Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3072 2023-08-07 12:57:54 |
2 format correct Meta information modification 3072 2023-08-08 08:27:11 | |
3 format correct Meta information modification 3072 2023-08-10 05:39:53 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Malaník, M.; Čulenová, M.; Sychrová, A.; Skiba, A.; Skalicka-Woźniak, K.; Šmejkal, K. Treating Epilepsy with Natural Products. Encyclopedia. Available online: https://encyclopedia.pub/entry/47739 (accessed on 01 July 2024).
Malaník M, Čulenová M, Sychrová A, Skiba A, Skalicka-Woźniak K, Šmejkal K. Treating Epilepsy with Natural Products. Encyclopedia. Available at: https://encyclopedia.pub/entry/47739. Accessed July 01, 2024.
Malaník, Milan, Marie Čulenová, Alice Sychrová, Adrianna Skiba, Krystyna Skalicka-Woźniak, Karel Šmejkal. "Treating Epilepsy with Natural Products" Encyclopedia, https://encyclopedia.pub/entry/47739 (accessed July 01, 2024).
Malaník, M., Čulenová, M., Sychrová, A., Skiba, A., Skalicka-Woźniak, K., & Šmejkal, K. (2023, August 07). Treating Epilepsy with Natural Products. In Encyclopedia. https://encyclopedia.pub/entry/47739
Malaník, Milan, et al. "Treating Epilepsy with Natural Products." Encyclopedia. Web. 07 August, 2023.
Treating Epilepsy with Natural Products
Edit

Epilepsy is a neurological disease characterized by recurrent seizures that can lead to uncontrollable muscle twitching, changes in sensitivity to sensory perceptions, and disorders of consciousness. Although modern medicine has effective antiepileptic drugs, the need for accessible and cost-effective medication is urgent, and products derived from plants could offer a solution.

epilepsy anticonvulsant natural products

1. Natural Products That Affect Voltage-Gated Na+ and Ca2+ Channels

Voltage-gated sodium channels (VGSCs) initiate and conduct action potentials in excitable cells such as neurons or muscle cells, whereas voltage-gated calcium channels (VGCCs) are activated during action potentials, and they conduct the influx of Ca2+ into cells to initiate physiological processes, such as neurotransmission and muscle contraction [1]. Both channels represent molecular targets for drugs used in the treatment of epilepsy (see above).
Not many plant-derived compounds that affect VGSCs and VGCCs have been reported (Table 1). The effects of monoterpenes on various ion channels have been reviewed recently by Oz et al. [2], but most of the compounds mentioned exerted anticonvulsant effects only at doses too high to be implemented in clinical practice. Therefore, alkaloids and coumarins seem to have better prospects than monoterpenoids.
Several Aconitum alkaloids have been proven to interact with Na+ channels [3][4][5][6][7][8]. Unfortunately, these alkaloids have been isolated only from Aconitum species endemic to some regions of China. Furthermore, the isolation of these compounds is very tricky, with small yields. Because the total synthesis of aconitine-type alkaloids remains elusive, the prospects for Aconitum alkaloids are vague.
On the other hand, piperine, the major bioactive component in black pepper (Piper nigrum), significantly inhibited Na+ channel activity in mice [9]. Moreover, co-administration with carbamazepine (CBZ) or phenytoin decreased the elimination of these AEDs and enhanced their bioavailability [10]. The metabolism of piperine in the liver is limited, and its high blood–brain barrier permeability has been demonstrated in the Caco-2 monolayer model [11]. Its only limiting factor is poor solubility in water, and this can be improved, e.g., by a nanoprecipitation method leading to enhanced oral bioavailability and brain delivery of piperine after oral administration [12]. Altogether, piperine represents a very promising candidate for further evaluation in clinical trials, albeit with the caveat that increased attention must be paid to piperine-mediated drug interactions [13].
Similarly, the co-administration of coumarins significantly reduced the ED50 values of AEDs in the MES test in mice, as observed for imperatorin (40 mg/kg, i.p.) in combination with CBZ, phenobarbital, or phenytoin [14], and for xanthotoxin (50 and 100 mg/kg, i.p.) in combination with CBZ and valproate (VPA), respectively [15]. Additionally, xanthotoxin increased the total brain concentration of CBZ and VPA by about 84% and 46%, respectively, probably by inhibiting P-glycoprotein [15]. Therefore, co-administering natural compounds with conventional AEDs could be a way to increase the anticonvulsant activity of AEDs and thus improve the comfort of patients suffering from epilepsy.
Table 1. Natural products that affect VGSCs, VGCCs, or VGPCs.

2. Natural Products That Affect the Gabaergic Transmission

γ-Aminobutyric acid (GABA) is the main inhibitory neurotransmitter in the central nervous system (CNS). It is synthesized from glutamate by glutamic acid decarboxylase (GAD). GABA receptors can be divided into the ionotropic GABAA receptors and the metabotropic GABAB receptors. GABAA receptors are ligand-gated ion channels that increase the flow of chloride ions into the cell and thus promote inhibitory effects in the brain [23]. The potentiation of GABAergic transmission is one of the main targets of the AEDs currently used in clinical practice (see above).
Whereas only a few natural products interact with VGSCs and VGCCs, phytoconstituents most often affect the modulation of GABAA receptors (Table 2). The most significant anticonvulsant effects mediated by GABAergic transmission have been observed for terpenoids. Manayi et al. [24] have reviewed the ability of natural terpenoids to modulate the GABAergic system, but most of the compounds identified exerted anticonvulsant effects only at doses too high to be implemented in clinical practice.
Monoterpenes and sesquiterpenes are constituents of essential oils, and they have properties plausible for a potential application. Their low molecular weight and high lipophilicity allow them to penetrate membranes and interact with epilepsy-related proteins [25]. Thymoquinone, the dominant compound in black seed oil (Nigella sativa), has shown the most notable activity of all the monoterpenes found in this search, inhibiting convulsions in male BALB/c mice at the relatively low dose of 40 mg/kg [26]. Preclinical findings led to a pilot study in children with refractory epilepsy. The results demonstrated that thymoquinone was effective and well-tolerated [27]. Moreover, thymoquinone potentiated the antiepileptic properties of VPA [24] and phenytoin [28]. It also displayed a neuroprotective effect in Sprague Dawley rats by phosphorylating cAMP response element-binding protein (CREB) [29] and downregulating TNF-α and COX-2 [30]. Unfortunately, thymoquinone inhibits the activity of cytochrome P450 2C9 (CYP2C9), and this must be taken into account when thymoquinone is co-administered with phenytoin [31]. To sum up, its anticonvulsant and neuroprotective properties make thymoquinone a very promising substance that deserves further investigation with emphasis on an in-depth exploration of its pharmacokinetics and potential interactions.
Iridoids seem to be effective in the inhibition of convulsions as well, and especially the reports of the anticonvulsant activity of the extracts of Valeriana species are increasingly frequent. Pretreatment with valepotriate (5, 10, 20 mg/kg/day, i.p.) protected mice against MES and PTZ-induced convulsions, but it was far less effective than diazepam (4 mg/kg/day, i.p.) [32]. Significant anticonvulsant activity has also been reported for paederosidic acid, a rare sulfur-containing iridoid [33], but such compounds are, unfortunately, very unstable under acidic conditions, which makes their peroral administration difficult [34].
Bilobalide, the main sesquiterpene trilactone found in the leaves of Ginkgo biloba, must be included. Bilobalide (30 mg/kg, p.o., once a day for 4 days) elevated GABA levels in the hippocampus, cerebral cortex, and striatum of male ddY mice, possibly through the potentiation of the activity of GAD [35]. Also considering its neuroprotective effect [36], makes bilobalide seems very promising, although Ng et al. point out that bilobalide negatively modulated the action of GABA at α1β2γ2L GABAA receptors [37].
An even more promising natural product affecting GABA transmission has been found in huperzine A (HupA), a dietary supplement used in the USA as a memory enhancer. HupA is an acetylcholinesterase inhibitor isolated from the Chinese club moss Huperzia serrata. It also exerts anti-inflammatory and neuroprotective effects by activating nicotinic cholinergic receptors [38]. In addition, HupA has delivered seizure relief in a 6 Hz model, with an ED50 value of 0.34 mg/kg in the 32 mA paradigm, being 57 times more potent than levetiracetam and 301 times more potent than VPA [39]. It also suppressed PTZ-induced seizures in rats by activating the cortical transmission of GABA [40]. These findings, together with the favorable pharmacokinetic properties of HupA in humans [41], have led to clinical testing.
Recently, several coumarins have been reported to effectively inhibit PTZ-induced seizures in the zebrafish larvae model of epilepsy at doses lower than those of the positive controls diazepam (10 mM) and VPA (1 mM), respectively [42][43]. Based on the seizure-inducing agent used (PTZ), it was postulated that the test coumarins interfered with the GABA transmission. Coumarins have previously been shown to inhibit the activity of GABA transaminase (GABA-T), the main degradative enzyme of GABA [44]. This hypothesis was supported by a molecular docking study of oxypeucedanin hydrate, the most active furanocoumarin, to the structural model of GABA-T. The results indicated that a bulky substituent at the C5 position is crucial for antiseizure activity, whereas an analogous bulky moiety substituted at the C8 position diminishes the activity [43]. Similar results have been reported by Singhuber et al. [45] in a study dealing with the modulation of GABA-induced chloride currents by selected coumarin derivatives on recombinant α1β2γ2S GABAA receptors expressed in Xenopus laevis oocytes [45]. On the other hand, exactly the opposite results were found for the mice MES test. C5-substituted furanocoumarins were inactive, whereas C8-substituted furanocoumarins exerted strong anticonvulsant activity [46]. Hence, more studies are needed to clarify the structure–activity relationship with respect to the model of epilepsy used as well as to find out more about bioavailability. However, as coumarins are simple molecules, they are ideal for chemical synthesis and modifications to improve their pharmacodynamic and pharmacokinetic properties.
Table 2. Natural products with influence on GABAergic transmission.

3. Natural Products That Reduce Postsynaptic Excitability by Affecting AMPA or NMDA Receptors

Glutamate is the principal excitatory neurotransmitter in the brain, and therefore, glutamate receptor agonists, such as α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid (AMPA), N-methyl-D-aspartate (NMDA), and kainate, act as elicitors of seizures. Glutamate receptors can be divided into ionotropic glutamate receptors (ligand-gated ion channels) and metabotropic glutamate receptors (G-protein-coupled receptors) [65]. Among the ionotropic glutamate receptors, the AMPA-type and NMDA-type glutamate receptors are the most important, as certain AEDs affect these ionotropic receptors (e.g., perampanel and topiramate).
Interestingly, no reports of natural products interacting with AMPA receptors have been found during some searches, except for one recent study describing the potent anticonvulsant activity of magnolol and honokiol in a model of therapy-resistant epilepsy. However, the authors postulate the involvement of not only AMPA receptors but also GABAA and cannabinoid receptors [66]. Magnolol and honokiol, the main bioactive substances in the bark of Magnolia officinalis, are known for antioxidant, anti-inflammatory, and neuroprotective properties that make them promising for further research in the field of epilepsy, especially at a time when knowledge of their toxicity and bioavailability is accumulating [67].
Panax ginseng is well-known for its neuroprotective properties with ginsenosides as the main active constituents. Among the test ginsenosides, 20(S)-Rg3 and 20(S)-Rh2 inhibited NMDA receptors. However, 20(S)-Rg3 interacted with the glycine site, while 20(S)-Rh2 likely did so with the polyamine-binding site of NMDA receptors. (R)-isomers were inactive and the mono-glycosylated moiety at C-3 was found to be essential for binding to polyamine sites [68]. Further assays showed that 20(S)-Rg3 regulated GABAA receptor activity by interacting with the γ2 subunit [50], demonstrating its multitarget mechanism of action. Nevertheless, the bioavailability of ginsenosides after oral administration is relatively poor. Low water solubility and easy degradation by gastric acid and gut microbiota are the crucial disadvantages. Therefore, their absorption needs to be enhanced by sophisticated formulation strategies [69].

4. Natural Products with Multiple Mechanisms of Action: Cannabinoids

Cannabis sativa contains more than 100 compounds (lipophilic phytocannabinoids) with different therapeutic potentials and because phytocannabinoids affect diverse epilepsy-related targets [70], they have been given a separate section. Medicinal marijuana is applicable as a treatment option mainly for patients with chronic, autoimmune, inflammatory, degenerative, or oncological illnesses, and also for palliative care [71]. Multiple in vitro and in vivo preclinical trials have reported antiepileptic effects for several constituents of medical marijuana. Compounds of interest include psychoactive ∆9-tetrahydrocannabinol (∆9-THC) and structurally similar cannabidiol (CBD), along with non-psychoactive ∆9-tetrahydrocannabivarin (∆9-THCV), cannabidivarin (CBDV), and ∆9-tetrahydrocannabinolic acid (Δ9-THCA). ∆9-THC acts as a potent partial agonist on endocannabinoid receptor CB1, influencing both GABAergic and glutamatergic synaptic transmission. The clinical use of medical marijuana is limited because the anticonvulsant effect of ∆9-THC is relatively unpredictable. It acts simultaneously on several receptor targets, such as the transient receptor potential (TRP) cation channels TRPA1, TRPV2, and TRPM8; the orphan G-coupled protein receptor GPR55; the 5-HT3A receptor; the peroxisome proliferator-activated receptor gamma (PPARγ); the μ- and δ-opioid receptors, the β-adrenoreceptors; and some subtypes of Ca2+, K+, and Na+ channels. Interestingly, some experiments have shown medical marijuana to have no or even a pro-convulsant effect [70]. Δ9-THCA is used to prevent seizures, e.g., in the USA. This metabolic precursor of Δ9-THC is more affordable and should have only minor psychoactive properties [72]. Experimental observations have demonstrated that it has anticonvulsant effects via the modulation of ion channels and enzymes crucial for the biosynthesis of the endocannabinoid 2-arachidonoylglycerol [70]. The mechanisms of anticonvulsant activity of Δ9-THCV and CBDV are not well understood. Both compounds probably exert their anticonvulsant effects via non-CB1/CB2 mechanisms. TRPV1, TRPV2, TRPA1, and TRPM8 channels are the likely molecular targets of Δ9-THCV and CBDV [73].
CBD is the most promising of these agents, with effects proved by several clinical trials, especially for the treatment of drug-resistant epilepsies. Recent studies have proposed that its anticonvulsant effect may involve agonistic activity at the TRPV1 channel, the blockade of human T-type VGCCs, the modulation of various receptors such as 5-HT1A, 5-HT2A, GPR55, adenosine A1 and A2, voltage-dependent anion-selective channel protein 1 (VDAC1), or an influence on the release of TNF-α [70].
 Smoked, inhaled, or vaporized Δ9-THC has a bioavailability in the range from ~10–35% in contrast to (~6%) by oral administration. CBD and CBDV are hardly soluble in water, and their bioavailability after oral administration is poor, but many trials have evaluated cannabinoids suspended in sesame oil or mixed with glycocholate to increase their bioavailability, and intranasal, sublingual, and transdermal applications are common. Unfortunately, because of their lipophilic properties, large amounts of cannabinoids accumulate in adipose and other tissues, especially with repeated administration. Negative interactions with other drugs metabolized by the cytochrome P450 system or isoenzymes CYP3A4, CYP2C19, CYP2C9, and CYP2D6 should be taken into account, especially in Europe, where the co-administration of CBD and clobazam has been approved as adjunctive treatment of Dravet syndrome (DS) and Lennox–Gastaut syndrome (LGS). The strong inhibition of CYP2C19 by CBD leads to a remarkable rise in clobazam concentration, which may contribute to side effects, including somnolence and sedation. Finally, potenti

References

  1. Catterall, W.A.; Lenaeus, M.J.; Gamal El-Din, T.M. Structure and Pharmacology of Voltage-Gated Sodium and Calcium Channels. Annu. Rev. Pharmacol. Toxicol. 2020, 60, 133–154.
  2. Oz, M.; Lozon, Y.; Sultan, A.; Yang, K.-H.S.; Galadari, S. Effects of Monoterpenes on Ion Channels of Excitable Cells. Pharmacol. Ther. 2015, 152, 83–97.
  3. Ameri, A.; Gleitz, J.; Peters, T. Inhibition of Neuronal Activity in Rat Hippocampal Slices by Aconitum Alkaloids. Brain Res. 1996, 738, 154–157.
  4. Ameri, A.; Shi, Q.; Aschoff, J.; Peters, T. Electrophysiological Effects of Aconitine in Rat Hippocampal Slices. Neuropharmacology 1996, 35, 13–22.
  5. Ameri, A. Inhibition of Rat Hippocampal Excitability by the Plant Alkaloid 3-Acetylaconitine Mediated by Interaction with Voltage-Dependent Sodium Channels. Naunyn-Schmiedebergs Arch. Pharmacol. 1997, 355, 273–280.
  6. Ameri, A. Structure-Dependent Differences in the Effects of the Aconitum Alkaloids Lappaconitine, N-Desacetyllappaconitine and Lappaconidine in Rat Hippocampal Slices. Brain Res. 1997, 769, 36–43.
  7. Ameri, A. Inhibition of Rat Hippocampal Excitability by the Aconitum Alkaloid, 1-Benzoylnapelline, but Not by Napelline. Eur. J. Pharmacol. 1997, 335, 145–152.
  8. Ameri, A. Effects of the Alkaloids 6-Benzoylheteratisine and Heteratisine on Neuronal Activity in Rat Hippocampal Slices. Neuropharmacology 1997, 36, 1039–1046.
  9. Mishra, A.; Punia, J.K.; Bladen, C.; Zamponi, G.W.; Goel, R.K. Anticonvulsant Mechanisms of Piperine, a Piperidine Alkaloid. Channels 2015, 9, 317–323.
  10. Tiwari, A.; Mahadik, K.R.; Gabhe, S.Y. Piperine: A Comprehensive Review of Methods of Isolation, Purification, and Biological Properties. Med. Drug Discov. 2020, 7, 100027.
  11. Ren, T.; Wang, Q.; Li, C.; Yang, M.; Zuo, Z. Efficient Brain Uptake of Piperine and Its Pharmacokinetics Characterization after Oral Administration. Xenobiotica 2018, 48, 1249–1257.
  12. Ren, T.; Hu, M.; Cheng, Y.; Shek, T.L.; Xiao, M.; Ho, N.J.; Zhang, C.; Leung, S.S.Y.; Zuo, Z. Piperine-Loaded Nanoparticles with Enhanced Dissolution and Oral Bioavailability for Epilepsy Control. Eur. J. Pharm. Sci. 2019, 137, 104988.
  13. Lee, S.H.; Kim, H.Y.; Back, S.Y.; Han, H.-K. Piperine-Mediated Drug Interactions and Formulation Strategy for Piperine: Recent Advances and Future Perspectives. Expert Opin. Drug Metab. Toxicol. 2018, 14, 43–57.
  14. Luszczki, J.J.; Glowniak, K.; Czuczwar, S.J. Imperatorin Enhances the Protective Activity of Conventional Antiepileptic Drugs against Maximal Electroshock-Induced Seizures in Mice. Eur. J. Pharmacol. 2007, 574, 133–139.
  15. Zagaja, M.; Pyrka, D.; Skalicka-Wozniak, K.; Glowniak, K.; Florek-Luszczki, M.; Glensk, M.; Luszczki, J.J. Effect of Xanthotoxin (8-Methoxypsoralen) on the Anticonvulsant Activity of Classical Antiepileptic Drugs against Maximal Electroshock-Induced Seizures in Mice. Fitoterapia 2015, 105, 1–6.
  16. Hino, H.; Takahashi, H.; Suzuki, Y.; Tanaka, J.; Ishii, E.; Fukuda, M. Anticonvulsive Effect of Paeoniflorin on Experimental Febrile Seizures in Immature Rats: Possible Application for Febrile Seizures in Children. PLoS ONE 2012, 7, e42920.
  17. Sancheti, J.; Shaikh, M.F.; Chaudhari, R.; Somani, G.; Patil, S.; Jain, P.; Sathaye, S. Characterization of Anticonvulsant and Antiepileptogenic Potential of Thymol in Various Experimental Models. Naunyn-Schmiedebergs Arch. Pharmacol. 2014, 387, 59–66.
  18. Zhang, C.; Chen, J.; Zhao, F.; Chen, R.; Yu, D.; Cao, Z. Iritectol G, a Novel Iridal-Type Triterpenoid from Iris Tectorum Displays Anti-Epileptic Activity in Vitro through Inhibition of Sodium Channels. Fitoterapia 2017, 122, 20–25.
  19. Wu, K.-C.; Chen, Y.-H.; Cheng, K.-S.; Kuo, Y.-H.; Yang, C.-T.; Wong, K.-L.; Tu, Y.-K.; Chan, P.; Leung, Y.-M. Suppression of Voltage-Gated Na+ Channels and Neuronal Excitability by Imperatorin. Eur. J. Pharmacol. 2013, 721, 49–55.
  20. Skalicka-Woźniak, K.; Orhan, I.E.; Cordell, G.A.; Nabavi, S.M.; Budzyńska, B. Implication of Coumarins towards Central Nervous System Disorders. Pharmacol. Res. 2016, 103, 188–203.
  21. Lin, T.-Y.; Huang, W.-J.; Wu, C.-C.; Lu, C.-W.; Wang, S.-J. Acacetin Inhibits Glutamate Release and Prevents Kainic Acid-Induced Neurotoxicity in Rats. PLoS ONE 2014, 9, e88644.
  22. Ribeiro, R.A.; Leite, J.R. Nantenine Alkaloid Presents Anticonvulsant Effect on Two Classical Animal Models. Phytomedicine 2003, 10, 563–568.
  23. Akyuz, E.; Polat, A.K.; Eroglu, E.; Kullu, I.; Angelopoulou, E.; Paudel, Y.N. Revisiting the Role of Neurotransmitters in Epilepsy: An Updated Review. Life Sci. 2021, 265, 118826.
  24. Manayi, A.; Nabavi, S.M.; Daglia, M.; Jafari, S. Natural Terpenoids as a Promising Source for Modulation of GABAergic System and Treatment of Neurological Diseases. Pharmacol. Rep. 2016, 68, 671–679.
  25. De Alvarenga, J.F.R.; Genaro, B.; Costa, B.L.; Purgatto, E.; Manach, C.; Fiamoncini, J. Monoterpenes: Current Knowledge on Food Source, Metabolism, and Health Effects. Crit. Rev. Food Sci. Nutr. 2023, 63, 1352–1389.
  26. Hosseinzadeh, H.; Parvardeh, S. Anticonvulsant Effects of Thymoquinone, the Major Constituent of Nigella Sativa Seeds, in Mice. Phytomedicine 2004, 11, 56–64.
  27. Akhondian, J.; Kianifar, H.; Raoofziaee, M.; Moayedpour, A.; Toosi, M.B.; Khajedaluee, M. The Effect of Thymoquinone on Intractable Pediatric Seizures (Pilot Study). Epilepsy Res. 2011, 93, 39–43.
  28. Pottoo, F.H.; Salahuddin, M.; Khan, F.A.; Alomar, F.; Al Dhamen, M.A.; Alhashim, A.F.; Alqattan, H.H.; Gomaa, M.S.; Alomary, M.N. Thymoquinone Potentiates the Effect of Phenytoin against Electroshock-Induced Convulsions in Rats by Reducing the Hyperactivation of m-TOR Pathway and Neuroinflammation: Evidence from In Vivo, In Vitro and Computational Studies. Pharmaceuticals 2021, 14, 1132.
  29. Ullah, I.; Badshah, H.; Naseer, M.I.; Lee, H.Y.; Kim, M.O. Thymoquinone and Vitamin C Attenuates Pentylenetetrazole-Induced Seizures Via Activation of GABAB1 Receptor in Adult Rats Cortex and Hippocampus. Neuromol. Med. 2015, 17, 35–46.
  30. Shao, Y.; Feng, Y.; Xie, Y.; Luo, Q.; Chen, L.; Li, B.; Chen, Y. Protective Effects of Thymoquinone Against Convulsant Activity Induced by Lithium-Pilocarpine in a Model of Status Epilepticus. Neurochem. Res. 2016, 41, 3399–3406.
  31. Wang, Z.; Wang, X.; Wang, Z.; Lv, X.; Yin, H.; Li, W.; Li, W.; Jiang, L.; Liu, Y. Potential Herb-Drug Interaction Risk of Thymoquinone and Phenytoin. Chem. Biol. Interact. 2022, 353, 109801.
  32. Wu, A.; Ye, X.; Huang, Q.; Dai, W.-M.; Zhang, J.-M. Anti-Epileptic Effects of Valepotriate Isolated from Valeriana Jatamansi Jones and Its Possible Mechanisms. Phcog. Mag. 2017, 13, 512.
  33. Yang, T.; Kong, B.; Gu, J.-W.; Kuang, Y.-Q.; Cheng, L.; Yang, W.-T.; Cheng, J.-M.; Ma, Y.; Yang, X.-K. Anticonvulsant and Sedative Effects of Paederosidic Acid Isolated from Paederia Scandens (Lour.) Merrill. in Mice and Rats. Pharmacol. Biochem. Behav. 2013, 111, 97–101.
  34. Ling, S.-K.; Tanaka, T.; Kouno, I. Effects of Iridoids on Lipoxygenase and Hyaluronidase Activities and Their Activation by β-Glucosidase in the Presence of Amino Acids. Biol. Pharm. Bull. 2003, 26, 352–356.
  35. Sasaki, K.; Hatta, S.; Haga, M.; Ohshika, H. Effects of Bilobalide on γ-Aminobutyric Acid Levels and Glutamic Acid Decarboxylase in Mouse Brain. Eur. J. Pharmacol. 1999, 367, 165–173.
  36. Jiang, M.; Li, J.; Peng, Q.; Liu, Y.; Liu, W.; Luo, C.; Peng, J.; Li, J.; Yung, K.K.L.; Mo, Z. Neuroprotective Effects of Bilobalide on Cerebral Ischemia and Reperfusion Injury Are Associated with Inhibition of Pro-Inflammatory Mediator Production and down-Regulation of JNK1/2 and P38 MAPK Activation. J. Neuroinflammation 2014, 11, 167.
  37. Ng, C.C.; Duke, R.K.; Hinton, T.; Johnston, G.A.R. Effects of Bilobalide, Ginkgolide B and Picrotoxinin on GABAA Receptor Modulation by Structurally Diverse Positive Modulators. Eur. J. Pharmacol. 2017, 806, 83–90.
  38. Damar, U.; Gersner, R.; Johnstone, J.T.; Schachter, S.; Rotenberg, A. Huperzine A as a Neuroprotective and Antiepileptic Drug: A Review of Preclinical Research. Expert Rev. Neurother. 2016, 16, 671–680.
  39. Bialer, M.; Johannessen, S.I.; Levy, R.H.; Perucca, E.; Tomson, T.; White, H.S. Progress Report on New Antiepileptic Drugs: A Summary of the Twelfth Eilat Conference (EILAT XII). Epilepsy Res. 2015, 111, 85–141.
  40. Gersner, R.; Ekstein, D.; Dhamne, S.C.; Schachter, S.C.; Rotenberg, A. Huperzine A Prophylaxis against Pentylenetetrazole-Induced Seizures in Rats Is Associated with Increased Cortical Inhibition. Epilepsy Res. 2015, 117, 97–103.
  41. Li, Y.X.; Zhang, R.Q.; Li, C.R.; Jiang, X.H. Pharmacokinetics of Huperzine a Following Oral Administration to Human Volunteers. Eur. J. Drug Metabol. Pharmacokinet. 2007, 32, 183–187.
  42. Kozioł, E.; Deniz, F.S.S.; Orhan, I.E.; Marcourt, L.; Budzyńska, B.; Wolfender, J.-L.; Crawford, A.D.; Skalicka-Woźniak, K. High-Performance Counter-Current Chromatography Isolation and Initial Neuroactivity Characterization of Furanocoumarin Derivatives from Peucedanum Alsaticum L (Apiaceae). Phytomedicine 2019, 54, 259–264.
  43. Kozioł, E.; Jóźwiak, K.; Budzyńska, B.; De Witte, P.A.M.; Copmans, D.; Skalicka-Woźniak, K. Comparative Antiseizure Analysis of Diverse Natural Coumarin Derivatives in Zebrafish. Int. J. Mol. Sci. 2021, 22, 11420.
  44. Choi, S.Y.; Ahn, E.-M.; Song, M.-C.; Kim, D.W.; Kang, J.H.; Kwon, O.-S.; Kang, T.-C.; Baek, N.-I. In Vitro GABA-Transaminase Inhibitory Compounds from the Root OfAngelica Dahurica. Phytother. Res. 2005, 19, 839–845.
  45. Singhuber, J.; Baburin, I.; Ecker, G.F.; Kopp, B.; Hering, S. Insights into Structure–Activity Relationship of GABAA Receptor Modulating Coumarins and Furanocoumarins. Eur. J. Pharmacol. 2011, 668, 57–64.
  46. Luszczki, J.J.; Wojda, E.; Andres-Mach, M.; Cisowski, W.; Glensk, M.; Glowniak, K.; Czuczwar, S.J. Anticonvulsant and Acute Neurotoxic Effects of Imperatorin, Osthole and Valproate in the Maximal Electroshock Seizure and Chimney Tests in Mice: A Comparative Study. Epilepsy Res. 2009, 85, 293–299.
  47. Ding, J.; Wang, J.-J.; Huang, C.; Wang, L.; Deng, S.; Xu, T.-L.; Ge, W.-H.; Li, W.-G.; Li, F. Curcumol from Rhizoma Curcumae Suppresses Epileptic Seizure by Facilitation of GABA(A) Receptors. Neuropharmacology 2014, 81, 244–255.
  48. Garlet, Q.I.; Pires, L.D.C.; Milanesi, L.H.; Marafiga, J.R.; Baldisserotto, B.; Mello, C.F.; Heinzmann, B.M. (+)-Dehydrofukinone Modulates Membrane Potential and Delays Seizure Onset by GABAa Receptor-Mediated Mechanism in Mice. Toxicol. Appl. Pharmacol. 2017, 332, 52–63.
  49. Muceniece, R.; Saleniece, K.; Rumaks, J.; Krigere, L.; Dzirkale, Z.; Mezhapuke, R.; Zharkova, O.; Klusa, V. Betulin Binds to γ-Aminobutyric Acid Receptors and Exerts Anticonvulsant Action in Mice. Pharmacol. Biochem. Behav. 2008, 90, 712–716.
  50. Lee, B.-H.; Kim, H.-J.; Chung, L.; Nah, S.-Y. Ginsenoside Rg3 Regulates GABAA Receptor Channel Activity: Involvement of Interaction with the γ2 Subunit. Eur. J. Pharmacol. 2013, 705, 119–125.
  51. Kazmi, I.; Afzal, M.; Gupta, G.; Anwar, F. Antiepileptic Potential of Ursolic Acid Stearoyl Glucoside by GABA Receptor Stimulation. CNS Neurosci. Ther. 2012, 18, 799–800.
  52. Kundap, U.P.; Choo, B.K.M.; Kumari, Y.; Ahmed, N.; Othman, I.B.; Shaikh, M.F. Embelin Protects Against Acute Pentylenetetrazole-Induced Seizures and Positively Modulates Cognitive Function in Adult Zebrafish. Front. Pharmacol. 2019, 10, 1249.
  53. Zaugg, J.; Eickmeier, E.; Rueda, D.C.; Hering, S.; Hamburger, M. HPLC-Based Activity Profiling of Angelica Pubescens Roots for New Positive GABAA Receptor Modulators in Xenopus Oocytes. Fitoterapia 2011, 82, 434–440.
  54. Grigoletto, J.; Oliveira, C.V.D.; Grauncke, A.C.B.; Souza, T.L.D.; Souto, N.S.; Freitas, M.L.D.; Furian, A.F.; Santos, A.R.S.; Oliveira, M.S. Rosmarinic Acid Is Anticonvulsant against Seizures Induced by Pentylenetetrazol and Pilocarpine in Mice. Epilepsy Behav. 2016, 62, 27–34.
  55. Aseervatham, G.S.B.; Suryakala, U.; Doulethunisha; Sundaram, S.; Bose, P.C.; Sivasudha, T. Expression Pattern of NMDA Receptors Reveals Antiepileptic Potential of Apigenin 8-C-Glucoside and Chlorogenic Acid in Pilocarpine Induced Epileptic Mice. Biomed. Pharmacother. 2016, 82, 54–64.
  56. An, S.-J.; Park, S.-K.; Hwang, I.K.; Choi, S.Y.; Kim, S.K.; Kwon, O.-S.; Jung, S.J.; Baek, N.-I.; Lee, H.Y.; Won, M.H.; et al. Gastrodin Decreases Immunoreactivities of Gamma-Aminobutyric Acid Shunt Enzymes in the Hippocampus of Seizure-Sensitive Gerbils. J. Neurosci. Res. 2003, 71, 534–543.
  57. Nassiri-Asl, M.; Shariati-Rad, S.; Zamansoltani, F. Anticonvulsive Effects of Intracerebroventricular Administration of Rutin in Rats. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2008, 32, 989–993.
  58. Park, H.G.; Yoon, S.Y.; Choi, J.Y.; Lee, G.S.; Choi, J.H.; Shin, C.Y.; Son, K.H.; Lee, Y.S.; Kim, W.K.; Ryu, J.H.; et al. Anticonvulsant Effect of Wogonin Isolated from Scutellaria Baicalensis. Eur. J. Pharmacol. 2007, 574, 112–119.
  59. Abbasi, E.; Nassiri-Asl, M.; Shafeei, M.; Sheikhi, M. Neuroprotective Effects of Vitexin, a Flavonoid, on Pentylenetetrazole-Induced Seizure in Rats: Neuroprotective Effects of Vitexin. Chem. Biol. Drug Des. 2012, 80, 274–278.
  60. Yang, B.; Wang, J.; Zhang, N. Effect of Nobiletin on Experimental Model of Epilepsy. Transl. Neurosci. 2018, 9, 211–219.
  61. Faggion, S.A.; Cunha, A.O.S.; Fachim, H.A.; Gavin, A.S.; Dos Santos, W.F.; Pereira, A.M.S.; Beleboni, R.O. Anticonvulsant Profile of the Alkaloids (+)-Erythravine and (+)-11-α-Hydroxy-Erythravine Isolated from the Flowers of Erythrina Mulungu Mart Ex Benth (Leguminosae–Papilionaceae). Epilepsy Behav. 2011, 20, 441–446.
  62. Tamboli, A.M.; Rub, R.A.; Ghosh, P.; Bodhankar, S. Antiepileptic Activity of Lobeline Isolated from the Leaf of Lobelia Nicotianaefolia and Its Effect on Brain GABA Level in Mice. Asian Pac. J. Trop. Biomed. 2012, 2, 537–542.
  63. Da Silva, A.; De Andrade, J.; Bevilaqua, L.; Desouza, M.; Izquierdo, I.; Henriques, A.; Zuanazzi, J. Anxiolytic-, Antidepressant- and Anticonvulsant-like Effects of the Alkaloid Montanine Isolated from Hippeastrum Vittatum. Pharmacol. Biochem. Behav. 2006, 85, 148–154.
  64. Da Cruz, G.M.P.; Felipe, C.F.B.; Scorza, F.A.; Da Costa, M.A.C.; Tavares, A.F.; Menezes, M.L.F.; De Andrade, G.M.; Leal, L.K.A.M.; Brito, G.A.C.; Da Graça Naffah-Mazzacoratti, M.; et al. Piperine Decreases Pilocarpine-Induced Convulsions by GABAergic Mechanisms. Pharmacol. Biochem. Behav. 2013, 104, 144–153.
  65. Hanada, T. Ionotropic Glutamate Receptors in Epilepsy: A Review Focusing on AMPA and NMDA Receptors. Biomolecules 2020, 10, 464.
  66. Li, J.; Copmans, D.; Partoens, M.; Hunyadi, B.; Luyten, W.; De Witte, P. Zebrafish-Based Screening of Antiseizure Plants Used in Traditional Chinese Medicine: Magnolia Officinalis Extract and Its Constituents Magnolol and Honokiol Exhibit Potent Anticonvulsant Activity in a Therapy-Resistant Epilepsy Model. ACS Chem. Neurosci. 2020, 11, 730–742.
  67. Lin, Y.; Li, Y.; Zeng, Y.; Tian, B.; Qu, X.; Yuan, Q.; Song, Y. Pharmacology, Toxicity, Bioavailability, and Formulation of Magnolol: An Update. Front. Pharmacol. 2021, 12, 632767.
  68. Lee, E.; Kim, S.; Chul Chung, K.; Choo, M.-K.; Kim, D.-H.; Nam, G.; Rhim, H. 20(S)-Ginsenoside Rh2, a Newly Identified Active Ingredient of Ginseng, Inhibits NMDA Receptors in Cultured Rat Hippocampal Neurons. Eur. J. Pharmacol. 2006, 536, 69–77.
  69. Pan, W.; Xue, B.; Yang, C.; Miao, L.; Zhou, L.; Chen, Q.; Cai, Q.; Liu, Y.; Liu, D.; He, H.; et al. Biopharmaceutical Characters and Bioavailability Improving Strategies of Ginsenosides. Fitoterapia 2018, 129, 272–282.
  70. Gaston, T.E.; Friedman, D. Pharmacology of Cannabinoids in the Treatment of Epilepsy. Epilepsy Behav. 2017, 70, 313–318.
  71. Mouhamed, Y.; Vishnyakov, A.; Qorri, B.; Sambi, M.; Frank, S.S.; Nowierski, C.; Lamba, A.; Bhatti, U.; Szewczuk, M.R. Therapeutic potential of medicinal marijuana: An educational primer for health care professionals. Drug Healthc. Patient Saf. 2018, 10, 45–66.
  72. Perucca, E. Cannabinoids in the Treatment of Epilepsy: Hard Evidence at Last? J. Epilepsy Res. 2017, 7, 61–76.
  73. Devinsky, O.; Cilio, M.R.; Cross, H.; Fernandez-Ruiz, J.; French, J.; Hill, C.; Katz, R.; Di Marzo, V.; Jutras-Aswad, D.; Notcutt, W.G.; et al. Cannabidiol: Pharmacology and Potential Therapeutic Role in Epilepsy and Other Neuropsychiatric Disorders. Epilepsia 2014, 55, 791–802.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 228
Revisions: 3 times (View History)
Update Date: 10 Aug 2023
1000/1000
Video Production Service