Next Article in Journal
The Biochemical Mechanism of Fork Regression in Prokaryotes and Eukaryotes—A Single Molecule Comparison
Previous Article in Journal
Molecular Mechanisms of Cytotoxicity of NCX4040, the Non-Steroidal Anti-Inflammatory NO-Donor, in Human Ovarian Cancer Cells
Previous Article in Special Issue
A Focal Impact Model of Traumatic Brain Injury in Xenopus Tadpoles Reveals Behavioral Alterations, Neuroinflammation, and an Astroglial Response
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aquatic Freshwater Vertebrate Models of Epilepsy Pathology: Past Discoveries and Future Directions for Therapeutic Discovery

School of Pharmacy, College of Health Sciences, University of Wyoming, Laramie, WY 82071, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8608; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23158608
Submission received: 6 May 2022 / Revised: 29 July 2022 / Accepted: 30 July 2022 / Published: 3 August 2022

Abstract

:
Epilepsy is an international public health concern that greatly affects patients’ health and lifestyle. About 30% of patients do not respond to available therapies, making new research models important for further drug discovery. Aquatic vertebrates present a promising avenue for improved seizure drug screening and discovery. Zebrafish (Danio rerio) and African clawed frogs (Xenopus laevis and tropicalis) are increasing in popularity for seizure research due to their cost-effective housing and rearing, similar genome to humans, ease of genetic manipulation, and simplicity of drug dosing. These organisms have demonstrated utility in a variety of seizure-induction models including chemical and genetic methods. Past studies with these methods have produced promising data and generated questions for further applications of these models to promote discovery of drug-resistant seizure pathology and lead to effective treatments for these patients.

1. Introduction

Epilepsy is a current public health concern with more than 50 million people affected around the world [1]. This disease takes a toll on the health and lifestyle of affected patients in the form of discrimination, comorbidities, and the constant stress of managing a chronic disease [2]. About 70% of patients attain adequate seizure control with available medications, but up to one-third of patients do not respond to available treatments [2,3,4,5,6,7]. The search for new epilepsy treatments requires biological models that can mirror the complexity of the human nervous system and allow researchers to discover new drug targets and drug molecules. Aquatic vertebrates provide promising alternative models for seizure modeling and drug screening and development. Zebrafish (Danio rerio) and Xenopus species of frog are two emerging aquatic models for seizure physiology and drug screening.

Classification and Etiology of Seizures

Seizures are bursts of excessive electrical activity in the brain. Uncontrolled electrical activity can lead to temporary abnormalities in both behavior and states of awareness. The International League Against Epilepsy (ILAE) has developed a standardized classification system for seizures [8] (Figure 1). Clinicians use guidelines for first-line treatment based on the seizure classification [9,10]. Overall, anti-seizure therapeutics vary based on the molecule’s pharmacokinetic properties, potential drug–drug interactions, side effects, and toxicities; however, most work mechanistically to suppress seizures [11]. Despite the large number of anti-seizure compounds on the market, drug-resistant epilepsy occurs in up to 30% of patients, refs. [2,3,4,5,6,7] and initial treatment fails in an additional 30% of patients due to intolerable side effects [12]. Discovery of therapeutic compounds that could affect disease progression is a required new direction for anti-seizure therapeutic discovery [13].
An additional challenge in developing effective therapies for seizures is that there are many underlying causes for seizures. For example, most seizures in newborns are often classified as acute symptomatic seizures and are usually due to a brain injury [14]. In contrast, neonatal-onset epilepsies are reported in ~12% of newborns with seizures and can be associated with genetic mutations [15,16,17,18,19], brain malformations [20,21], and metabolic disorders [22,23]. Model organisms that are amendable to genetic manipulations, brain imaging, and high-throughput screening of new compounds are critical for discovery of new therapeutics with different properties.

2. Aquatic Freshwater Vertebrate Animal Model Advantages

Both zebrafish (Danio rerio) and frogs (Xenopus laevis and Xenopus tropicalis) are cornerstones in biomedical research. They offer many advantages over the traditionally used mammalian models. Both zebrafish and frogs can be more easily raised and housed in large numbers, particularly at larval stages [24,25]. The transparency of zebrafish embryos and Xenopus tadpoles permits live imaging at the organismal level with a number of tissue-specific transgenic lines available to permit tracking of cellular dynamics in vivo [26,27]. High-throughput screens are possible for both drug discovery and toxicology [28,29,30,31,32,33,34]. A number of compounds identified in zebrafish screens are in early clinical trials further underscoring the translational potential of conducting drug screens in zebrafish [35]. With the advent of whole-genome sequencing, both zebrafish and Xenopus have stood out as model organisms among their mammalian counterparts. Over 70% of human genes have at least one zebrafish orthologue [36], with the Xenopus genome including orthologs of ~80% of human disease genes [37]. More recently, methods to increase the genome editing efficiency of the Clustered Regularly Interspaced Short Palindromic Repeat– (CRISPR–) Cas9 system in zebrafish [38,39,40,41,42,43] and Xenopus [44,45,46,47,48] have led to new human disease models.
Although zebrafish and Xenopus do not have the same level of regulatory recognition as their mammalian counterparts, government agencies such as the National Toxicology Program are now funding programs to increase the utility of aquatic vertebrates as pre-clinical models [49]. Indeed, a number of academic and pharmaceutical companies have teamed up with contract research service (CRO) companies that focus on zebrafish and Xenopus for preclinical drug development [50]. A growing list of drug treatments that have recently entered clinical trials after research starting in zebrafish is reviewed in [35].

2.1. Aquatic Freshwater Vertebrates as Seizure Models

Xenopus laevis oocytes have a long history of contributing to epilepsy research. Intact oocytes are a versatile expression system for functional investigation of ion channels and transporters [51]. As early as the 1980s, laboratories isolated RNA from mammalian brains, injected total RNA into oocytes, and recorded membrane currents [52,53,54,55,56,57]. Later, laboratories cloned and injected ion channel mRNA to characterize how genetic mutations linked to epilepsy affect the biophysical properties of specific channels [58,59]. Furthermore, treatment with epileptogenic agents identified the role voltage-gated potassium (Kv) channels and N-methyl-D-aspartate (NMDA) receptors play in seizure generation [60,61].
At the organismal level, zebrafish larvae have been used to model seizure disorders for some time while Xenopus tadpoles are a more recently developed aquatic vertebrate gaining notoriety in the field. Both of these organisms have advantages as models for seizures versus traditional mammalian models. In addition to the efficient rearing, low housing costs, and transparency described above, the central nervous system (CNS) of the zebrafish and Xenopus have similar organization to the human CNS [62,63,64,65,66,67,68,69]. Zebrafish and Xenopus have myelinated axons and are used as models for demyelination [70,71,72,73,74,75]. Aquatic freshwater vertebrate models of demyelination are particularly useful given that children with epilepsy have abnormal myelin development [76,77] and patients with demyelinating disease also suffer from seizures [78]. Lastly, zebrafish and Xenopus use gamma-aminobutyric acid (GABA) and glutamate receptors to control CNS activity including signaling required for movement [79,80,81,82,83]. Therefore, the process of seizure induction between freshwater aquatic vertebrates and humans are likely conserved, as GABA and glutamate are the main neurotransmitters contributing to the pathophysiology of epilepsy.

2.2. Advantages of Aquatic Freshwater Vertebrates as Seizure Models

A major advantage of zebrafish larva and Xenopus tadpoles is the ability to visualize the nervous system during seizure activity in real time. Using calcium imaging, scientists can look at both excitatory and inhibitory activity throughout the brain, identify areas of high activity, and assess the rate of spread of the seizure to other parts of the brain [84,85,86,87,88,89,90,91]. Calcium imaging can be combined with a variety of electrophysiological approaches to get a direct correlation between neuronal activity and electrical field potential [92,93,94,95]. With recent advantages in calcium imaging, it is now also possible to get single cell resolution in an acute zebrafish seizure model [96].
In addition to their transparency, the aquatic nature of these organisms allows for easy drug delivery by dosing the water they reside in. Specific concentrations of drugs can be delivered by creating various baths and placing the desired animals in the solution. Drug exposure using this method allows for control of dose, frequency, and length of exposure more so than administering a drug orally or parenterally. The ability to cultivate large numbers of larvae or tadpoles allows for high-throughput screening options for drugs [97]. By dosing the water and allowing zebrafish and Xenopus to behave freely, the data are not confounded by anesthetics or invasive surgical procedures as found in mammalian models providing cleaner results due to freedom from interference in the neural circuits [98]. Taken together, aquatic vertebrates provide an animal platform with improved control of dose- and timing-related phenomena.
Perhaps the largest advantage aquatic freshwater vertebrates have in modeling seizures is their genetic tractability. Recent studies have found approximately 900 genes associated with epilepsy [99], making genetic animal models of seizures ever more useful for studying the disease. The ease of genetic manipulation in zebrafish and Xenopus affords them a major advantage over mammalian models as mutations in multiple types of proteins can be made in combination. For example, genetic seizure disorders are highly variable and prone to drug resistance [100]. However, it is less understood whether resistance is due to mutations that cause changes in the pharmacokinetics and pharmacodynamics of the therapeutic molecules or inherent differences in the pathophysiology of the seizure. The extensive genetic toolkit available in aquatic freshwater vertebrates permits modeling of knockdown [101,102,103,104,105,106,107], knockout [39,44,46,48,108,109,110,111,112,113,114,115,116], ectopic, and overexpression [38,117,118,119,120,121] of seizure-associated genes in combination with metabolic enzymes, transporters, and other proteins required for absorption, distribution, degradation, and excretion of therapeutics. A unique aspect of Xenopus is that their large eggs and embryos can be genetically modified by injecting the embryo on only one side at the two-cell stage, providing an internal control [120]. This causes the alteration to occur only on the injected side, so the resulting phenotype can be compared to the contralateral side for off-target effects. Seizure activity can be monitored by similar behaviors in both species in combination with calcium imaging and electrophysiology.
Lastly, researchers in both model organism communities use a similar rating scale for seizures. The rating scale consists of five categories ranging from barely noticeable locomotor changes (category 1) to C-shaped contractions (category 5) [98,122,123,124]. Although zebrafish can exhibit some human symptoms, such as tonic-like behavior, this scoring system does not directly relate to the ILAE classification. Instead, this scoring system provides a method for standardization of experimental technique, which allows the data to be reproducible and translatable across laboratories and different aquatic species. Other behavioral metrics used to analyze seizure activity include total distance traveled and thigmotaxis.

2.3. Disadvantages of Aquatic Freshwater Vertebrates as Seizure Models

Like most model organisms, both zebrafish and Xenopus have disadvantages. Adults develop pigment, making live imaging more difficult in older animals. Genetic mutants which lack pigment may alleviate live imaging limitations of adult animals. As techniques are developed to image brain activity in older animals, the use of juvenile and adult models of seizure disorders should increase. Indeed, a genetic model of juvenile myoclonic epilepsy in older zebrafish has demonstrated convulsive seizure generation in response to light [125]. More recently, an EEG system was developed for adult zebrafish, permitting a direct comparison between zebrafish and mammalian models [126].
Although dosing the water is facile, uptake of the compounds through skin and gills varies, creating pharmacokinetic challenges such as indirect measurements of drug absorption into plasma. Though most studies in zebrafish do not measure blood concentration of compounds, these measurements could prove essential for translating the therapeutic potential of aquatic vertebrate models to the clinic [127]. Given blood samples from both zebrafish and Xenopus are small, pooled samples may be necessary to get the required pharmacokinetic data [128,129]. Small molecules with poor water solubility are also not absorbed efficiently, limiting the chemical libraries that can be screened by adding the molecule to the water. Furthermore, drugs which do not readily cross the mammalian blood–brain barrier are also not observed in the zebrafish brain after water administration [130].

3. Current Aquatic Vertebrate Seizure Models

Chemical and genetic seizure induction are the most common options in the zebrafish and Xenopus communities. Below we highlight the progress made toward understanding seizure pathophysiology and treatment using these methods.

3.1. Chemical Induction of Seizures

Seizures in humans are due to hyperexcitability and hypersynchronous activity of cortical neurons. Therefore, most chemically induced models of aquatic vertebrates are generated using compounds that disrupt the inhibitory and excitatory balance in the brain of the animal (Figure 2). Possible chemical induction agents include bicuculline, picrotoxin, tetramethylenedisulfotetramine (TETS), kainic acid, pilocarpine, 4-aminopyridine (4-AP), and pentylenetetrazole (PTZ) [98]. Most of the studies in aquatic organisms that have utilized non-PTZ chemical induction methods are in zebrafish. Chemical induction agents are advantageous due to rapid seizure induction and ease of use. They also facilitate high-throughput screening by permitting rapid dose response studies in multiple animal ages to determine desired characteristics prior to the main data-generating experiment. Zebrafish studies have used animals as young as 2 days post-fertilization (2 dpf) up to adulthood whereas the Xenopus community typically uses tadpoles at developmental stages 42–49.
Bicuculline is a competitive GABAA antagonist [131] originally used to elucidate details surrounding synapses and GABAA transmission [132,133]. It is a light-sensitive molecule, which easily decomposes in solution, making it difficult to administer [131]. Therefore, studies using this compound are limited in free-swimming aquatic vertebrates. One advantage of bicuculline is the ability to induce seizures in multiple models from C. elegans to cats [134,135,136], providing a clear translational path from target identification in aquatic organisms to mammals. In rats, natural product screens identified sakuranetin and melittin as protective against bicuculline-induced seizures [137,138]. However, exposure to bicuculline does not cause significantly different electrical recordings than those generated from other chemically induced seizures [134]. Given the difficulty working with bicuculline and lack of unique electrical changes induced, its utility in aquatic vertebrates is still limited.
Picrotoxin is also a GABAA antagonist but unlike bicuculline, is a non-competitive inhibitor [132,133]. In zebrafish larvae (5 dpf), picrotoxin exposure increased locomotion in a dose-dependent manner. Similarly, Xenopus tadpoles also displayed concentration-dependent seizures [98]. In addition, high doses increased thigmotaxis in zebrafish larvae, which is a common measure for anxiety. Furthermore, higher doses of acute anti-seizure medications are required to decrease locomotor seizure symptoms when induced with this compound [94]. Therefore, picrotoxin may be advantageous in modeling treatment-resistant seizures. Picrotoxin-treated adult zebrafish exhibited increased hyperactivity and cortisol levels following seizure [139]. The data from larvae and adults are consistent with studies where low doses of picrotoxin elevate anxiety and corticosterone in mice [140] and lysosomal dysfunction in rats [141,142]. Therefore, the picrotoxin model in adult zebrafish may provide a useful tool to probe seizure-induced effects on the endocrine system. The seizure-inducing properties of picrotoxin makes it a credible chemical threat to humans. In fact, the NIH Countermeasures Against Chemical Threats (CounterACT) program listed picrotoxin as an agent of interest. In addition to seizure modeling, picrotoxin in zebrafish is a good platform to screen symptoms and potential treatments for the use of picrotoxin against human populations in warfare [94].
Similar to picrotoxin, TETS is also a non-competitive GABAA antagonist that is considered a potential chemical warfare agent. While TETS is a potent rodenticide, it also causes human seizures with lasting neurological effects, leading to a worldwide ban [143]. The recurrent nature of seizures after TETS exposure in humans as well as the lack of a targeted treatment make it attractive as a seizure-induction agent for research using aquatic vertebrate models. Compared to picrotoxin, in zebrafish larvae less TETS is required to evoke seizure behavior [94,144]. TETS also triggered high-frequency electrical discharges which were different from electrical measurements following picrotoxin exposure. Benzodiazepine treatment in larvae attenuated some of the electrical changes without a full return to baseline, consistent with what is seen in human exposure. Future studies aimed at identifying TETS-protective compounds in zebrafish will facilitate the development of effective antidotes for this poison.
Kainic acid is an analog of glutamate, which promotes excitability leading to seizure activity. It activates both kainate and AMPA receptors. One advantage of this drug is that it causes focal seizures in both mammals and primates [145,146,147]. Kainic acid causes cell death and damage to the brain, but zebrafish can regenerate brain cells [148]. Upon regeneration, after kainic acid treatment, the brain becomes disorganized which leads to a chronic spontaneous seizure state similar to epilepsy in humans. One study showed a lack of response of this seizure model to clinically used anti-seizure medications [149]. This lack of response underscores the potential of this model to identify drugs for treatment of drug-resistant seizures. The major disadvantage of this model is that even in aquatic organisms, kainic acid is not well absorbed and must be injected, making administration more technically challenging and time intensive.
Pilocarpine induces seizures through activation of the cholinergic system in the brain, specifically agonizing the muscarinic (M1) receptor. Activation of the cholinergic system leads to activation of NMDA receptors and hyperexcitability [150]. Some evidence suggests that pilocarpine seizures have different behavioral characteristics than PTZ-induced seizures. This makes pilocarpine beneficial for modeling different types of seizures possibly leading to new therapeutic options [150]. Multiple studies have used pilocarpine especially to model seizures in adult zebrafish with emphasis on chronic seizure modeling through repeated dosing [150,151]. Repeated dosing is particularly attractive because patients typically present with recurrent seizures, so modeling this will allow for better drug discovery leading to treatment of recurrent versus acute seizures.
4-AP is a voltage-sensitive potassium (K+) channel blocker. K+ channel inhibition by 4-AP results in increased cholinergic signaling in the CNS and neuromuscular junctions resulting in clonic seizures [152]. 4-AP was originally used to repel and kill birds, and it is toxic to mammals as well [153]. Despite its toxicity, 4-AP is used as a therapeutic for multiple sclerosis and spinal cord injury patients [154,155,156]. Therefore, 4-AP-induced seizures are a good model for seizure-threshold lowering therapeutics. In Xenopus, 4-AP induces both behavioral and rhythmic high-amplitude electrical discharges [98]. Similarly, zebrafish larvae also exhibit increased swimming activity when treated with 4-AP. Recently, a screen to test the efficacy of current anti-seizure compounds was performed in hippocampal-entorhinal slices of adult rats with 4-AP-induced seizures [152]. Given the ease of drug application in aquatic vertebrates, it is possible to simultaneously administer a seizure-threshold lowering drug such as 4-AP at a low dose and screen for seizure-protective compounds.
The most common chemical induction model in zebrafish and Xenopus is PTZ. PTZ is a GABAA antagonist and as such PTZ-induced seizures are sensitive to drugs acting directly on GABAA receptors. The PTZ model is ideal for its translation to mammalian models. PTZ has been used to induce seizures in rodents [157], canines [158], and primates [159]. Using the PTZ model, early zebrafish studies focused on zebrafish larvae and their changes in electrical activity [97]. Pharmacologically, PTZ seizures and identified anti-seizure compounds from these screens in zebrafish larvae are consistent with results from rodent studies [160,161]. The PTZ-induced seizure model in zebrafish larvae has also been used to screen a variety of natural compounds for anti-seizure activity [162,163,164,165]. These screens aimed to discover novel anti-seizure therapies for treating refractory seizure disorders. While larvae may be more versatile for imaging and replicating developmental neural tissue, adult zebrafish still have a role in anti-seizure drug research with potential applications to older human patients. One study used adult zebrafish instead of larvae to screen the anti-seizure properties of leaf extracts [166]. Taken together, the PTZ-zebrafish model is a first-line drug-screening tool for discovery of anti-seizure medications. Further, the model shows promise as a first line tool for identifying the adverse effects of seizure induction in medications indicated for other diseases.
Xenopus models have also demonstrated efficacy as a drug-discovery tool for seizures. Most Xenopus studies use PTZ for chemical seizure induction. In fact, the Haas laboratory tested the different seizure-induction agents described above to determine the most consistent seizure model in Xenopus [98]. PTZ is preferred for its wide therapeutic window and consistent timing and intensity of seizure activity in the tadpole [98]. Similar to zebrafish, the Xenopus community uses intracellular calcium variations to complement locomotor assays. Furthermore, the PTZ-induced Xenopus seizure model exhibits electrical events consistent with those seen in other vertebrates [167].

3.2. Genetic Induction of Seizures

Originally, genetic seizure models were generally limited to forward-genetic screens using N-ethyl-N-nitrosourea (ENU) mutagenesis. As genetic engineering techniques bloomed, multiple models became available. For example, gene knockdown with antisense morpholinos has been used to model autosomal dominant partial epilepsy with auditory features and temporal lobe epilepsy [168,169]. Morpholino knockdown can also be used to probe the mechanism of disease-linked genes in epilepsy formation [170]. Unfortunately, morpholinos come with their own set of limitations including off-target effects, transient effects that wane after days, and noted discrepancies between morphants and mutant phenotypes.
Given the growing list of concerns surrounding morpholinos, laboratories derived genetic seizure models from clinically relevant mutations in humans that resulted in seizure disorders. The Baraban laboratory published a phenotypic analysis of 40 single-gene mutant zebrafish lines based on genes implicated in childhood epilepsy [171]. For example, Dravet syndrome is due to mutations in the NaV1.1 voltage-gated sodium channel. A phenotype-based screen in mutant zebrafish larvae identified clemizole as an anti-seizure compound [172]. In addition, zebrafish with mutations in the syntaxin binding protein, stxbp1, also respond to clemizole [173]. These and future genetic models are expected to provide much needed information on the pathophysiology of childhood epilepsies and identify new classes of anti-seizure compounds paving the way for patient-specific therapeutics.
Xenopus models are also genetically amendable. Using cRNA injection technology, a Xenopus oocyte model was used to find a non-conventional seizure treatment with direct application to a human patient [174]. This study directly demonstrates the translational relevance of data from these aquatic vertebrates. A Xenopus model using CRISPR-Cas9-mediated genome editing to deplete neurod2, a gene implicated in early infantile epileptic encephalopathy, induces spontaneous seizures in tadpoles mimicking the human disease [175]. In addition to genetic manipulation, Xenopus tadpoles and oocytes have been used to determine molecular details of clinically utilized and novel drug molecules [176,177,178]. For example, the commonly used drug for absence seizures, ethosuximide, inhibits specific G-protein activated inwardly rectifying potassium channels (GIRK) which has the potential to affect other systems beyond the brain [179]. Another study demonstrated how multiple compounds can interact pharmacodynamically to produce increased effects for seizure treatment [180]. Additionally, the effects of polyamine synthesis during seizure activity are controversial. Data from the Xenopus model support the hypothesis that polyamines are protective against recurrent seizure activity which may indicate novel therapeutic targets [181]. As there are data indicating polyamines are not protective in mammals [182], further studies across species may be necessary to make a final determination as to the protective effect of post-seizure polyamine generation.

3.3. New but Less Established Aquatic Vertebrate Seizure Models

Not all anti-seizure drugs are active against multiple types of induced seizures. Screening of currently approved anti-seizure drugs showed differences in efficacy against PTZ-induced seizures and one model of genetic seizures [183]. To identify new models for anti-seizure efficacy, the Kurrasch laboratory developed zebrafish monitoring for seizure activity using mitochondrial respiration [183]. This platform identified a new compound, vorinostatcan that decreased seizures in both genetic and chemical induction models of epilepsy. Unfortunately, locomotor seizure activity does not correlate well with increased mitochondrial respiration, requiring more refinement of this model.
Zebrafish have also been used to model traumatic brain injuries and the subsequent seizure pathology seen in humans [184]. Pre-treatment of injured zebrafish with sonic hedgehog signaling inhibitors after injury rendered them resistant to anti-seizure compounds. This combinatory model of injury and drug resistance has potential as a model for drug-resistant epilepsies as these seizures are characteristically difficult to control. Traumatic brain injury models may prove useful for understanding other adult-onset acquired seizure disorders. For example, one study evaluated the development of seizures in zebrafish after injury and then tested the effectiveness of anti-seizure therapeutics using a T-maze setup to test cognitive ability before and after treatment [122]. This study highlights the importance of using multiple types of data (electrophysiological and locomotion) to track seizure activity and the effectiveness of potential treatments. Given the rise of another aquatic vertebrate, Medaka (Oryzias latipes), for studying traumatic brain injuries [185], this paradigm has the potential to expand our ability to model acquired seizure disorders in both a pro-regenerative and less regenerative species.

4. Looking to the Future for Aquatic Seizure Models

Although the PTZ model remains the most characterized and commonly used chemically induced seizure model from zebrafish to primates, aquatic vertebrates provide an opportunity to develop and validate new chemically induced seizure models. Using a variety of compounds that act through distinct pathways to induce seizures provides a broader set of models for screening anti-seizure therapeutics. These additional chemical models may uncover different anti-seizure therapeutics that could benefit treatment-resistant epilepsies in humans. Similarly, studies have shown that not all clinically effective antiepileptic drugs are active in a single seizure-induction model; therefore, aquatic vertebrates provide a model in which to test potential new drug compounds against multiple induction models before declaring them not useful. This leads to the possibility that a new pro-convulsive agent, that has yet to be considered, could be the key to modeling treatment-resistant seizure disorders.
In addition to chemically induced seizure models, aquatic vertebrates offer numerous genetic options. Stable genetic models are available for specific epilepsies such as Dravet syndrome and epilepsies linked to rare syndromes such as Angelman syndrome. Many of these models serve as a starting point for future studies aimed at gaining mechanistic insight into seizure generation and propagation. Furthermore, screening for anti-seizure therapeutics in these models permits the discovery of gene-specific therapeutics opening the door for personalized medicine in human patients.
Aquatic vertebrates permit the combination of different seizure models to determine the differences in seizure pathology and thus provide a path toward novel therapeutic options especially for drug-resistant epilepsies. The ease of housing, large clutches, transparent nature, aquatic environment, and potential for high-throughput screening make zebrafish and Xenopus ideal model organisms for research into the pathology and treatment of seizure disorders. The plethora of genetic resources in these models promises to provide new research opportunities and drug development directions. As genetic models are developed in new aquatic vertebrates, such as Medaka, the ability to compare evolutionary development of seizures and mechanisms of drug-resistance across models exists [186].
Continued development of imaging and screening approaches in aquatic vertebrates will increase the number and types of neurophysiological questions that can be addressed when studying seizure generation and treatment. The combination of new tools and genetic resources will accelerate the contribution of studies to translational research.

Author Contributions

Conceptualization, R.E.W. and K.M.; writing—original draft preparation, R.E.W.; writing—review and editing, R.E.W. and K.M.; supervision, K.M.; project administration, K.M.; funding acquisition, K.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institutes of Health (GM143565 and GM121310) to K.M.

Acknowledgments

We thank members of the Mruk lab for critical reading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. Epilepsy. 2022. Available online: https://www.who.int/health-topics/epilepsy#tab=tab_1 (accessed on 27 April 2022).
  2. Moshe, S.L.; Perucca, E.; Ryvlin, P.; Tomson, T. Epilepsy: New advances. Lancet 2015, 385, 884–898. [Google Scholar] [CrossRef]
  3. Go, C.; Snead, O.C., 3rd. Pharmacologically intractable epilepsy in children: Diagnosis and preoperative evaluation. Neurosurg Focus 2008, 25, E2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Annegers, J.F.; Hauser, W.A.; Elveback, L.R. Remission of seizures and relapse in patients with epilepsy. Epilepsia 1979, 20, 729–737. [Google Scholar] [CrossRef] [PubMed]
  5. Cockerell, O.C.; Johnson, A.L.; Sander, J.W.; Hart, Y.M.; Shorvon, S.D. Remission of epilepsy: Results from the National General Practice Study of Epilepsy. Lancet 1995, 346, 140–144. [Google Scholar] [CrossRef]
  6. Kwan, P.; Arzimanoglou, A.; Berg, A.T.; Brodie, M.J.; Allen Hauser, W.; Mathern, G.; Moshe, S.L.; Perucca, E.; Wiebe, S.; French, J. Definition of drug resistant epilepsy: Consensus proposal by the ad hoc Task Force of the ILAE Commission on Therapeutic Strategies. Epilepsia 2010, 51, 1069–1077. [Google Scholar] [CrossRef] [PubMed]
  7. Kwan, P.; Brodie, M.J. Early identification of refractory epilepsy. N. Engl. J. Med. 2000, 342, 314–319. [Google Scholar] [CrossRef]
  8. Fisher, R.S.; Cross, J.H.; French, J.A.; Higurashi, N.; Hirsch, E.; Jansen, F.E.; Lagae, L.; Moshé, S.L.; Peltola, J.; Roulet Perez, E.; et al. Operational classification of seizure types by the International League Against Epilepsy: Position Paper of the ILAE Commission for Classification and Terminology. Epilepsia 2017, 58, 522–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Kanner, A.M.; Ashman, E.; Gloss, D.; Harden, C.; Bourgeois, B.; Bautista, J.F.; Abou-Khalil, B.; Burakgazi-Dalkilic, E.; Llanas Park, E.; Stern, J.; et al. Practice guideline update summary: Efficacy and tolerability of the new antiepileptic drugs I: Treatment of new-onset epilepsy: Report of the Guideline Development, Dissemination, and Implementation Subcommittee of the American Academy of Neurology and the American Epilepsy Society. Neurology 2018, 91, 74–81. [Google Scholar] [CrossRef] [Green Version]
  10. Kanner, A.M.; Ashman, E.; Gloss, D.; Harden, C.; Bourgeois, B.; Bautista, J.F.; Abou-Khalil, B.; Burakgazi-Dalkilic, E.; Park, E.L.; Stern, J.; et al. Practice guideline update summary: Efficacy and tolerability of the new antiepileptic drugs I: Treatment of new-onset epilepsy: Report of the American Epilepsy Society and the Guideline Development, Dissemination, and Implementation Subcommittee of the American Academy of Neurology. Epilepsy Curr. 2018, 18, 260–268. [Google Scholar] [CrossRef] [PubMed]
  11. French, J.A.; Perucca, E. Time to Start Calling Things by Their Own Names? The Case for Antiseizure Medicines. Epilepsy Curr. 2020, 20, 69–72. [Google Scholar] [CrossRef] [Green Version]
  12. Hakami, T. Efficacy and tolerability of antiseizure drugs. Ther. Adv. Neurol. Disord. 2021, 14, 17562864211037430. [Google Scholar] [CrossRef] [PubMed]
  13. Löscher, W.; Schmidt, D. Modern antiepileptic drug development has failed to deliver: Ways out of the current dilemma. Epilepsia 2011, 52, 657–678. [Google Scholar] [CrossRef] [PubMed]
  14. Shellhaas, R.A. Seizure classification, etiology, and management. Handb. Clin. Neurol. 2019, 162, 347–361. [Google Scholar] [CrossRef]
  15. Maljevic, S.; Vejzovic, S.; Bernhard, M.K.; Bertsche, A.; Weise, S.; Döcker, M.; Lerche, H.; Lemke, J.R.; Merkenschlager, A.; Syrbe, S. Novel KCNQ3 Mutation in a Large Family with Benign Familial Neonatal Epilepsy: A Rare Cause of Neonatal Seizures. Mol. Syndromol. 2016, 7, 189–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Piro, E.; Nardello, R.; Gennaro, E.; Fontana, A.; Taglialatela, M.; Mangano, G.D.; Corsello, G.; Mangano, S. A novel mutation in KCNQ3-related benign familial neonatal epilepsy: Electroclinical features and neurodevelopmental outcome. Epileptic Disord. Int. Epilepsy J. Videotape 2019, 21, 87–91. [Google Scholar] [CrossRef]
  17. Li, H.; Li, N.; Shen, L.; Jiang, H.; Yang, Q.; Song, Y.; Guo, J.; Xia, K.; Pan, Q.; Tang, B. A novel mutation of KCNQ3 gene in a Chinese family with benign familial neonatal convulsions. Epilepsy Res. 2008, 79, 1–5. [Google Scholar] [CrossRef]
  18. Miceli, F.; Striano, P.; Soldovieri, M.V.; Fontana, A.; Nardello, R.; Robbiano, A.; Bellini, G.; Elia, M.; Zara, F.; Taglialatela, M.; et al. A novel KCNQ3 mutation in familial epilepsy with focal seizures and intellectual disability. Epilepsia 2015, 56, e15–e20. [Google Scholar] [CrossRef]
  19. Nardello, R.; Mangano, G.D.; Miceli, F.; Fontana, A.; Piro, E.; Salpietro, V. Benign familial infantile epilepsy associated with KCNQ3 mutation: A rare occurrence or an underestimated event? Epileptic Disord. Int. Epilepsy J. Videotape 2020, 22, 807–810. [Google Scholar] [CrossRef] [PubMed]
  20. Guerrini, R.; Barba, C. Focal cortical dysplasia: An update on diagnosis and treatment. Expert Rev. Neurother. 2021, 21, 1213–1224. [Google Scholar] [CrossRef]
  21. Koenig, M.; Dobyns, W.B.; Di Donato, N. Lissencephaly: Update on diagnostics and clinical management. Eur. J. Paediatr. Neurol. EJPN Off. J. Eur. Paediatr. Neurol. Soc. 2021, 35, 147–152. [Google Scholar] [CrossRef] [PubMed]
  22. Hundallah, K.; Tabarki, B. Treatable inherited metabolic epilepsies. Neurosciences 2021, 26, 229–235. [Google Scholar] [CrossRef]
  23. Pearl, P.L.; Bennett, H.D.; Khademian, Z. Seizures and metabolic disease. Curr. Neurol. Neurosci. Rep. 2005, 5, 127–133. [Google Scholar] [CrossRef] [PubMed]
  24. Choi, T.Y.; Choi, T.I.; Lee, Y.R.; Choe, S.K.; Kim, C.H. Zebrafish as an animal model for biomedical research. Exp. Mol. Med. 2021, 53, 310–317. [Google Scholar] [CrossRef]
  25. Ishibashi, S.; Saldanha, F.Y.L.; Amaya, E. Chapter 14—Xenopus as a Model Organism for Biomedical Research. In Basic Science Methods for Clinical Researchers; Jalali, M., Saldanha, F.Y.L., Jalali, M., Eds.; Academic Press: Boston, MA, USA, 2017; pp. 263–290. [Google Scholar]
  26. Kwan, K.M.; Fujimoto, E.; Grabher, C.; Mangum, B.D.; Hardy, M.E.; Campbell, D.S.; Parant, J.M.; Yost, H.J.; Kanki, J.P.; Chien, C.B. The Tol2kit: A multisite gateway-based construction kit for Tol2 transposon transgenesis constructs. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2007, 236, 3088–3099. [Google Scholar] [CrossRef] [PubMed]
  27. Horb, M.; Wlizla, M.; Abu-Daya, A.; McNamara, S.; Gajdasik, D.; Igawa, T.; Suzuki, A.; Ogino, H.; Noble, A.; Centre de Ressource Biologique Xenope Team in France; et al. Xenopus Resources: Transgenic, Inbred and Mutant Animals, Training Opportunities, and Web-Based Support. Front. Physiol. 2019, 10, 387. [Google Scholar] [CrossRef]
  28. Thessen, A.E.; Marvel, S.; Achenbach, J.C.; Fischer, S.; Haendel, M.A.; Hayward, K.; Klüver, N.; Könemann, S.; Legradi, J.; Lein, P.; et al. Implementation of Zebrafish Ontologies for Toxicology Screening. Front. Toxicol. 2022, 4, 817999. [Google Scholar] [CrossRef]
  29. Padilla, S.; Corum, D.; Padnos, B.; Hunter, D.L.; Beam, A.; Houck, K.A.; Sipes, N.; Kleinstreuer, N.; Knudsen, T.; Dix, D.J.; et al. Zebrafish developmental screening of the ToxCast™ Phase I chemical library. Reprod. Toxicol. 2012, 33, 174–187. [Google Scholar] [CrossRef]
  30. Truong, L.; Reif, D.M.; St Mary, L.; Geier, M.C.; Truong, H.D.; Tanguay, R.L. Multidimensional in vivo hazard assessment using zebrafish. Toxicol. Sci. Off. J. Soc. Toxicol. 2014, 137, 212–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Tice, R.R.; Austin, C.P.; Kavlock, R.J.; Bucher, J.R. Improving the human hazard characterization of chemicals: A Tox21 update. Environ. Health Perspect. 2013, 121, 756–765. [Google Scholar] [CrossRef] [Green Version]
  32. Mannioui, A.; Vauzanges, Q.; Fini, J.B.; Henriet, E.; Sekizar, S.; Azoyan, L.; Thomas, J.L.; Pasquier, D.D.; Giovannangeli, C.; Demeneix, B.; et al. The Xenopus tadpole: An in vivo model to screen drugs favoring remyelination. Mult. Scler. 2018, 24, 1421–1432. [Google Scholar] [CrossRef] [Green Version]
  33. Mughal, B.B.; Demeneix, B.A.; Fini, J.B. Evaluating Thyroid Disrupting Chemicals In Vivo Using Xenopus laevis. Methods Mol. Biol. 2018, 1801, 183–192. [Google Scholar] [CrossRef]
  34. Sullivan, K.G.; Levin, M. Inverse Drug Screening of Bioelectric Signaling and Neurotransmitter Roles: Illustrated Using a Xenopus Tail Regeneration Assay. Cold Spring Harb. Protoc. 2018, 2018. [Google Scholar] [CrossRef] [PubMed]
  35. Patton, E.E.; Zon, L.I.; Langenau, D.M. Zebrafish disease models in drug discovery: From preclinical modelling to clinical trials. Nat. Rev. Drug Discov. 2021, 20, 611–628. [Google Scholar] [CrossRef]
  36. Howe, K.; Clark, M.D.; Torroja, C.F.; Torrance, J.; Berthelot, C.; Muffato, M.; Collins, J.E.; Humphray, S.; McLaren, K.; Matthews, L.; et al. The zebrafish reference genome sequence and its relationship to the human genome. Nature 2013, 496, 498–503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Hellsten, U.; Harland, R.M.; Gilchrist, M.J.; Hendrix, D.; Jurka, J.; Kapitonov, V.; Ovcharenko, I.; Putnam, N.H.; Shu, S.; Taher, L.; et al. The genome of the Western clawed frog Xenopus tropicalis. Science 2010, 328, 633–636. [Google Scholar] [CrossRef] [Green Version]
  38. Chen, S.; Jiao, Y.; Pan, F.; Guan, Z.; Cheng, S.H.; Sun, D. Knock-in of a Large Reporter Gene via the High-Throughput Microinjection of the CRISPR/Cas9 System. IEEE Trans. Bio-Med. Eng. 2022. [Google Scholar] [CrossRef] [PubMed]
  39. Colijn, S.; Yin, Y.; Stratman, A.N. High-throughput methodology to identify CRISPR-generated Danio rerio mutants using fragment analysis with unmodified PCR products. Dev. Biol. 2022, 484, 22–29. [Google Scholar] [CrossRef]
  40. Guo, S.; Gao, G.; Zhang, C.; Peng, G. Multiplexed Genome Editing for Efficient Phenotypic Screening in Zebrafish. Vet. Sci. 2022, 9, 92. [Google Scholar] [CrossRef] [PubMed]
  41. Han, B.; Zhang, Y.; Zhou, Y.; Zhang, B.; Krueger, C.J.; Bi, X.; Zhu, Z.; Tong, X.; Zhang, B. ErCas12a and T5exo-ErCas12a Mediate Simple and Efficient Genome Editing in Zebrafish. Biology 2022, 11, 411. [Google Scholar] [CrossRef]
  42. Hernandez-Huertas, L.; Kushawah, G.; Diaz-Moscoso, A.; Tomas-Gallardo, L.; Moreno-Sanchez, I.; da Silva Pescador, G.; Bazzini, A.A.; Moreno-Mateos, M.A. Optimized CRISPR-RfxCas13d system for RNA targeting in zebrafish embryos. STAR Protoc. 2022, 3, 101058. [Google Scholar] [CrossRef]
  43. Thumberger, T.; Tavhelidse-Suck, T.; Gutierrez-Triana, J.A.; Cornean, A.; Medert, R.; Welz, B.; Freichel, M.; Wittbrodt, J. Boosting targeted genome editing using the hei-tag. eLife 2022, 11, e70558. [Google Scholar] [CrossRef]
  44. Blitz, I.L.; Nakayama, T. CRISPR-Cas9 Mutagenesis in Xenopus tropicalis for Phenotypic Analyses in the F(0) Generation and Beyond. Cold Spring Harb. Protoc. 2022, 2022. [Google Scholar] [CrossRef]
  45. Corkins, M.E.; DeLay, B.D.; Miller, R.K. Tissue-Targeted CRISPR-Cas9-Mediated Genome Editing of Multiple Homeologs in F(0)-Generation Xenopus laevis Embryos. Cold Spring Harb. Protoc. 2022, 2022. [Google Scholar] [CrossRef]
  46. Godden, A.M.; Antonaci, M.; Ward, N.J.; van der Lee, M.; Abu-Daya, A.; Guille, M.; Wheeler, G.N. An efficient miRNA knockout approach using CRISPR-Cas9 in Xenopus. Dev. Biol. 2022, 483, 66–75. [Google Scholar] [CrossRef]
  47. Parain, K.; Lourdel, S.; Donval, A.; Chesneau, A.; Borday, C.; Bronchain, O.; Locker, M.; Perron, M. CRISPR/Cas9-Mediated Models of Retinitis Pigmentosa Reveal Differential Proliferative Response of Müller Cells between Xenopus laevis and Xenopus tropicalis. Cells 2022, 11, 807. [Google Scholar] [CrossRef] [PubMed]
  48. Tanouchi, M.; Igawa, T.; Suzuki, N.; Suzuki, M.; Hossain, N.; Ochi, H.; Ogino, H. Optimization of CRISPR/Cas9-mediated gene disruption in Xenopus laevis using a phenotypic image analysis technique. Dev. Growth Differ. 2022. [Google Scholar] [CrossRef]
  49. Hamm, J.T.; Ceger, P.; Allen, D.; Stout, M.; Maull, E.A.; Baker, G.; Zmarowski, A.; Padilla, S.; Perkins, E.; Planchart, A.; et al. Characterizing sources of variability in zebrafish embryo screening protocols. Altex 2019, 36, 103–120. [Google Scholar] [CrossRef] [PubMed]
  50. Alzualde, A.; Behl, M.; Sipes, N.S.; Hsieh, J.H.; Alday, A.; Tice, R.R.; Paules, R.S.; Muriana, A.; Quevedo, C. Toxicity profiling of flame retardants in zebrafish embryos using a battery of assays for developmental toxicity, neurotoxicity, cardiotoxicity and hepatotoxicity toward human relevance. Neurotoxicol. Teratol. 2018, 70, 40–50. [Google Scholar] [CrossRef]
  51. O’Connell, D.; Mruk, K.; Rocheleau, J.M.; Kobertz, W.R. Xenopus laevis oocytes infected with multi-drug-resistant bacteria: Implications for electrical recordings. J. Gen. Physiol. 2011, 138, 271–277. [Google Scholar] [CrossRef] [Green Version]
  52. Gundersen, C.B.; Miledi, R.; Parker, I. Serotonin receptors induced by exogenous messenger RNA in Xenopus oocytes. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1983, 219, 103–109. [Google Scholar] [CrossRef]
  53. Gundersen, C.B.; Miledi, R.; Parker, I. Glutamate and kainate receptors induced by rat brain messenger RNA in Xenopus oocytes. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1984, 221, 127–143. [Google Scholar] [CrossRef]
  54. Houamed, K.M.; Bilbe, G.; Smart, T.G.; Constanti, A.; Brown, D.A.; Barnard, E.A.; Richards, B.M. Expression of functional GABA, glycine and glutamate receptors in Xenopus oocytes injected with rat brain mRNA. Nature 1984, 310, 318–321. [Google Scholar] [CrossRef] [PubMed]
  55. Sumikawa, K.; Parker, I.; Miledi, R. Partial purification and functional expression of brain mRNAs coding for neurotransmitter receptors and voltage-operated channels. Proc. Natl. Acad. Sci. USA 1984, 81, 7994–7998. [Google Scholar] [CrossRef] [Green Version]
  56. Sumikawa, K.; Parker, I.; Miledi, R. Messenger RNA from rat brain induces noradrenaline and dopamine receptors in Xenopus oocytes. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1984, 223, 255–260. [Google Scholar] [CrossRef]
  57. Jardemark, K.; Nystrom, B.; Rydenhag, B.; Hamberger, A.; Jacobson, I. Expression of Ca(2+)-ion permeable alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) receptors in Xenopus oocytes injected with total RNA from human epileptic temporal lobe. Neurosci. Lett. 1995, 194, 93–96. [Google Scholar] [CrossRef]
  58. Spauschus, A.; Eunson, L.; Hanna, M.G.; Kullmann, D.M. Functional characterization of a novel mutation in KCNA1 in episodic ataxia type 1 associated with epilepsy. Ann. N. Y. Acad. Sci. 1999, 868, 442–446. [Google Scholar] [CrossRef] [PubMed]
  59. Yang, W.P.; Levesque, P.C.; Little, W.A.; Conder, M.L.; Ramakrishnan, P.; Neubauer, M.G.; Blanar, M.A. Functional expression of two KvLQT1-related potassium channels responsible for an inherited idiopathic epilepsy. J. Biol. Chem. 1998, 273, 19419–19423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Madeja, M.; Stocker, M.; Musshoff, U.; Pongs, O.; Speckmann, E.J. Potassium currents in epilepsy: Effects of the epileptogenic agent pentylenetetrazol on a cloned potassium channel. Brain Res. 1994, 656, 287–294. [Google Scholar] [CrossRef]
  61. Musshoff, U.; Madeja, M.; Bloms-Funke, P.; Speckmann, E.J. Effects of the epileptogenic agent bicuculline methiodide on membrane currents induced by N-methyl-D-aspartate and kainate (oocyte; Xenopus laevis). Brain Res. 1994, 639, 135–138. [Google Scholar] [CrossRef] [PubMed]
  62. Airaksinen, M.S.; Panula, P. Comparative neuroanatomy of the histaminergic system in the brain of the frog Xenopus laevis. J. Comp. Neurol. 1990, 292, 412–423. [Google Scholar] [CrossRef]
  63. Gupta, T.; Marquart, G.D.; Horstick, E.J.; Tabor, K.M.; Pajevic, S.; Burgess, H.A. Morphometric analysis and neuroanatomical mapping of the zebrafish brain. Methods 2018, 150, 49–62. [Google Scholar] [CrossRef]
  64. Herget, U.; Wolf, A.; Wullimann, M.F.; Ryu, S. Molecular neuroanatomy and chemoarchitecture of the neurosecretory preoptic-hypothalamic area in zebrafish larvae. J. Comp. Neurol. 2014, 522, 1542–1564. [Google Scholar] [CrossRef]
  65. Kunst, M.; Laurell, E.; Mokayes, N.; Kramer, A.; Kubo, F.; Fernandes, A.M.; Förster, D.; Dal Maschio, M.; Baier, H. A Cellular-Resolution Atlas of the Larval Zebrafish Brain. Neuron 2019, 103, 21–38 e25. [Google Scholar] [CrossRef] [PubMed]
  66. Panula, P.; Chen, Y.C.; Priyadarshini, M.; Kudo, H.; Semenova, S.; Sundvik, M.; Sallinen, V. The comparative neuroanatomy and neurochemistry of zebrafish CNS systems of relevance to human neuropsychiatric diseases. Neurobiol. Dis. 2010, 40, 46–57. [Google Scholar] [CrossRef]
  67. Ridd, K. Neuroanatomy: From fin to forelimb. Nature 2010, 466, 701. [Google Scholar] [CrossRef] [PubMed]
  68. Roberts, A.; Alford, S.T. Descending projections and excitation during fictive swimming in Xenopus embryos: Neuroanatomy and lesion experiments. J. Comp. Neurol. 1986, 250, 253–261. [Google Scholar] [CrossRef]
  69. Roberts, A.; Clarke, J.D. The neuroanatomy of an amphibian embryo spinal cord. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 1982, 296, 195–212. [Google Scholar] [CrossRef]
  70. Zhu, X.Y.; Guo, S.Y.; Xia, B.; Li, C.Q.; Wang, L.; Wang, Y.H. Development of zebrafish demyelination model for evaluation of remyelination compounds and RORgammat inhibitors. J. Pharmacol. Toxicol. Methods 2019, 98, 106585. [Google Scholar] [CrossRef] [PubMed]
  71. Fang, Y.; Lei, X.; Li, X.; Chen, Y.; Xu, F.; Feng, X.; Wei, S.; Li, Y. A novel model of demyelination and remyelination in a GFP-transgenic zebrafish. Biol. Open 2014, 4, 62–68. [Google Scholar] [CrossRef] [PubMed]
  72. Chung, A.Y.; Kim, P.S.; Kim, S.; Kim, E.; Kim, D.; Jeong, I.; Kim, H.K.; Ryu, J.H.; Kim, C.H.; Choi, J.; et al. Generation of demyelination models by targeted ablation of oligodendrocytes in the zebrafish CNS. Mol. Cells 2013, 36, 82–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Kaya, F.; Mannioui, A.; Chesneau, A.; Sekizar, S.; Maillard, E.; Ballagny, C.; Houel-Renault, L.; Dupasquier, D.; Bronchain, O.; Holtzmann, I.; et al. Live imaging of targeted cell ablation in Xenopus: A new model to study demyelination and repair. J. Neurosci. Off. J. Soc. Neurosci. 2012, 32, 12885–12895. [Google Scholar] [CrossRef] [Green Version]
  74. Sekizar, S.; Mannioui, A.; Azoyan, L.; Colin, C.; Thomas, J.L.; Du Pasquier, D.; Mallat, M.; Zalc, B. Remyelination by Resident Oligodendrocyte Precursor Cells in a Xenopus laevis Inducible Model of Demyelination. Dev. Neurosci. 2015, 37, 232–242. [Google Scholar] [CrossRef] [PubMed]
  75. Mannioui, A.; Zalc, B. Conditional Demyelination and Remyelination in a Transgenic Xenopus laevis. Methods Mol. Biol. 2019, 1936, 239–248. [Google Scholar] [CrossRef]
  76. Drenthen, G.S.; Fonseca Wald, E.L.A.; Backes, W.H.; Debeij-Van Hall, M.; Hendriksen, J.G.M.; Aldenkamp, A.P.; Vermeulen, R.J.; Klinkenberg, S.; Jansen, J.F.A. Lower myelin-water content of the frontal lobe in childhood absence epilepsy. Epilepsia 2019, 60, 1689–1696. [Google Scholar] [CrossRef] [Green Version]
  77. Bencurova, P.; Laakso, H.; Salo, R.A.; Paasonen, E.; Manninen, E.; Paasonen, J.; Michaeli, S.; Mangia, S.; Bares, M.; Brazdil, M.; et al. Infantile status epilepticus disrupts myelin development. Neurobiol. Dis. 2022, 162, 105566. [Google Scholar] [CrossRef] [PubMed]
  78. de Curtis, M.; Garbelli, R.; Uva, L. A hypothesis for the role of axon demyelination in seizure generation. Epilepsia 2021, 62, 583–595. [Google Scholar] [CrossRef] [PubMed]
  79. Reith, C.A.; Sillar, K.T. Development and role of GABA(A) receptor-mediated synaptic potentials during swimming in postembryonic Xenopus laevis tadpoles. J. Neurophysiol. 1999, 82, 3175–3187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Schmidt, C.; Hollmann, M. Molecular and functional characterization of Xenopus laevis N-methyl-d-aspartate receptors. Mol. Cell. Neurosci. 2009, 42, 116–127. [Google Scholar] [CrossRef] [PubMed]
  81. Schmidt, C.; Klein, C.; Hollmann, M. Xenopus laevis oocytes endogenously express all subunits of the ionotropic glutamate receptor family. J. Mol. Biol. 2009, 390, 182–195. [Google Scholar] [CrossRef]
  82. Cox, J.A.; Kucenas, S.; Voigt, M.M. Molecular characterization and embryonic expression of the family of N-methyl-D-aspartate receptor subunit genes in the zebrafish. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2005, 234, 756–766. [Google Scholar] [CrossRef] [PubMed]
  83. Wahlstrom-Helgren, S.; Montgomery, J.E.; Vanpelt, K.T.; Biltz, S.L.; Peck, J.H.; Masino, M.A. Glutamate receptor subtypes differentially contribute to optogenetically activated swimming in spinally transected zebrafish larvae. J. Neurophysiol. 2019, 122, 2414–2426. [Google Scholar] [CrossRef]
  84. Turrini, L.; Fornetto, C.; Marchetto, G.; Müllenbroich, M.C.; Tiso, N.; Vettori, A.; Resta, F.; Masi, A.; Mannaioni, G.; Pavone, F.S.; et al. Optical mapping of neuronal activity during seizures in zebrafish. Sci. Rep. 2017, 7, 3025. [Google Scholar] [CrossRef] [Green Version]
  85. Rosch, R.E.; Hunter, P.R.; Baldeweg, T.; Friston, K.J.; Meyer, M.P. Calcium imaging and dynamic causal modelling reveal brain-wide changes in effective connectivity and synaptic dynamics during epileptic seizures. PLoS Comput. Biol. 2018, 14, e1006375. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, J.; Baraban, S.C. Network Properties Revealed during Multi-Scale Calcium Imaging of Seizure Activity in Zebrafish. eNeuro 2019, 6. [Google Scholar] [CrossRef] [Green Version]
  87. Burrows, D.R.W.; Samarut, É.; Liu, J.; Baraban, S.C.; Richardson, M.P.; Meyer, M.P.; Rosch, R.E. Imaging epilepsy in larval zebrafish. Eur. J. Paediatr. Neurol. EJPN Off. J. Eur. Paediatr. Neurol. Soc. 2020, 24, 70–80. [Google Scholar] [CrossRef] [Green Version]
  88. Zheng, J.; Hsieh, F.; Ge, L. A Data-Driven Approach to Predict and Classify Epileptic Seizures from Brain-Wide Calcium Imaging Video Data. IEEE/ACM Trans. Comput. Biol. Bioinform. 2020, 17, 1858–1870. [Google Scholar] [CrossRef]
  89. Niemeyer, J.E.; Gadamsetty, P.; Chun, C.; Sylvester, S.; Lucas, J.P.; Ma, H.; Schwartz, T.H.; Aksay, E.R.F. Seizures initiate in zones of relative hyperexcitation in a zebrafish epilepsy model. Brain A J. Neurol. 2022. [Google Scholar] [CrossRef]
  90. Cline, H.T. Imaging Structural and Functional Dynamics in Xenopus Neurons. Cold Spring Harb. Protoc. 2022, 2022, top106773. [Google Scholar] [CrossRef] [PubMed]
  91. Li, V.J.; Schohl, A.; Ruthazer, E.S. Topographic map formation and the effects of NMDA receptor blockade in the developing visual system. Proc. Natl. Acad. Sci. USA 2022, 119, e2107899119. [Google Scholar] [CrossRef]
  92. Banote, R.K.; Koutarapu, S.; Chennubhotla, K.S.; Chatti, K.; Kulkarni, P. Oral gabapentin suppresses pentylenetetrazole-induced seizure-like behavior and cephalic field potential in adult zebrafish. Epilepsy Behav. EB 2013, 27, 212–219. [Google Scholar] [CrossRef] [PubMed]
  93. Meyer, M.; Dhamne, S.C.; LaCoursiere, C.M.; Tambunan, D.; Poduri, A.; Rotenberg, A. Microarray Noninvasive Neuronal Seizure Recordings from Intact Larval Zebrafish. PLoS ONE 2016, 11, e0156498. [Google Scholar] [CrossRef] [Green Version]
  94. Bandara, S.B.; Carty, D.R.; Singh, V.; Harvey, D.J.; Vasylieva, N.; Pressly, B.; Wulff, H.; Lein, P.J. Susceptibility of larval zebrafish to the seizurogenic activity of GABA type A receptor antagonists. Neurotoxicology 2020, 76, 220–234. [Google Scholar] [CrossRef] [PubMed]
  95. Cozzolino, O.; Sicca, F.; Paoli, E.; Trovato, F.; Santorelli, F.M.; Ratto, G.M.; Marchese, M. Evolution of Epileptiform Activity in Zebrafish by Statistical-Based Integration of Electrophysiology and 2-Photon Ca(2+) Imaging. Cells 2020, 9, 769. [Google Scholar] [CrossRef] [Green Version]
  96. Hadjiabadi, D.; Lovett-Barron, M.; Raikov, I.G.; Sparks, F.T.; Liao, Z.; Baraban, S.C.; Leskovec, J.; Losonczy, A.; Deisseroth, K.; Soltesz, I. Maximally selective single-cell target for circuit control in epilepsy models. Neuron 2021, 109, 2556–2572.e2556. [Google Scholar] [CrossRef]
  97. Baraban, S.C.; Taylor, M.R.; Castro, P.A.; Baier, H. Pentylenetetrazole induced changes in zebrafish behavior, neural activity and c-fos expression. Neuroscience 2005, 131, 759–768. [Google Scholar] [CrossRef] [PubMed]
  98. Hewapathirane, D.S.; Dunfield, D.; Yen, W.; Chen, S.; Haas, K. In vivo imaging of seizure activity in a novel developmental seizure model. Exp Neurol 2008, 211, 480–488. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, J.; Lin, Z.J.; Liu, L.; Xu, H.Q.; Shi, Y.W.; Yi, Y.H.; He, N.; Liao, W.P. Epilepsy-associated genes. Seizure 2017, 44, 11–20. [Google Scholar] [CrossRef] [Green Version]
  100. Cárdenas-Rodríguez, N.; Carmona-Aparicio, L.; Pérez-Lozano, D.L.; Ortega-Cuellar, D.; Gómez-Manzo, S.; Ignacio-Mejía, I. Genetic variations associated with pharmacoresistant epilepsy (Review). Mol. Med. Rep. 2020, 21, 1685–1701. [Google Scholar] [CrossRef] [PubMed]
  101. Pauli, A.; Montague, T.G.; Lennox, K.A.; Behlke, M.A.; Schier, A.F. Antisense Oligonucleotide-Mediated Transcript Knockdown in Zebrafish. PLoS ONE 2015, 10, e0139504. [Google Scholar] [CrossRef] [Green Version]
  102. Dubińska-Magiera, M.; Chmielewska, M.; Kozioł, K.; Machowska, M.; Hutchison, C.J.; Goldberg, M.W.; Rzepecki, R. Xenopus LAP2β protein knockdown affects location of lamin B and nucleoporins and has effect on assembly of cell nucleus and cell viability. Protoplasma 2016, 253, 943–956. [Google Scholar] [CrossRef] [Green Version]
  103. Crossley, M.P.; Krude, T. Targeting Functional Noncoding RNAs. Methods Mol. Biol. 2017, 1565, 151–160. [Google Scholar] [CrossRef] [PubMed]
  104. Stainier, D.Y.R.; Raz, E.; Lawson, N.D.; Ekker, S.C.; Burdine, R.D.; Eisen, J.S.; Ingham, P.W.; Schulte-Merker, S.; Yelon, D.; Weinstein, B.M.; et al. Guidelines for morpholino use in zebrafish. PLoS Genet. 2017, 13, e1007000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kim, J.; Clark, K.; Barton, C.; Tanguay, R.; Moulton, H. A Novel Zebrafish Model for Assessing In Vivo Delivery of Morpholino Oligomers. Methods Mol. Biol. 2018, 1828, 293–306. [Google Scholar] [CrossRef]
  106. Paraiso, K.D.; Blitz, I.L.; Zhou, J.J.; Cho, K.W.Y. Morpholinos Do Not Elicit an Innate Immune Response during Early Xenopus Embryogenesis. Dev. Cell 2019, 49, 643–650 e643. [Google Scholar] [CrossRef] [PubMed]
  107. Van Gils, M.; Vanakker, O.M. Morpholino-Mediated Gene Knockdown in Zebrafish: It Is All About Dosage and Validation. J. Investig. Dermatol. 2019, 139, 1599–1600. [Google Scholar] [CrossRef]
  108. Iida, M.; Suzuki, M.; Sakane, Y.; Nishide, H.; Uchiyama, I.; Yamamoto, T.; Suzuki, K.T.; Fujii, S. A simple and practical workflow for genotyping of CRISPR-Cas9-based knockout phenotypes using multiplexed amplicon sequencing. Genes Cells Devoted Mol. Cell. Mech. 2020, 25, 498–509. [Google Scholar] [CrossRef] [PubMed]
  109. Kim, B.H.; Zhang, G. Generating Stable Knockout Zebrafish Lines by Deleting Large Chromosomal Fragments Using Multiple gRNAs. G3 (Bethesda Md.) 2020, 10, 1029–1037. [Google Scholar] [CrossRef] [Green Version]
  110. Kroll, F.; Powell, G.T.; Ghosh, M.; Gestri, G.; Antinucci, P.; Hearn, T.J.; Tunbak, H.; Lim, S.; Dennis, H.W.; Fernandez, J.M.; et al. A simple and effective F0 knockout method for rapid screening of behaviour and other complex phenotypes. eLife 2021, 10, e59683. [Google Scholar] [CrossRef]
  111. Zhang, C.; Li, J.; Tarique, I.; Zhang, Y.; Lu, T.; Wang, J.; Chen, A.; Wen, F.; Zhang, Z.; Zhang, Y.; et al. A Time-Saving Strategy to Generate Double Maternal Mutants by an Oocyte-Specific Conditional Knockout System in Zebrafish. Biology 2021, 10, 777. [Google Scholar] [CrossRef] [PubMed]
  112. Aksoy, Y.A.; Yang, B.; Chen, W.; Hung, T.; Kuchel, R.P.; Zammit, N.W.; Grey, S.T.; Goldys, E.M.; Deng, W. Spatial and Temporal Control of CRISPR-Cas9-Mediated Gene Editing Delivered via a Light-Triggered Liposome System. ACS Appl. Mater. Interfaces 2020, 12, 52433–52444. [Google Scholar] [CrossRef]
  113. Quick, R.E.; Buck, L.D.; Parab, S.; Tolbert, Z.R.; Matsuoka, R.L. Highly Efficient Synthetic CRISPR RNA/Cas9-Based Mutagenesis for Rapid Cardiovascular Phenotypic Screening in F0 Zebrafish. Front. Cell Dev. Biol. 2021, 9, 735598. [Google Scholar] [CrossRef] [PubMed]
  114. Sharma, P.; Sharma, B.S.; Verma, R.J. CRISPR-based genome editing of zebrafish. Prog. Mol. Biol. Transl. Sci. 2021, 180, 69–84. [Google Scholar] [CrossRef] [PubMed]
  115. Uribe-Salazar, J.M.; Kaya, G.; Sekar, A.; Weyenberg, K.; Ingamells, C.; Dennis, M.Y. Evaluation of CRISPR gene-editing tools in zebrafish. BMC Genom. 2022, 23, 12. [Google Scholar] [CrossRef]
  116. Zhang, C.; Lu, T.; Zhang, Y.; Li, J.; Tarique, I.; Wen, F.; Chen, A.; Wang, J.; Zhang, Z.; Zhang, Y.; et al. Rapid generation of maternal mutants via oocyte transgenic expression of CRISPR-Cas9 and sgRNAs in zebrafish. Sci. Adv. 2021, 7, eabg4243. [Google Scholar] [CrossRef] [PubMed]
  117. Abdelrahman, D.; Hasan, W.; Da’as, S.I. Microinjection quality control in zebrafish model for genetic manipulations. MethodsX 2021, 8, 101418. [Google Scholar] [CrossRef]
  118. Lane, M.; Mis, E.K.; Khokha, M.K. Microinjection of Xenopus tropicalis Embryos. Cold Spring Harb. Protoc. 2022, 2022, prot107644. [Google Scholar] [CrossRef]
  119. Moody, S.A. Microinjection of mRNAs and Oligonucleotides. Cold Spring Harb. Protoc. 2018, 2018. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, C.; Szaro, B.G. A method for using direct injection of plasmid DNA to study cis-regulatory element activity in F0 Xenopus embryos and tadpoles. Dev. Biol. 2015, 398, 11–23. [Google Scholar] [CrossRef] [Green Version]
  121. Yasuoka, Y.; Taira, M. Microinjection of DNA Constructs into Xenopus Embryos for Gene Misexpression and cis-Regulatory Module Analysis. Cold Spring Harb. Protoc. 2019, 2019. [Google Scholar] [CrossRef]
  122. Cho, S.J.; Park, E.; Baker, A.; Reid, A.Y. Post-Traumatic Epilepsy in Zebrafish Is Drug-Resistant and Impairs Cognitive Function. J. Neurotrauma 2021, 38, 3174–3183. [Google Scholar] [CrossRef]
  123. Kumari, S.; Sharma, P.; Mazumder, A.G.; Rana, A.K.; Sharma, S.; Singh, D. Development and validation of chemical kindling in adult zebrafish: A simple and improved chronic model for screening of antiepileptic agents. J. Neurosci. Methods 2020, 346, 108916. [Google Scholar] [CrossRef]
  124. Mussulini, B.H.; Leite, C.E.; Zenki, K.C.; Moro, L.; Baggio, S.; Rico, E.P.; Rosemberg, D.B.; Dias, R.D.; Souza, T.M.; Calcagnotto, M.E.; et al. Seizures induced by pentylenetetrazole in the adult zebrafish: A detailed behavioral characterization. PLoS ONE 2013, 8, e54515. [Google Scholar] [CrossRef]
  125. Samarut, E.; Swaminathan, A.; Riche, R.; Liao, M.; Hassan-Abdi, R.; Renault, S.; Allard, M.; Dufour, L.; Cossette, P.; Soussi-Yanicostas, N.; et al. gamma-Aminobutyric acid receptor alpha 1 subunit loss of function causes genetic generalized epilepsy by impairing inhibitory network neurodevelopment. Epilepsia 2018, 59, 2061–2074. [Google Scholar] [CrossRef] [Green Version]
  126. Lee, Y.; Lee, K.J.; Jang, J.W.; Lee, S.I.; Kim, S. An EEG system to detect brain signals from multiple adult zebrafish. Biosens. Bioelectron. 2020, 164, 112315. [Google Scholar] [CrossRef]
  127. Cassar, S.; Breidenbach, L.; Olson, A.; Huang, X.; Britton, H.; Woody, C.; Sancheti, P.; Stolarik, D.; Wicke, K.; Hempel, K.; et al. Measuring drug absorption improves interpretation of behavioral responses in a larval zebrafish locomotor assay for predicting seizure liability. J. Pharmacol. Toxicol. Methods 2017, 88, 56–63. [Google Scholar] [CrossRef]
  128. Van Wijk, R.C.; Krekels, E.H.J.; Kantae, V.; Ordas, A.; Kreling, T.; Harms, A.C.; Hankemeier, T.; Spaink, H.P.; van der Graaf, P.H. Mechanistic and Quantitative Understanding of Pharmacokinetics in Zebrafish Larvae through Nanoscale Blood Sampling and Metabolite Modeling of Paracetamol. J. Pharmacol. Exp. Ther. 2019, 371, 15–24. [Google Scholar] [CrossRef] [PubMed]
  129. Howard, A.M.; Papich, M.G.; Felt, S.A.; Long, C.T.; McKeon, G.P.; Bond, E.S.; Torreilles, S.L.; Luong, R.H.; Green, S.L. The pharmacokinetics of enrofloxacin in adult African clawed frogs (Xenopus laevis). J. Am. Assoc. Lab. Anim. Sci. JAALAS 2010, 49, 800–804. [Google Scholar]
  130. Fleming, A.; Diekmann, H.; Goldsmith, P. Functional characterisation of the maturation of the blood-brain barrier in larval zebrafish. PLoS ONE 2013, 8, e77548. [Google Scholar] [CrossRef]
  131. Johnston, G.A. Advantages of an antagonist: Bicuculline and other GABA antagonists. Br. J. Pharmacol. 2013, 169, 328–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Krishek, B.J.; Moss, S.J.; Smart, T.G. A functional comparison of the antagonists bicuculline and picrotoxin at recombinant GABAA receptors. Neuropharmacology 1996, 35, 1289–1298. [Google Scholar] [CrossRef]
  133. Simmonds, M.A. Evidence that bicuculline and picrotoxin act at separate sites to antagonize gamma-aminobutyric acid in rat cuneate nucleus. Neuropharmacology 1980, 19, 39–45. [Google Scholar] [CrossRef]
  134. de Feo, M.R.; Mecarelli, O.; Ricci, G.F. Bicuculline- and allylglycine-induced epilepsy in developing rats. Exp. Neurol. 1985, 90, 411–421. [Google Scholar] [CrossRef]
  135. Jones, A.; Barker-Haliski, M.; Ilie, A.S.; Herd, M.B.; Baxendale, S.; Holdsworth, C.J.; Ashton, J.P.; Placzek, M.; Jayasekera, B.A.P.; Cowie, C.J.A.; et al. A multiorganism pipeline for antiseizure drug discovery: Identification of chlorothymol as a novel γ-aminobutyric acidergic anticonvulsant. Epilepsia 2020, 61, 2106–2118. [Google Scholar] [CrossRef] [PubMed]
  136. Murao, K.; Shingu, K.; Miyamoto, E.; Ikeda, S.; Nakao, S.; Masuzawa, M.; Yamada, M. Anticonvulsant effects of sevoflurane on amygdaloid kindling and bicuculline-induced seizures in cats: Comparison with isoflurane and halothane. J. Anesth. 2002, 16, 34–43. [Google Scholar] [CrossRef] [PubMed]
  137. Soares-Silva, B.; Beserra-Filho, J.I.A.; Morera, P.M.A.; Custódio-Silva, A.C.; Maria-Macêdo, A.; Silva-Martins, S.; Alexandre-Silva, V.; Silva, S.P.; Silva, R.H.; Ribeiro, A.M. The bee venom active compound melittin protects against bicuculline-induced seizures and hippocampal astrocyte activation in rats. Neuropeptides 2022, 91, 102209. [Google Scholar] [CrossRef]
  138. Vicente-Silva, W.; Silva-Freitas, F.R.; Beserra-Filho, J.I.A.; Cardoso, G.N.; Silva-Martins, S.; Sarno, T.A.; Silva, S.P.; Soares-Silva, B.; Dos Santos, J.R.; da Silva, R.H.; et al. Sakuranetin exerts anticonvulsant effect in bicuculline-induced seizures. Fundam. Clin. Pharmacol. 2022. [Google Scholar] [CrossRef]
  139. Wong, K.; Stewart, A.; Gilder, T.; Wu, N.; Frank, K.; Gaikwad, S.; Suciu, C.; Dileo, J.; Utterback, E.; Chang, K.; et al. Modeling seizure-related behavioral and endocrine phenotypes in adult zebrafish. Brain Res. 2010, 1348, 209–215. [Google Scholar] [CrossRef] [PubMed]
  140. Stankevicius, D.; Rodrigues-Costa, E.C.; Camilo Florio, J.; Palermo-Neto, J. Neuroendocrine, behavioral and macrophage activity changes induced by picrotoxin effects in mice. Neuropharmacology 2008, 54, 300–308. [Google Scholar] [CrossRef]
  141. Acharya, M.M.; Khamesra, S.H.; Katyare, S.S. Picrotoxin-induced convulsions and lysosomal function in the rat brain. Indian J. Clin. Biochem. IJCB 2005, 20, 56–60. [Google Scholar] [CrossRef] [Green Version]
  142. Acharya, M.M.; Khamesra, S.H.; Katyare, S.S. Effect of repeated intraperitoneal exposure to picrotoxin on rat liver lysosomal function. Indian J. Exp. Biol. 2004, 42, 808–811. [Google Scholar]
  143. Li, J.M.; Gan, J.; Zeng, T.F.; Sander, J.W.; Zhou, D. Tetramethylenedisulfotetramine intoxication presenting with de novo Status Epilepticus: A case series. Neurotoxicology 2012, 33, 207–211. [Google Scholar] [CrossRef] [PubMed]
  144. Mundy, P.C.; Pressly, B.; Carty, D.R.; Yaghoobi, B.; Wulff, H.; Lein, P.J. The efficacy of gamma-aminobutyric acid type A receptor (GABA AR) subtype-selective positive allosteric modulators in blocking tetramethylenedisulfotetramine (TETS)-induced seizure-like behavior in larval zebrafish with minimal sedation. Toxicol. Appl. Pharmacol. 2021, 426, 115643. [Google Scholar] [CrossRef]
  145. Maroso, M.; Balosso, S.; Ravizza, T.; Liu, J.; Aronica, E.; Iyer, A.M.; Rossetti, C.; Molteni, M.; Casalgrandi, M.; Manfredi, A.A.; et al. Toll-like receptor 4 and high-mobility group box-1 are involved in ictogenesis and can be targeted to reduce seizures. Nat. Med. 2010, 16, 413–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Bragin, A.; Azizyan, A.; Almajano, J.; Wilson, C.L.; Engel, J., Jr. Analysis of chronic seizure onsets after intrahippocampal kainic acid injection in freely moving rats. Epilepsia 2005, 46, 1592–1598. [Google Scholar] [CrossRef]
  147. Bragin, A.; Wilson, C.L.; Engel, J., Jr. Chronic epileptogenesis requires development of a network of pathologically interconnected neuron clusters: A hypothesis. Epilepsia 2000, 41 (Suppl. S6), S144–S152. [Google Scholar] [CrossRef] [Green Version]
  148. Zupanc, G.K.; Hinsch, K.; Gage, F.H. Proliferation, migration, neuronal differentiation, and long-term survival of new cells in the adult zebrafish brain. J. Comp. Neurol. 2005, 488, 290–319. [Google Scholar] [CrossRef] [PubMed]
  149. Heylen, L.; Pham, D.H.; De Meulemeester, A.S.; Samarut, E.; Skiba, A.; Copmans, D.; Kazwiny, Y.; Vanden Berghe, P.; de Witte, P.A.M.; Siekierska, A. Pericardial Injection of Kainic Acid Induces a Chronic Epileptic State in Larval Zebrafish. Front. Mol. Neurosci. 2021, 14, 753936. [Google Scholar] [CrossRef]
  150. Paudel, Y.N.; Kumari, Y.; Abidin, S.A.Z.; Othman, I.; Shaikh, M.F. Pilocarpine Induced Behavioral and Biochemical Alterations in Chronic Seizure-Like Condition in Adult Zebrafish. Int J. Mol. Sci 2020, 21, 2492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Paudel, Y.N.; Othman, I.; Shaikh, M.F. Anti-High Mobility Group Box-1 Monoclonal Antibody Attenuates Seizure-Induced Cognitive Decline by Suppressing Neuroinflammation in an Adult Zebrafish Model. Front. Pharm. 2020, 11, 613009. [Google Scholar] [CrossRef] [PubMed]
  152. Heuzeroth, H.; Wawra, M.; Fidzinski, P.; Dag, R.; Holtkamp, M. The 4-Aminopyridine Model of Acute Seizures in vitro Elucidates Efficacy of New Antiepileptic Drugs. Front. Neurosci. 2019, 13, 677. [Google Scholar] [CrossRef] [Green Version]
  153. Gupta, R. Encyclopedia of Toxicology; Academic Press: Cambridge, MA, USA, 2014. [Google Scholar]
  154. Blight, A.R.; Henney, H.R., 3rd; Cohen, R. Development of dalfampridine, a novel pharmacologic approach for treating walking impairment in multiple sclerosis. Ann. N. Y. Acad. Sci. 2014, 1329, 33–44. [Google Scholar] [CrossRef]
  155. DeForge, D.; Nymark, J.; Lemaire, E.; Gardner, S.; Hunt, M.; Martel, L.; Curran, D.; Barbeau, H. Effect of 4-aminopyridine on gait in ambulatory spinal cord injuries: A double-blind, placebo-controlled, crossover trial. Spinal Cord 2004, 42, 674–685. [Google Scholar] [CrossRef] [Green Version]
  156. Hayes, K.C. The use of 4-aminopyridine (fampridine) in demyelinating disorders. CNS Drug Rev. 2004, 10, 295–316. [Google Scholar] [CrossRef]
  157. Rahimi, N.; Modabberi, S.; Faghir-Ghanesefat, H.; Shayan, M.; Farzad Maroufi, S.; Asgari Dafe, E.; Reza Dehpour, A. The possible role of nitric oxide signaling and NMDA receptors in allopurinol effect on maximal electroshock- and pentylenetetrazol-induced seizures in mice. Neurosci. Lett. 2022, 778, 136620. [Google Scholar] [CrossRef]
  158. van der Linde, H.; Kreir, M.; Teisman, A.; Gallacher, D.J. Seizure-induced Torsades de pointes: In a canine drug-induced long-QT1 model. J. Pharmacol. Toxicol. Methods 2021, 111, 107086. [Google Scholar] [CrossRef]
  159. Nagata, S.; Fujiwara, K.; Kuga, K.; Ozaki, H. Prediction of GABA receptor antagonist-induced convulsion in cynomolgus monkeys by combining machine learning and heart rate variability analysis. J. Pharmacol. Toxicol. Methods 2021, 112, 107127. [Google Scholar] [CrossRef] [PubMed]
  160. Loscher, W. Critical review of current animal models of seizures and epilepsy used in the discovery and development of new antiepileptic drugs. Seizure 2011, 20, 359–368. [Google Scholar] [CrossRef] [Green Version]
  161. Alachkar, A.; Ojha, S.K.; Sadeq, A.; Adem, A.; Frank, A.; Stark, H.; Sadek, B. Experimental Models for the Discovery of Novel Anticonvulsant Drugs: Focus on Pentylenetetrazole-Induced Seizures and Associated Memory Deficits. Curr. Pharm. Des. 2020, 26, 1693–1711. [Google Scholar] [CrossRef]
  162. Chipiti, T.; Viljoen, A.M.; Cordero-Maldonado, M.L.; Veale, C.G.L.; Van Heerden, F.R.; Sandasi, M.; Chen, W.; Crawford, A.D.; Enslin, G.M. Anti-seizure activity of African medicinal plants: The identification of bioactive alkaloids from the stem bark of Rauvolfia caffra using an in vivo zebrafish model. J. Ethnopharmacol. 2021, 279, 114282. [Google Scholar] [CrossRef]
  163. Copmans, D.; Orellana-Paucar, A.M.; Steurs, G.; Zhang, Y.; Ny, A.; Foubert, K.; Exarchou, V.; Siekierska, A.; Kim, Y.; De Borggraeve, W.; et al. Methylated flavonoids as anti-seizure agents: Naringenin 4′,7-dimethyl ether attenuates epileptic seizures in zebrafish and mouse models. Neurochem. Int. 2018, 112, 124–133. [Google Scholar] [CrossRef] [Green Version]
  164. Copmans, D.; Rateb, M.; Tabudravu, J.N.; Perez-Bonilla, M.; Dirkx, N.; Vallorani, R.; Diaz, C.; Perez Del Palacio, J.; Smith, A.J.; Ebel, R.; et al. Zebrafish-Based Discovery of Antiseizure Compounds from the Red Sea: Pseurotin A2 and Azaspirofuran A. ACS Chem. Neurosci. 2018, 9, 1652–1662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Koziol, E.; Jozwiak, K.; Budzynska, B.; de Witte, P.A.M.; Copmans, D.; Skalicka-Wozniak, K. Comparative Antiseizure Analysis of Diverse Natural Coumarin Derivatives in Zebrafish. Int. J. Mol. Sci. 2021, 22, 1420. [Google Scholar] [CrossRef]
  166. Choo, B.K.M.; Kundap, U.P.; Kumari, Y.; Hue, S.M.; Othman, I.; Shaikh, M.F. Orthosiphon stamineus Leaf Extract Affects TNF-alpha and Seizures in a Zebrafish Model. Front. Pharm. 2018, 9, 139. [Google Scholar] [CrossRef] [Green Version]
  167. Xia, X.; Vishwanath, M.; Zhang, J.; Sarafan, S.; Trigo Torres, R.S.; Le, T.; Lau, M.P.H.; Nguyen, A.H.; Cao, H. Microelectrode array membranes to simultaneously assess cardiac and neurological signals of xenopus laevis under chemical exposures and environmental changes. Biosens. Bioelectron. 2022, 210, 114292. [Google Scholar] [CrossRef] [PubMed]
  168. Teng, Y.; Xie, X.; Walker, S.; Saxena, M.; Kozlowski, D.J.; Mumm, J.S.; Cowell, J.K. Loss of zebrafish lgi1b leads to hydrocephalus and sensitization to pentylenetetrazol induced seizure-like behavior. PLoS ONE 2011, 6, e24596. [Google Scholar] [CrossRef]
  169. Johnson, M.R.; Behmoaras, J.; Bottolo, L.; Krishnan, M.L.; Pernhorst, K.; Santoscoy, P.L.M.; Rossetti, T.; Speed, D.; Srivastava, P.K.; Chadeau-Hyam, M.; et al. Systems genetics identifies Sestrin 3 as a regulator of a proconvulsant gene network in human epileptic hippocampus. Nat. Commun. 2015, 6, 6031. [Google Scholar] [CrossRef] [Green Version]
  170. Mei, X.; Wu, S.; Bassuk, A.G.; Slusarski, D.C. Mechanisms of prickle1a function in zebrafish epilepsy and retinal neurogenesis. Dis. Models Mech. 2013, 6, 679–688. [Google Scholar] [CrossRef] [Green Version]
  171. Baraban, S.C. A zebrafish-centric approach to antiepileptic drug development. Dis. Models Mech. 2021, 14. [Google Scholar] [CrossRef]
  172. Baraban, S.C.; Dinday, M.T.; Hortopan, G.A. Drug screening in Scn1a zebrafish mutant identifies clemizole as a potential Dravet syndrome treatment. Nat. Commun. 2013, 4, 2410. [Google Scholar] [CrossRef] [Green Version]
  173. Moog, M.; Baraban, S.C. Clemizole and Trazodone are Effective Antiseizure Treatments in a Zebrafish Model of STXBP1 Disorder. Epilepsia Open 2022. [Google Scholar] [CrossRef]
  174. Xu, Y.; Song, R.; Chen, W.; Strong, K.; Shrey, D.; Gedela, S.; Traynelis, S.F.; Zhang, G.; Yuan, H. Recurrent seizure-related GRIN1 variant: Molecular mechanism and targeted therapy. Ann. Clin. Transl. Neurol. 2021, 8, 1480–1494. [Google Scholar] [CrossRef] [PubMed]
  175. Sega, A.G.; Mis, E.K.; Lindstrom, K.; Mercimek-Andrews, S.; Ji, W.; Cho, M.T.; Juusola, J.; Konstantino, M.; Jeffries, L.; Khokha, M.K.; et al. De novo pathogenic variants in neuronal differentiation factor 2 (NEUROD2) cause a form of early infantile epileptic encephalopathy. J. Med. Genet. 2019, 56, 113–122. [Google Scholar] [CrossRef] [PubMed]
  176. Kobayashi, T.; Hirai, H.; Iino, M.; Fuse, I.; Mitsumura, K.; Washiyama, K.; Kasai, S.; Ikeda, K. Inhibitory effects of the antiepileptic drug ethosuximide on G protein-activated inwardly rectifying K+ channels. Neuropharmacology 2009, 56, 499–506. [Google Scholar] [CrossRef] [PubMed]
  177. Ottosson, N.E.; Silvera Ejneby, M.; Wu, X.; Estrada-Mondragon, A.; Nilsson, M.; Karlsson, U.; Schupp, M.; Rognant, S.; Jepps, T.A.; Konradsson, P.; et al. Synthetic resin acid derivatives selectively open the hKV 7.2/7.3 channel and prevent epileptic seizures. Epilepsia 2021, 62, 1744–1758. [Google Scholar] [CrossRef] [PubMed]
  178. Stadler, M.; Monticelli, S.; Seidel, T.; Luger, D.; Salzer, I.; Boehm, S.; Holzer, W.; Schwarzer, C.; Urban, E.; Khom, S.; et al. Design, Synthesis, and Pharmacological Evaluation of Novel beta2/3 Subunit-Selective gamma-Aminobutyric Acid Type A (GABAA) Receptor Modulators. J. Med. Chem. 2019, 62, 317–341. [Google Scholar] [CrossRef]
  179. Kobayashi, T.; Ikeda, K.; Kumanishi, T. Inhibition by various antipsychotic drugs of the G-protein-activated inwardly rectifying K(+) (GIRK) channels expressed in xenopus oocytes. Br. J. Pharmacol. 2000, 129, 1716–1722. [Google Scholar] [CrossRef] [Green Version]
  180. Anderson, L.L.; Absalom, N.L.; Abelev, S.V.; Low, I.K.; Doohan, P.T.; Martin, L.J.; Chebib, M.; McGregor, I.S.; Arnold, J.C. Coadministered cannabidiol and clobazam: Preclinical evidence for both pharmacodynamic and pharmacokinetic interactions. Epilepsia 2019, 60, 2224–2234. [Google Scholar] [CrossRef] [Green Version]
  181. Bell, M.R.; Belarde, J.A.; Johnson, H.F.; Aizenman, C.D. A neuroprotective role for polyamines in a Xenopus tadpole model of epilepsy. Nat. Neurosci. 2011, 14, 505–512. [Google Scholar] [CrossRef]
  182. Najm, I.; el-Skaf, G.; Tocco, G.; Vanderklish, P.; Lynch, G.; Baudry, M. Seizure activity-induced changes in polyamine metabolism and neuronal pathology during the postnatal period in rat brain. Brain Res. Dev. Brain Res. 1992, 69, 11–21. [Google Scholar] [CrossRef]
  183. Ibhazehiebo, K.; Gavrilovici, C.; de la Hoz, C.L.; Ma, S.C.; Rehak, R.; Kaushik, G.; Meza Santoscoy, P.L.; Scott, L.; Nath, N.; Kim, D.Y.; et al. A novel metabolism-based phenotypic drug discovery platform in zebrafish uncovers HDACs 1 and 3 as a potential combined anti-seizure drug target. Brain J. Neurol. 2018, 141, 744–761. [Google Scholar] [CrossRef] [Green Version]
  184. Hentig, J.; Campbell, L.J.; Cloghessy, K.; Lee, M.; Boggess, W.; Hyde, D.R. Prophylactic Activation of Shh Signaling Attenuates TBI-Induced Seizures in Zebrafish by Modulating Glutamate Excitotoxicity through Eaat2a. Biomedicines 2021, 10, 32. [Google Scholar] [CrossRef] [PubMed]
  185. Shimizu, Y.; Kawasaki, T. Stab Wound Injury Model of the Adult Optic Tectum using Zebrafish and Medaka for the Comparative Analysis of Regenerative Capacity. J. Vis. Exp. JoVE 2022. [Google Scholar] [CrossRef] [PubMed]
  186. Uemura, N.; Koike, M.; Ansai, S.; Kinoshita, M.; Ishikawa-Fujiwara, T.; Matsui, H.; Naruse, K.; Sakamoto, N.; Uchiyama, Y.; Todo, T.; et al. Viable neuronopathic Gaucher disease model in Medaka (Oryzias latipes) displays axonal accumulation of alpha-synuclein. PLoS Genet. 2015, 11, e1005065. [Google Scholar] [CrossRef]
Figure 1. Classification of seizures according to the International League Against Epilepsy.
Figure 1. Classification of seizures according to the International League Against Epilepsy.
Ijms 23 08608 g001
Figure 2. Simplified schematic of chemical inducing agents used to model seizures in aquatic vertebrates.
Figure 2. Simplified schematic of chemical inducing agents used to model seizures in aquatic vertebrates.
Ijms 23 08608 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Williams, R.E.; Mruk, K. Aquatic Freshwater Vertebrate Models of Epilepsy Pathology: Past Discoveries and Future Directions for Therapeutic Discovery. Int. J. Mol. Sci. 2022, 23, 8608. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23158608

AMA Style

Williams RE, Mruk K. Aquatic Freshwater Vertebrate Models of Epilepsy Pathology: Past Discoveries and Future Directions for Therapeutic Discovery. International Journal of Molecular Sciences. 2022; 23(15):8608. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23158608

Chicago/Turabian Style

Williams, Rachel E., and Karen Mruk. 2022. "Aquatic Freshwater Vertebrate Models of Epilepsy Pathology: Past Discoveries and Future Directions for Therapeutic Discovery" International Journal of Molecular Sciences 23, no. 15: 8608. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23158608

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop