemanche, R., Bilodeau, G.-A., & Carmant, L.

1623 (2009). Synchronized gamma oscillations (30–50 Hz) in the amygdalo-hippocampal network in re-

1624 lation with seizure propagation and severity. Neurobiology of Disease, 35(2), 209–218. JOUR.

1625 https://doi.org/10.1016/j.nbd.2009.04.011

1626

1627 Lévesque, M., & Avoli, M. (2013). The kainic acid model of temporal lobe epilepsy. Neuroscience &

1628 Biobehavioral Reviews, 37(10, Part 2), 2887–2899. JOUR.

1629 https://doi.org/10.1016/j.neubiorev.2013.10.011

1630

1631 Lévesque, M., Salami, P., Gotman, J., & Avoli, M. (2012). Two seizure-onset types reveal specific

1632 patterns of high-frequency oscillations in a model of temporal lobe epilepsy. The Journal of Neuro-

1633 science : The Official Journal of the Society for Neuroscience, 32(38), 13264–13272. JOUR.

1634 http://doi.org/10.1523/JNEUROSCI.5086-11.2012

1635

1636

58

1637 Li, T., Lytle, N., Lan, J.-Q., Sandau, U. S., & Boison, D. (2012). Local disruption of glial adenosine

1638 homeostasis in mice associates with focal electrographic seizures: a first step in epileptogenesis?

1639 Glia, 60(1), 83–95. JOUR. http://doi.org/10.1002/glia.21250

1640

1641 Liu, G., Gu, B., He, X.-P., Joshi, R. B., Wackerle, H. D., Rodriguiz, R. M., … McNamara, J. O.

1642 (2013). Transient Inhibition of TrkB Kinase after Status Epilepticus Prevents Development of Tem-

1643 poral Lobe Epilepsy. Neuron, 79(1), 31–38. JOUR. https://doi.org/10.1016/j.neuron.2013.04.027

1644

1645 LÖSCHER, W., & EBERT, U. (1996). THE ROLE OF THE PIRIFORM CORTEX IN KINDLING.

1646 Progress in Neurobiology, 50(5), 427–481. JOUR. https://doi.org/10.1016/S0301-0082(96)00036-6

1647

1648 Löscher, W., Hirsch, L. J., & Schmidt, D. (2015). The enigma of the latent period in the develop-

1649 ment of symptomatic acquired epilepsy — Traditional view versus new concepts. Epilepsy & Be-

1650 havior, 52, 78–92. https://doi.org/10.1016/j.yebeh.2015.08.037

1651

1652 Lothman, E. W., Bertram, E. H., Bekenstein, J. W., & Perlin, J. B. (1989). Self-sustaining limbic

1653 status epilepticus induced by ‘continuous’ hippocampal stimulation: electrographic and behavioral

1654 characteristics. Epilepsy research, 3(2), 107-119.

1655

1656 Lowry, E. R., Kruyer, A., Norris, E. H., Cederroth, C. R., & Strickland, S. (2013). The GluK4 kainate

1657 receptor subunit regulates memory, mood, and excitotoxic neurodegeneration. Neuroscience, 235,

1658 215–225. https://doi.org/10.1016/j.neuroscience.2013.01.029

1659

1660 Lu, M.-O., Zhang, X.-M., Mix, E., Quezada, H. C., Jin, T., Zhu, J., & Adem, A. (2008). TNF-α recep-

1661 tor 1 deficiency enhances kainic acid–induced hippocampal injury in mice. Journal of Neuroscience

1662 Research, 86(7), 1608–1614. JOUR. http://doi.org/10.1002/jnr.21600

1663

59

1664 Luo, L., Jin, Y., Kim, I.-D., & Lee, J.-K. (2013). Glycyrrhizin Attenuates Kainic Acid-Induced Neu-

1665 ronal Cell Death in the Mouse Hippocampus. Exp Neurobiol, 22(2), 107–115. JOUR.

1666 http://synapse.koreamed.org/DOIx.php?id=10.5607%2Fen.2013.22.2.107

1667

1668 Maillard, L., Vignal, J. P., Gavaret, M., Guye, M., Biraben, A., McGonigal, A., Chauvel, P., & Barto-

1669 lomei, F. (2004). Semiologic and electrophysiologic correlations in temporal lobe seizure subtypes.

1670 Epilepsia, 45(12), 1590–1599. https://doi.org/10.1111/j.0013-9580.2004.09704.x

1671

1672 Malik, R., Dougherty, K. A., Parikh, K., Byrne, C., & Johnston, D. (2016). Mapping the electrophys-

1673 iological and morphological properties of CA1 pyramidal neurons along the longitudinal hippocam-

1674 pal axis. Hippocampus, 26(3), 341–361. https://doi.org/10.1002/hipo.22526

1675

1676 Manfredini, R., Vergine, G., Boari, B., Faggioli, R., & Borgna-Pignatti, C. (2004). Circadian and

1677 seasonal variation of first febrile seizures. The Journal of Pediatrics, 145(6), 838–839. JOUR.

1678 https://doi.org/10.1016/j.jpeds.2004.06.079

1679

1680 Manouze, H., Ghestem, A., Poillerat, V., Bennis, M., Ba-M’hamed, S., Benoliel, J. J., … Bernard,

1681 C. (2019). Effects of Single Cage Housing on Stress, Cognitive, and Seizure Parameters in the Rat

1682 and Mouse Pilocarpine Models of Epilepsy. Eneuro, 6(4), ENEURO.0179-18.2019. JOUR.

1683 http://doi.org/10.1523/ENEURO.0179-18.2019

1684

1685 Mathern, G. W., Babb, T. L., Vickrey, B. G., Melendez, M., & Pretorius, J. K. (1995). The clinical-

1686 pathogenic mechanisms of hippocampal neuron loss and surgical outcomes in temporal lobe epi-

1687 lepsy. Brain, 118(1), 105–118. JOUR. http://doi.org/10.1093/brain/118.1.105

1688

1689 Mazarati, A. M., Wasterlain, C. G., Sankar, R., & Shin, D. (1998). Self-sustaining status epilepticus

1690 after brief electrical stimulation of the perforant path. Brain Research, 801(1), 251–253.

1691 https://doi.org/10.1016/S0006-8993(98)00606-4

1692 60

1693 McCord, M. C., Lorenzana, A., Bloom, C. S., Chancer, Z. O., & Schauwecker, P. E. (2008). Effect

1694 of age on kainate-induced seizure severity and cell death. Neuroscience, 154(3), 1143–1153.

1695 JOUR. http://doi.org/10.1016/j.neuroscience.2008.03.082

1696

1697 McIntyre, D. C., Kelly, M. E., & Armstrong, J. N. (1993). Kindling in the perirhinal cortex. Brain Re-

1698 search, 615(1), 1–6. JOUR. https://doi.org/10.1016/0006-8993(93)91108-5

1699

1700 McKhann, G. M., Wenzel, H. J., Robbins, C. A., Sosunov, A. A., & Schwartzkroin, P. A. (2003).

1701 Mouse strain differences in kainic acid sensitivity, seizure behavior, mortality, and hippocampal

1702 pathology. Neuroscience, 122(2), 551–561. JOUR. https://doi.org/10.1016/S0306-4522(03)00562-

1703 1

1704

1705 McLin, J. P., & Steward, O. (2006). Comparison of seizure phenotype and neurodegeneration in-

1706 duced by systemic kainic acid in inbred, outbred, and hybrid mouse strains. European Journal of

1707 Neuroscience, 24(8), 2191–2202. JOUR. http://doi.org/10.1111/j.1460-9568.2006.05111.x

1708

1709 Medel-Matus, J. S., Shin, D., Sankar, R., & Mazarati, A. (2017). Inherent vulnerabilities in mono-

1710 aminergic pathways predict the emergence of depressive impairments in an animal model of

1711 chronic epilepsy. Epilepsia, 58(8), e116–e121. https://doi.org/10.1111/epi.13822

1712

1713

1714 Mikati, M. A., Werner, S., Shehab, L., Shamseddine, A., Simaan, E., Liu, Z., … Holmes, G. (2003).

1715 Stages of status epilepticus in the developing brain. Epilepsy Research, 55(1), 9–19. JOUR.

1716 https://doi.org/10.1016/S0920-1211(03)00089-5

1717

1718 Miyamoto, M., Ishida, M., & Shinozaki, H. (1997). Anticonvulsive and neuroprotective actions of a

1719 potent agonist (DCG-IV) for Group II metabotropic glutamate receptors against intraventricular

1720 kainate in the rat. Neuroscience, 77(1), 131–140. JOUR. https://doi.org/10.1016/S0306-

1721 4522(96)00442-3 61

1722

1723 Mizuseki, K., Diba, K., Pastalkova, E., & Buzsáki, G. (2011). Hippocampal CA1 pyramidal cells

1724 form functionally distinct sublayers. Nature neuroscience, 14(9), 1174–1181.

1725 https://doi.org/10.1038/nn.2894

1726

1727 Morimoto, K., Fahnestock, M., & Racine, R. J. (2004). Kindling and status epilepticus models of

1728 epilepsy: rewiring the brain. Progress in Neurobiology, 73(1), 1–60. JOUR.

1729 http://doi.org/https://doi.org/10.1016/j.pneurobio.2004.03.009

1730

1731 Mouri, G., Jimenez-Mateos, E., Engel, T., Dunleavy, M., Hatazaki, S., Paucard, A., … Henshall, D.

1732 C. (2008). Unilateral hippocampal CA3-predominant damage and short latency epileptogenesis

1733 after intra-amygdala microinjection of kainic acid in mice. Brain Research, 1213, 140–151. JOUR.

1734 https://doi.org/10.1016/j.brainres.2008.03.061

1735

1736 Mulle, C., Sailer, A., Pérez-Otaño, I., Dickinson-Anson, H., Castillo, P. E., Bureau, I., … Heine-

1737 mann, S. F. (1998). Altered synaptic physiology and reduced susceptibility to kainate-induced sei-

1738 zures in GluR6-deficient mice. Nature, 392(6676), 601–605. JOUR. http://doi.org/10.1038/33408

1739

1740 Muñoz, A., Ballesteros‐ Yáñez, I., DeFelipe, J., Arellano, J. I., & Sola, R. G. (2004). Histopathology

1741 and reorganization of chandelier cells in the human epileptic sclerotic hippocampus. Brain, 127(1),

1742 45–64. JOUR. http://doi.org/10.1093/brain/awh004

1743

1744 Murphy, T. H., Miyamoto, M., Sastre, A., Schnaar, R. L., & Coyle, J. T. (1989). Glutamate toxicity

1745 in a neuronal cell line involves inhibition of cystine transport leading to oxidative stress. Neuron,

1746 2(6), 1547–1558. JOUR. https://doi.org/10.1016/0896-6273(89)90043-3

1747

1748 NADLER, J. V., PERRY, B. W., & COTMAN, C. W. (1978). Intraventricular kainic acid preferential-

1749 ly destroys hippocampal pyramidal cells. Nature, 271(5646), 676–677. JOUR.

1750 http://doi.org/10.1038/271676a0 62

1751

1752 Nguyen, D., Alavi, M. V, Kim, K.-Y., Kang, T., Scott, R. T., Noh, Y. H., … Ju, W.-K. (2011). A new

1753 vicious cycle involving glutamate excitotoxicity, oxidative stress and mitochondrial dynamics. Cell

1754 Death & Disease, 2(12), e240–e240. JOUR. http://doi.org/10.1038/cddis.2011.117

1755

1756 Ogren, J. A., Wilson, C. L., Bragin, A., Lin, J. J., Salamon, N., Dutton, R. A., … Staba, R. J. (2009).

1757 Three-dimensional surface maps link local atrophy and fast ripples in human epileptic hippocam-

1758 pus. Annals of Neurology, 66(6), 783–791. JOUR. http://doi.org/10.1002/ana.21703

1759

1760 Palizvan, M. R., Fathollahi, Y., Semnanian, S., Hajezadeh, S., & Mirnajafizadh, J. (2001). Differen-

1761 tial effects of pentylenetetrazol-kindling on long-term potentiation of population excitatory postsyn-

1762 aptic potentials and population spikes in the CA1 region of rat hippocampus. Brain Research,

1763 898(1), 82–90. JOUR. https://doi.org/10.1016/S0006-8993(01)02146-1

1764

1765 Pallud, J., Häussler, U., Langlois, M., Hamelin, S., Devaux, B., Deransart, C., & Depaulis, A.

1766 (2011). Dentate gyrus and hilus transection blocks seizure propagation and granule cell dispersion

1767 in a mouse model for mesial temporal lobe epilepsy. Hippocampus, 21(3), 334–343. JOUR.

1768 http://doi.org/10.1002/hipo.20795

1769

1770 Pandis, C., Sotiriou, E., Kouvaras, E., Asprodini, E., Papatheodoropoulos, C., & Angelatou, F.

1771 (2006). Differential expression of NMDA and AMPA receptor subunits in rat dorsal and ventral hip-

1772 pocampus. Neuroscience, 140(1), 163–175. https://doi.org/10.1016/j.neuroscience.2006.02.003

1773

1774 Parent, J. M. (2007). Adult neurogenesis in the intact and epileptic dentate gyrus. In H. E. B. T.-P.

1775 in B. R. Scharfman (Ed.), The Dentate Gyrus: A Comprehensive Guide to Structure, Function, and

1776 Clinical Implications (Vol. 163, pp. 529–817). Elsevier. https://doi.org/10.1016/S0079-

1777 6123(07)63028-3

1778

63

1779 Parent, J. M., & Kron, M. M. (2012). Neurogenesis and epilepsy. In Jasper's Basic Mechanisms of

1780 the Epilepsies [Internet]. 4th edition. National Center for Biotechnology Information (US).

1781

1782 Park, H.J., Kim, H.J., Park, H.J., Ra, J., Zheng, L.T., Yim, S.V. and Chung, J.‐ H. (2008), Protec-

1783 tive effect of topiramate on kainic acid–induced cell death in mice hippocampus. Epilepsia, 49:

1784 163-167. doi:10.1111/j.1528-1167.2007.01308.x

1785

1786 Patel, S., Meldrum, B. S., & Collins, J. F. (1986). Distribution of [3H]kainic acid and binding sites in

1787 the rat brain: in vivo and in vitro receptor autoradiography. Neuroscience Letters, 70(3), 301–307.

1788 JOUR. https://doi.org/10.1016/0304-3940(86)90569-0

1789

1790 Paternain, A. V, Herrera, M. T., Nieto, M. A., & Lerma, J. (2000). GluR5 and GluR6 Kainate Recep-

1791 tor Subunits Coexist in Hippocampal Neurons and Coassemble to Form Functional Receptors. The

1792 Journal of Neuroscience, 20(1), 196 LP – 205. https://doi.org/10.1523/JNEUROSCI.20-01-

1793 00196.2000

1794

1795 Pernot, F., Heinrich, C., Barbier, L., Peinnequin, A., Carpentier, P., Dhote, F., … Dorandeu, F.

1796 (2011). Inflammatory changes during epileptogenesis and spontaneous seizures in a mouse model

1797 of mesiotemporal lobe epilepsy. Epilepsia, 52(12), 2315–2325. JOUR.

1798 http://doi.org/10.1111/j.1528-1167.2011.03273.x

1799

1800 Perucca, P., Dubeau, F., & Gotman, J. (2014). Intracranial electroencephalographic seizure-onset

1801 patterns: effect of underlying pathology. Brain : a journal of neurology, 137(Pt 1), 183–196.

1802 https://doi.org/10.1093/brain/awt299

1803

1804 Pinheiro, P. S., Lanore, F., Veran, J., Artinian, J., Blanchet, C., Crépel, V., Perrais, D., & Mulle, C.

1805 (2013). Selective Block of Postsynaptic Kainate Receptors Reveals Their Function at Hippocampal

1806 Mossy Fiber Synapses. Cerebral Cortex, 23(2), 323–331. https://doi.org/10.1093/cercor/bhs022

1807 64

1808 Pitkänen, A., & Mcintosh, T. K. (2006). Animal Models of Post-Traumatic Epilepsy. Journal of Neu-

1809 rotrauma, 23(2), 241–261. JOUR. http://doi.org/10.1089/neu.2006.23.241

1810

1811 Pitkänen, A., & Immonen, R. (2014). Epilepsy Related to Traumatic Brain Injury. Neurotherapeu-

1812 tics, 11(2), 286–296. JOUR. http://doi.org/10.1007/s13311-014-0260-7

1813

1814 Pitsch, J., Kuehn, J. C., Gnatkovsky, V., Müller, J. A., van Loo, K. M. J., de Curtis, M., … Becker,

1815 A. J. (2019). Anti-epileptogenic and Anti-convulsive Effects of Fingolimod in Experimental Tem-

1816 poral Lobe Epilepsy. Molecular Neurobiology, 56(3), 1825–1840. JOUR.

1817 http://doi.org/10.1007/s12035-018-1181-y

1818

1819 Puttachary, S., Sharma, S., Tse, K., Beamer, E., Sexton, A., Crutison, J., & Thippeswamy, T.

1820 (2015). Immediate Epileptogenesis after Kainate-Induced Status Epilepticus in C57BL/6J Mice:

1821 Evidence from Long Term Continuous Video-EEG Telemetry. PLOS ONE, 10(7), e0131705.

1822 JOUR. https://doi.org/10.1371/journal.pone.0131705

1823

1824 Quigg, M. (2000). Circadian rhythms: interactions with seizures and epilepsy. Epilepsy Research,

1825 42(1), 43–55. JOUR. https://doi.org/10.1016/S0920-1211(00)00157-1

1826

1827 Racine, R. J. (1972). Modification of seizure activity by electrical stimulation: II. Motor seizure.

1828 Electroencephalography and Clinical Neurophysiology, 32(3), 281–294. JOUR.

1829 https://doi.org/10.1016/0013-4694(72)90177-0

1830

1831 Raedt, R., Van Dycke, A., Van Melkebeke, D., De Smedt, T., Claeys, P., Wyckhuys, T., … Boon,

1832 P. (2009). Seizures in the intrahippocampal kainic acid epilepsy model: characterization using

1833 long-term video-EEG monitoring in the rat. Acta Neurologica Scandinavica, 119(5), 293–303.

1834 JOUR. http://doi.org/10.1111/j.1600-0404.2008.01108.x

1835

65

1836 Rakhade, S. N., Klein, P. M., Huynh, T., Hilario-Gomez, C., Kosaras, B., Rotenberg, A., & Jensen,

1837 F. E. (2011). Development of later life spontaneous seizures in a rodent model of hypoxia-induced

1838 neonatal seizures. Epilepsia, 52(4), 753–765. JOUR. http://doi.org/10.1111/j.1528-

1839 1167.2011.02992.x

1840

1841 Rao, M. S., Hattiangady, B., Reddy, D. S., & Shetty, A. K. (2006). Hippocampal neurodegenera-

1842 tion, spontaneous seizures, and mossy fiber sprouting in the F344 rat model of temporal lobe epi-

1843 lepsy. Journal of Neuroscience Research, 83(6), 1088–1105. JOUR.

1844 http://doi.org/10.1002/jnr.20802

1845

1846 Rapp, P. R., Stack, E. C., & Gallagher, M. (1999). Morphometric studies of the aged hippocampus:

1847 I. Volumetric analysis in behaviorally characterized rats. Journal of Comparative Neurology,

1848 403(4), 459–470. JOUR. http://doi.org/10.1002/(SICI)1096-9861(19990125)403:4<459::AID-

1849 CNE3>3.0.CO;2-9

1850

1851 Rattka, M., Brandt, C., & Löscher, W. (2013). The intrahippocampal kainate model of temporal lobe

1852 epilepsy revisited: Epileptogenesis, behavioral and cognitive alterations, pharmacological re-

1853 sponse, and hippoccampal damage in epileptic rats. Epilepsy Research, 103(2), 135–152. JOUR.

1854 https://doi.org/10.1016/j.eplepsyres.2012.09.015

1855

1856 Riban, V., Bouilleret, V., Pham-Lê, B. T., Fritschy, J.-M., Marescaux, C., & Depaulis, A. (2002).

1857 Evolution of hippocampal epileptic activity during the development of hippocampal sclerosis in a

1858 mouse model of temporal lobe epilepsy. Neuroscience, 112(1), 101–111. JOUR.

1859 https://doi.org/10.1016/S0306-4522(02)00064-7

1860

1861 ROGAWSKI, M. A., GRYDER, D., CASTANEDA, D., YONEKAWA, W., BANKS, M. K., & LI, H. E.

1862 (2003). GluR5 Kainate Receptors, Seizures, and the Amygdala. Annals of the New York Academy

1863 of Sciences, 985(1), 150–162. JOUR. http://doi.org/10.1111/j.1749-6632.2003.tb07079.x

1864 66

1865 Roux, E., & Borrel, A. (1898). Tétanos cérébral et immunité contre le tétanos. Ann. Inst. Pasteur,

1866 12, 225–239. JOUR.

1867

1868 Rowley, S., & Patel, M. (2013). Mitochondrial involvement and oxidative stress in temporal lobe

1869 epilepsy. Free Radical Biology & Medicine, 62, 121–131. JOUR.

1870 http://doi.org/10.1016/j.freeradbiomed.2013.02.002

1871

1872 Royle, S. J., Collins, F. C., Rupniak, H. T., Barnes, J. C., & Anderson, R. (1999). Behavioural anal-

1873 ysis and susceptibility to CNS injury of four inbred strains of mice. Brain Research, 816(2), 337–

1874 349. JOUR. https://doi.org/10.1016/S0006-8993(98)01122-6

1875

1876 Sabilallah, M., Fontanaud, P., Linck, N., Boussadia, B., Peyroutou, R., Lasgouzes, T., … Hirbec,

1877 H. E. (2016). Evidence for Status Epilepticus and Pro-Inflammatory Changes after Intranasal

1878 Kainic Acid Administration in Mice. PLOS ONE, 11(3), e0150793. JOUR.

1879 https://doi.org/10.1371/journal.pone.0150793

1880

1881 Saggio, M. L., Crisp, D., Scott, J. M., Karoly, P., Kuhlmann, L., Nakatani, M., … Stacey, W. C.

1882 (2020). A taxonomy of seizure dynamotypes. eLife, 9, e55632. JOUR.

1883 http://doi.org/10.7554/eLife.55632

1884

1885 Sánchez Fernández, I., Gaínza-Lein, M., Abend, N. S., Amengual-Gual, M., Anderson, A., Arya,

1886 R., Brenton, J. N., Carpenter, J. L., Chapman, K. E., Clark, J., Farias-Moeller, R., Davis Gaillard,

1887 W., Glauser, T. A., Goldstein, J., Goodkin, H. P., Guerriero, R. M., Hecox, K., Jackson, M., Kapur,

1888 K., Kelley, S. A., … Pediatric Status Epilepticus Research Group (2019). The onset of pediatric re-

1889 fractory status epilepticus is not distributed uniformly during the day. Seizure, 70, 90–96.

1890 https://doi.org/10.1016/j.seizure.2019.06.017

1891

1892 Sandow, N., Kim, S., Raue, C., Päsler, D., Klaft, Z. J., Antonio, L. L., Hollnagel, J. O., Kovacs, R.,

1893 Kann, O., Horn, P., Vajkoczy, P., Holtkamp, M., Meencke, H. J., Cavalheiro, E. A., Pragst, F., Ga- 67

1894 briel, S., Lehmann, T. N., & Heinemann, U. (2015). Drug resistance in cortical and hippocampal

1895 slices from resected tissue of epilepsy patients: no significant impact of p-glycoprotein and multi-

1896 drug resistance-associated proteins. Frontiers in neurology, 6, 30.

1897 https://doi.org/10.3389/fneur.2015.00030

1898

1899 Scharfman, H. E., Mercurio, T. C., Goodman, J. H., Wilson, M. A., & MacLusky, N. J. (2003). Hip-

1900 pocampal excitability increases during the estrous cycle in the rat: a potential role for brain-derived

1901 neurotrophic factor. The Journal of Neuroscience : The Official Journal of the Society for Neurosci-

1902 ence, 23(37), 11641–11652.

1903

1904 Scharfman, H. E., Hintz, T. M., Gomez, J., Stormes, K. A., Barouk, S., Malthankar-Phatak, G. H.,

1905 … Maclusky, N. J. (2007). Changes in hippocampal function of ovariectomized rats after sequential

1906 low doses of estradiol to simulate the preovulatory estrogen surge. The European Journal of Neu-

1907 roscience, 26(9), 2595–2612. http://doi.org/10.1111/j.1460-9568.2007.05848.x

1908

1909 Schauwecker P. E. (2012). Strain differences in seizure-induced cell death following pilocarpine-

1910 induced status epilepticus. Neurobiology of disease, 45(1), 297–304.

1911 https://doi.org/10.1016/j.nbd.2011.08.013

1912

1913 Schauwecker, P. E., & Steward, O. (1997). Genetic determinants of susceptibility to excitotoxic cell

1914 death: implications for gene targeting approaches. Proceedings of the National Academy of Sci-

1915 ences of the United States of America, 94(8), 4103–4108. JOUR. Retrieved from

1916 https://www.ncbi.nlm.nih.gov/pubmed/9108112

1917

1918 Schwob, J. E., Fuller, T., Price, J. L., & Olney, J. W. (1980). Widespread patterns of neuronal

1919 damage following systemic or intracerebral injections of kainic acid: A histological study. Neurosci-

1920 ence, 5(6), 991–1014. JOUR. https://doi.org/10.1016/0306-4522(80)90181-5

68

1921 Sharif, F., Tayebi, B., Buzsáki, G., Royer, S., & Fernandez-Ruiz, A. (2021). Subcircuits of Deep

1922 and Superficial CA1 Place Cells Support Efficient Spatial Coding across Heterogeneous Environ-

1923 ments. Neuron. https://doi.org/10.1016/j.neuron.2020.10.034

1924

1925 Sharma, A. K., Reams, R. Y., Jordan, W. H., Miller, M. A., Thacker, H. L., & Snyder, P. W. (2007).

1926 Mesial Temporal Lobe Epilepsy: Pathogenesis, Induced Rodent Models and Lesions. Toxicologic

1927 Pathology, 35(7), 984–999. JOUR. http://doi.org/10.1080/01926230701748305

1928

1929 Sharma, A. K., Jordan, W. H., Reams, R. Y., Hall, D. G., & Snyder, P. W. (2008). Temporal Profile

1930 of Clinical Signs and Histopathologic Changes in an F-344 Rat Model of Kainic Acid–induced Me-

1931 sial Temporal Lobe Epilepsy. Toxicologic Pathology, 36(7), 932–943. JOUR.

1932 http://doi.org/10.1177/0192623308326093

1933

1934 Sharma, S., Puttachary, S., Thippeswamy, A., Kanthasamy, A. G., & Thippeswamy, T. (2018). Sta-

1935 tus Epilepticus: Behavioral and Electroencephalography Seizure Correlates in Kainate Experi-

1936 mental Models. Frontiers in Neurology, 9, 7. https://doi.org/10.3389/fneur.2018.00007

1937

1938 Sheybani, L., Birot, G., Contestabile, A., Seeck, M., Kiss, J. Z., Schaller, K., Michel, C. M., & Quai-

1939 riaux, C. (2018). Electrophysiological Evidence for the Development of a Self-Sustained Large-

1940 Scale Epileptic Network in the Kainate Mouse Model of Temporal Lobe Epilepsy. The Journal of

1941 Neuroscience, 38(15), 3776 LP – 3791. https://doi.org/10.1523/JNEUROSCI.2193-17.2018

1942

1943 Shinoda, S., Araki, T., Lan, J.-Q., Schindler, C. K., Simon, R. P., Taki, W., & Henshall, D. C.

1944 (2004). Development of a model of seizure-induced hippocampal injury with features of pro-

1945 grammed cell death in the BALB/c mouse. Journal of Neuroscience Research, 76(1), 121–128.

1946 JOUR. http://doi.org/10.1002/jnr.20064

1947

1948 Skucas, V. A., Duffy, A. M., Harte-Hargrove, L. C., Magagna-Poveda, A., Radman, T.,

1949 Chakraborty, G., … Scharfman, H. E. (2013). Testosterone depletion in adult male rats increases 69

1950 mossy fiber transmission, LTP, and sprouting in area CA3 of hippocampus. The Journal of Neuro-

1951 science : The Official Journal of the Society for Neuroscience, 33(6), 2338–2355. JOUR.

1952 http://doi.org/10.1523/JNEUROSCI.3857-12.2013

1953

1954 Sloviter R. S. (1987). Decreased hippocampal inhibition and a selective loss of interneurons in ex-

1955 perimental epilepsy. Science (New York, N.Y.), 235(4784), 73–76.

1956 https://doi.org/10.1126/science.2879352

1957

1958 Sloviter, R. S., Sollas, A. L., Barbaro, N. M., & Laxer, K. D. (1991). Calcium-binding protein (cal-

1959 bindin-D28K) and parvalbumin immunocytochemistry in the normal and epileptic human hippo-

1960 campus. Journal of Comparative Neurology, 308(3), 381–396. JOUR.

1961 http://doi.org/10.1002/cne.903080306

1962

1963 Słowiński, P., Sheybani, L., Michel, C. M., Richardson, M. P., Quairiaux, C., Terry, J. R., & Good-

1964 fellow, M. (2019). Background EEG Connectivity Captures the Time-Course of Epileptogenesis in

1965 a Mouse Model of Epilepsy. Eneuro, 6(4), ENEURO.0059-19.2019.

1966 https://doi.org/10.1523/ENEURO.0059-19.2019

1967

1968 Sommer, C., Roth, S. U., & Kiessling, M. (2001). Kainate-induced epilepsy alters protein expres-

1969 sion of AMPA receptor subunits GluR1, GluR2 and AMPA receptor binding protein in the rat hippo-

1970 campus. Acta Neuropathologica, 101(5), 460–468. JOUR. http://doi.org/10.1007/s004010000310

1971

1972 Song, C., Xu, W., Zhang, X. et al. (2016). CXCR4 Antagonist AMD3100 Suppresses the Long-

1973 Term Abnormal Structural Changes of Newborn Neurons in the Intraventricular Kainic Acid Model

1974 of Epilepsy. Mol Neurobiol, 53: 1518. https://doi.org/10.1007/s12035-015-9102-9

1975

1976 Sotiriou, E., Papatheodoropoulos, C., & Angelatou, F. (2005). Differential expression of gamma-

1977 aminobutyric acid--a receptor subunits in rat dorsal and ventral hippocampus. Journal of neurosci-

1978 ence research, 82(5), 690–700. https://doi.org/10.1002/jnr.20670 70

1979

1980 Spencer, D. C., Sun, F. T., Brown, S. N., Jobst, B. C., Fountain, N. B., Wong, V. S. S., … Quigg,

1981 M. (2016). Circadian and ultradian patterns of epileptiform discharges differ by seizure-onset loca-

1982 tion during long-term ambulatory intracranial monitoring. Epilepsia, 57(9), 1495–1502. JOUR.

1983 http://doi.org/10.1111/epi.13455

1984

1985 Sperk, G., Lassmann, H., Baran, H., Kish, S. J., Seitelberger, F., & Hornykiewicz, O. (1983). Kainic

1986 acid induced seizures: Neurochemical and histopathological changes. Neuroscience, 10(4), 1301–

1987 1315. JOUR. https://doi.org/10.1016/0306-4522(83)90113-6

1988

1989 Stafstrom, C. E., Thompson, J. L., & Holmes, G. L. (1992). Kainic acid seizures in the developing

1990 brain: status epilepticus and spontaneous recurrent seizures. Developmental Brain Research,

1991 65(2), 227–236. JOUR. https://doi.org/10.1016/0165-3806(92)90184-X

1992

1993 Stanley, D. P., & Shetty, A. K. (2004). Aging in the rat hippocampus is associated with widespread

1994 reductions in the number of glutamate decarboxylase-67 positive interneurons but not interneuron

1995 degeneration. Journal of Neurochemistry, 89(1), 204–216. JOUR. http://doi.org/10.1111/j.1471-

1996 4159.2004.02318.x

1997

1998 Suárez, L. M., Cid, E., Gal, B., Inostroza, M., Brotons-Mas, J. R., Gómez-Domínguez, D., … Solís,

1999 J. M. (2012). Systemic Injection of Kainic Acid Differently Affects LTP Magnitude Depending on its

2000 Epileptogenic Efficiency. PLOS ONE, 7(10), e48128. JOUR.

2001 https://doi.org/10.1371/journal.pone.0048128

2002

2003 Sutula, T., Cascino, G., Cavazos, J., Parada, I., & Ramirez, L. (1989). Mossy fiber synaptic reor-

2004 ganization in the epileptic human temporal lobe. Annals of Neurology, 26(3), 321–330. JOUR.

2005 http://doi.org/10.1002/ana.410260303

2006

71

2007 Tabori, N. E., Stewart, L. S., Znamensky, V., Romeo, R. D., Alves, S. E., McEwen, B. S., & Milner,

2008 T. A. (2005). Ultrastructural evidence that androgen receptors are located at extranuclear sites in

2009 the rat hippocampal formation. Neuroscience, 130(1), 151–163. JOUR.

2010 https://doi.org/10.1016/j.neuroscience.2004.08.048

2011

2012 Tanaka, T., Tanaka, S., Fujita, T., Takano, K., Fukuda, H., Sako, K., & Yonemasu, Y. (1992). Ex-

2013 perimental complex partial seizures induced by a microinjection of kainic acid into limbic structures.

2014 Progress in Neurobiology, 38(3), 317–334. JOUR. https://doi.org/10.1016/0301-0082(92)90023-8

2015

2016 Tanaka, K., Jimenez-Mateos, E. M., Matsushima, S., Taki, W., & Henshall, D. C. (2010). Hippo-

2017 campal damage after intra-amygdala kainic acid-induced status epilepticus and seizure precondi-

2018 tioning-mediated neuroprotection in SJL mice. Epilepsy Research, 88(2), 151–161. JOUR.

2019 https://doi.org/10.1016/j.eplepsyres.2009.10.012

2020

2021 Tauck, D. L., & Nadler, J. V. (1985). Evidence of functional mossy fiber sprouting in hippocampal

2022 formation of kainic acid-treated rats. The Journal of Neuroscience, 5(4), 1016 LP-1022. JOUR.

2023 http://doi.org/10.1523/JNEUROSCI.05-04-01016.1985

2024

2025 Telfeian, A. E., Federoff, H. J., Leone, P., During, M. J., & Williamson, A. (2000). Overexpression

2026 of GluR6 in Rat Hippocampus Produces Seizures and Spontaneous Nonsynaptic Bursting in Vitro.

2027 Neurobiology of Disease, 7(4), 362–374. JOUR. https://doi.org/10.1006/nbdi.2000.0294

2028

2029 Thom M. (2014). Review: Hippocampal sclerosis in epilepsy: a neuropathology review. Neuropa-

2030 thology and applied neurobiology, 40(5), 520–543. https://doi.org/10.1111/nan.12150

2031

2032 Thompson, C. L., Pathak, S. D., Jeromin, A., Ng, L. L., MacPherson, C. R., Mortrud, M. T., Cusick,

2033 A., Riley, Z. L., Sunkin, S. M., Bernard, A., Puchalski, R. B., Gage, F. H., Jones, A. R., Bajic, V. B.,

2034 Hawrylycz, M. J., & Lein, E. S. (2008). Genomic anatomy of the hippocampus. Neuron, 60(6),

2035 1010–1021. https://doi.org/10.1016/j.neuron.2008.12.008 72

2036

2037 Torres-Reveron, A., Khalid, S., Williams, T. J., Waters, E. M., Jacome, L., Luine, V. N., … Milner,

2038 T. A. (2009). Hippocampal dynorphin immunoreactivity increases in response to gonadal steroids

2039 and is positioned for direct modulation by ovarian steroid receptors. Neuroscience, 159(1), 204–

2040 216. JOUR. http://doi.org/10.1016/j.neuroscience.2008.12.023

2041

2042 Tóth, K., Eross, L., Vajda, J., Halász, P., Freund, T. F., & Maglóczky, Z. (2010). Loss and reorgan-

2043 ization of calretinin-containing interneurons in the epileptic human hippocampus. Brain : A Journal

2044 of Neurology, 133(9), 2763–2777. JOUR. http://doi.org/10.1093/brain/awq149

2045

2046 Toth, M., Grimsby, J., Buzsaki, G., & Donovan, G. P. (1995). Epileptic seizures caused by inactiva-

2047 tion of a novel gene, jerky, related to centromere binding protein–B in transgenic mice. Nature Ge-

2048 netics, 11(1), 71–75. https://doi.org/10.1038/ng0995-71

2049

2050 Toyoda, I., Fujita, S., Thamattoor, A. K., & Buckmaster, P. S. (2015). Unit Activity of Hippocampal

2051 Interneurons before Spontaneous Seizures in an Animal Model of Temporal Lobe Epilepsy. The

2052 Journal of Neuroscience, 35(16), 6600 LP-6618. JOUR. http://doi.org/10.1523/JNEUROSCI.4786-

2053 14.2015

2054

2055 Trinka, E., Cock, H., Hesdorffer, D., Rossetti, A. O., Scheffer, I. E., Shinnar, S., … Lowenstein, D.

2056 H. (2015). A definition and classification of status epilepticus – Report of the ILAE Task Force on

2057 Classification of Status Epilepticus. Epilepsia, 56(10), 1515–1523. JOUR.

2058 http://doi.org/10.1111/epi.13121

2059

2060 Tse, K., Puttachary, S., Beamer, E., Sills, G. J., & Thippeswamy, T. (2014). Advantages of repeat-

2061 ed low dose against single high dose of kainate in C57BL/6J mouse model of status epilepticus:

2062 behavioral and electroencephalographic studies. PloS One, 9(5), e96622–e96622. JOUR.

2063 http://doi.org/10.1371/journal.pone.0096622

2064 73

2065 Turski, W. A., Cavalheiro, E. A., Schwarz, M., Czuczwar, S. J., Kleinrok, Z., & Turski, L. (1983).

2066 Limbic seizures produced by pilocarpine in rats: Behavioural, electroencephalographic and neuro-

2067 pathological study. Behavioural Brain Research, 9(3), 315–335. JOUR.

2068 http://doi.org/https://doi.org/10.1016/0166-4328(83)90136-5

2069

2070 Twele, F., Bankstahl, M., Klein, S., Römermann, K., & Löscher, W. (2015). The AMPA receptor

2071 antagonist NBQX exerts anti-seizure but not antiepileptogenic effects in the intrahippocampal

2072 kainate mouse model of mesial temporal lobe epilepsy. Neuropharmacology, 95, 234–242. JOUR.

2073 https://doi.org/10.1016/j.neuropharm.2015.03.014

2074

2075 Ullal, G., Fahnestock, M., & Racine, R. (2005). Time-dependent Effect of Kainate-induced Seizures

2076 on Glutamate Receptor GluR5, GluR6, and GluR7 mRNA and Protein Expression in Rat Hippo-

2077 campus. Epilepsia, 46(5), 616–623. JOUR. http://doi.org/10.1111/j.1528-1167.2005.49604.x

2078

2079 Umpierre, A. D., Bennett, I. V, Nebeker, L. D., Newell, T. G., Tian, B. B., Thomson, K. E., … Wil-

2080 cox, K. S. (2016). Repeated low-dose kainate administration in C57BL/6J mice produces temporal

2081 lobe epilepsy pathology but infrequent spontaneous seizures. Experimental Neurology, 279, 116–

2082 126. JOUR. http://doi.org/10.1016/j.expneurol.2016.02.014

2083

2084 Van Nieuwenhuyse, B., Raedt, R., Sprengers, M., Dauwe, I., Gadeyne, S., Carrette, E., … Vonck,

2085 K. (2015). The systemic kainic acid rat model of temporal lobe epilepsy: Long-term EEG monitor-

2086 ing. Brain Research, 1627, 1–11. JOUR. https://doi.org/10.1016/j.brainres.2015.08.016

2087

2088 Velasco, A. L., Wilson, C. L., Babb, T. L., & Engel Jr, J. (2000). Functional and anatomic corre-

2089 lates of two frequently observed temporal lobe seizure-onset patterns. Neural Plasticity, 7(1–2),

2090 49–63. JOUR. http://doi.org/10.1155/NP.2000.49

2091

74

2092 Vezzani, A., & Sperk, G. (2004). Overexpression of NPY and Y2 receptors in epileptic brain tissue:

2093 an endogenous neuroprotective mechanism in temporal lobe epilepsy? Neuropeptides, 38(4),

2094 245–252. JOUR. https://doi.org/10.1016/j.npep.2004.05.004

2095

2096 Wada, J. A., Osawa, T., & Mizoguchi, T. (1975). Recurrent spontaneous seizure state induced by

2097 prefrontal kindling in senegalese baboons, Papio papio. The Canadian journal of neurological sci-

2098 ences. Le journal canadien des sciences neurologiques, 2(4), 477–492.

2099 https://doi.org/10.1017/s031716710002062x

2100

2101 Wasterlain, C.G., Fujikawa, D.G., Penix, L. and Sankar, R. (1993), Pathophysiological Mecha-

2102 nisms of Brain Damage from Status Epilepticus. Epilepsia, 34: S37-S53. doi:10.1111/j.1528-

2103 1157.1993.tb05905.x

2104

2105 Wenk, G. L., & Barnes, C. A. (2000). Regional changes in the hippocampal density of AMPA and

2106 NMDA receptors across the lifespan of the rat. Brain Research, 885(1), 1–5. JOUR.

2107 https://doi.org/10.1016/S0006-8993(00)02792-X

2108

2109 White, A., Williams, P. A., Hellier, J. L., Clark, S., Dudek, F. E., & Staley, K. J. (2010). EEG spike

2110 activity precedes epilepsy after kainate-induced status epilepticus. Epilepsia, 51(3), 371–383.

2111 JOUR. http://doi.org/10.1111/j.1528-1167.2009.02339.x

2112

2113 Williams, P., White, A., Ferraro, D., Clark, S., Staley, K., & Dudek, F. E. (2006). The use of radiote-

2114 lemetry to evaluate electrographic seizures in rats with kainate-induced epilepsy. Journal of Neu-

2115 roscience Methods, 155(1), 39–48. JOUR. https://doi.org/10.1016/j.jneumeth.2005.12.035

2116

2117 Williams, P. A., White, A. M., Clark, S., Ferraro, D. J., Swiercz, W., Staley, K. J., & Dudek, F. E.

2118 (2009). Development of Spontaneous Recurrent Seizures after Kainate-Induced Status Epilepti-

2119 cus. The Journal of Neuroscience, 29(7), 2103 LP-2112. JOUR.

2120 http://doi.org/10.1523/JNEUROSCI.0980-08.2009 75

2121

2122 Wisden, W., & Seeburg, P. H. (1993). A complex mosaic of high-affinity kainate receptors in rat

2123 brain. The Journal of Neuroscience, 13(8), 3582 LP-3598. JOUR.

2124 http://doi.org/10.1523/JNEUROSCI.13-08-03582.1993

2125

2126 Wozniak, D. F., Stewart, G. R., Philip Miller, J., & Olney, J. W. (1991). Age-related sensitivity to

2127 kainate neurotoxicity. Experimental Neurology, 114(2), 250–253. JOUR.

2128 https://doi.org/10.1016/0014-4886(91)90042-B

2129

2130 Zeidler, Z., Brandt-Fontaine, M., Leintz, C., Krook-Magnuson, C., Netoff, T., & Krook-Magnuson, E.

2131 (2018). Targeting the Mouse Ventral Hippocampus in the Intrahippocampal Kainic Acid Model of

2132 Temporal Lobe Epilepsy. eNeuro, 5(4), ENEURO.0158-18.2018. JOUR.

2133 http://doi.org/10.1523/ENEURO.0158-18.2018

2134

2135 Zeng, L.-H., Rensing, N. R., Zhang, B., Gutmann, D. H., Gambello, M. J., & Wong, M. (2011). Tsc2

2136 gene inactivation causes a more severe epilepsy phenotype than Tsc1 inactivation in a mouse

2137 model of Tuberous Sclerosis Complex. Human Molecular Genetics, 20(3), 445–454.

2138 https://doi.org/10.1093/hmg/ddq491

2139

2140 Zhang, X.-M., Zhu, S.-W., Duan, R.-S., Mohammed, A. H., Winblad, B., & Zhu, J. (2008). Gender

2141 differences in susceptibility to kainic acid-induced neurodegeneration in aged C57BL/6 mice. Neu-

2142 roToxicology, 29(3), 406–412. JOUR. https://doi.org/10.1016/j.neuro.2008.01.006

2143

76

2144 Table 1. Neuropathological alterations in different KA models in rats and mice.

2145

Administration Neuropathology References

mode

Rats Mice Rats Mice

Systemic Bilateral damage Bilateral damage Stafstrom et al., 1992 Hu et al., 1998

Entire hippo- CA3/CA1 Lado et al., 2006 McKhann et al., 2003

campus Cortical areas Williams et al., 2009 Benkovic et al., 2004

Subiculum Lateral amygdala White et al., 2010 Umpierre et al., 2017

Entorhinal cortex Dorsal thalamus Drexel et al., 2012

MF sprouting MF sprouting Bertoglio et al., 2017

ICV Unilateral lesion Unilateral damage Miyamoto et al., 1997 Cho et al., 2002

CA3/CA4 areas CA3/CA1 areas of Jing et al., 2009 Jeon et al., 2004

of the hippo- the hippocampus Gordon et al., 2014 Park et al., 2008

campus MF sprouting Song et al., 2016 Jin et al., 2009

MF sprouting Gao et al., 2019

Intra-HC Unilateral lesion Unilateral lesion Cavalheiro et al., 1982 Bouilleret et al., 1999

CA3/CA4 CA3/CA1 Bragin et al., 2005 Riban et al., 2002

DG granule cells DG granule cells Arkhipov et al., 2008 Gouder et al., 2003

dispersion dispersion Rattka et al., 2013 Gröticke et al., 2008

MF sprouting MF sprouting Klee et al., 2017 Lee et al., 2012

Zeidler et al., 2018

Intraamygdaloid Ipsilateral Am Ipsilateral Am Ben-Ari et al., 1979 Araki et al., 2002

Contralateral HC Contralateral HC Tanaka et al., 1992 Shinoda et al., 2004

Contralateral Am Contralateral Am Ueda et al., 2001 Mouri et al., 2008

CA3/CA1 CA3/CA1 Takebayashi et al., Tanaka et al., 2010

MF sprouting MF sprouting 2007 Li et al., 2012

Extratemporal Extratemporal Gurbanova et al., 2008 Liu et al., 2013

77

Administration Neuropathology References

mode

Supra-HC - Unilateral lesion - Bedner et al., 2015

CA3/CA1 Jefferys et al., 2016

DG granule cells Pitsch et al., 2019

dispersion

Intranasal - Bilateral damage - Chen et al., 2002

CA3 area of the Duan et al., 2006

hippocampus Zhang et al., 2008

Olfactory bulbs Lu et al., 2008

Sabilallah et al., 2016

2146 2147 2148 2149

2150 Am — amygdala; DG — dentate gyrus; HC — hippocampus; ICV — intraventricular; MF — mossy

2151 fibers;

2152

2153

78

2154 Table 2. Post-KA chronic epilepsy in rats.

2155

Administra- SE Latent SRS Comorbidities References

tion mode period

Systemic Starts within 2h 12-36 days 8 per day Cognitive im- Stafstrom et al., 1992

Lasts for 5-9 h (± 5.4) pairment Lado et al., 2006

22% mortality Aggressiveness Williams et al., 2009

Drexel et al., 2012

Bertoglio et al., 2017

ICV Starts within 7-14 days 5 per day Cognitive im- Miyamoto et al., 1997

10-30 min (± 3.5) pairment Jing et al., 2009

Lasts for 2-6h Gordon et al., 2014

<10% mortality Song et al., 2016

Gao et al., 2019

Intra-HC Starts within 5- 13-30 days 8 per day Cognitive im- Cavalheiro et al., 1982

20 min (± 6.3) pairment Bragin et al., 2005

Lasts for Perseverative Arkhipov et al., 2008

3-15 h behavior Rattka et al., 2013

<5% mortality Hyperexcitablty Klee et al., 2017

Intra-Am Starts within 5- 11-24 days 12 per day No data Ben-Ari et al., 1979

30 min (± 6.2) Tanaka et al., 1992

Lasts for 4-6 h Ueda et al., 2001

10% mortality Takebayashi et al., 2007

Gurbanova et al., 2008

2156 2157 2158 2159

2160

2161 Am — amygdala; HC — hippocampus; ICV — intraventricular; SE — status epilepticus; SRS — 79

2162 spontaneous recurrent seizures;

80

2163 Table 3. Post-KA chronic epilepsy in mice.

Administration SE Latent pe- SRS Comorbidities References

mode riod

Systemic Starts within 2-3 days Rarely ob- Hyperactivity Hu et al., 1998

15-30 min served Delayed mortality McKhann et al., 2003

Lasts for Infrequent Benkovic et al., 2004

2-6 h Puttachary et al., 2015

27% mortality Umpierre et al., 2016

Intra-HC Starts within 30 2-14 days 1-2 per Cognitive impair- Bouilleret et al., 1999

min week ment Riban et al., 2002

Lasts for Highly vari- Hyperexcitabiity Gouder et al., 2003

3-12 h able Depression (le- Gröticke et al., 2008

<5% mortality eSRS +++ sioned ventral Lee et al., 2012

hippocampus) Zeidler et al., 2018

Intra-Am Starts within 20 3-12 days 3 per day (± Anxiety Araki et al., 2002

min 2.5) Shinoda et al., 2004

Usually termi- Mouri et al., 2008

nated at 30-40 Tanaka et al., 2010

min Li et al., 2012

No mortality Liu et al., 2013

Supra-HC Lasts for 2-7 h 4-13 days 2.8 per day No data Bedner et al., 2015

<10% mortality (± 3.5) Jefferys et al., 2016

Pitsch et al., 2019

Intranasal Starts within 15-30 days Reported Increased locomo- Chen et al., 2002

15-30 min tion Duan et al., 2006

Lasts for 1-5h Zhang et al., 2008

~ 6.5% mortali- Lu et al., 2008

ty Sabilallah et al., 2016

2164 81

2165 2166 2167

2168

2169 Am — amygdala; eSRS — electrographic spontaneous recurrent seizures; HC — hippocampus;

2170 SE — status epilepticus; SRS — spontaneous recurrent seizures;

82