Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-06-21T21:14:41.547Z Has data issue: false hasContentIssue false

Chapter 22 - The borderland between epilepsy and movement disorders

from Section IV - Movement disorders in general neurology

Published online by Cambridge University Press:  05 April 2014

Tasneem Peeraully
Affiliation:
Department of Neurology, Ronald Reagan UCLA Medical Center, Los Angeles, CA, USA
Ryuji Kaji
Affiliation:
Department of Neurology, Tokushima University, Tokushima, Japan
Eng-King Tan
Affiliation:
Neuroscience & Behavioral Disorders Program, Duke-NUS Graduate Medical School, Singapore
Werner Poewe
Affiliation:
Medical University Innsbruck
Joseph Jankovic
Affiliation:
Baylor College of Medicine, Texas
Get access

Summary

Introduction

The term “borderland of epilepsy” was coined by William Richard Gowers in 1907 when he described the borderland as disorders “near it (epilepsy), but not of it” (Crompton and Berkovic 2009).

Brief paroxysmal stereotyped events are the cardinal feature in both epilepsy and paroxysmal movement disorders (PMD). A seizure is an abnormal synchronous burst of neuronal activity, with epilepsy defined as a propensity to having seizures. Epilepsy is differentiated from movement disorders by the presence of characteristic ictal and interictal discharges on electroencephalography (EEG) and events may evolve into or be associated with generalized tonic-clonic or other seizures. Cortical involvement is a key factor in epileptic seizures, with the disruption between cortical and subcortical pathways influencing the semiology of the events. Altered functioning of subcortical structures is primarily implicated in generating PMD, although hypersynchronous discharges have not been disproved as the basis for PMD (Berkovic 2000).

Unlike generalized or complex partial seizures, movement disorders, in general, do not cause impaired consciousness. However, some seizures may cause very subtle alterations in mentation and in a number of sleep-related movement disorders patients have no recollection of the events. In addition, distinguishing between epilepsy and PMD in patients with developmental delay, impaired cognition, or altered mental status may present even greater difficulty.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2014

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abbott, G. W., Butler, M. H., Bendahhou, S., Dalakas, M. C., Ptacek, L. J., and Goldstein, S. A. (2001). “MiRP2 forms potassium channels in skeletal muscle with Kv3.4 and is associated with periodic paralysis,” Cell 104: 217–31.CrossRefGoogle ScholarPubMed
Alexopoulos, H. and Dalakas, M. C. (2010). “A critical update on the immunopathogenesis of Stiff Person Syndrome,” Eur. J. Clin. Invest. 40: 1018–25.CrossRefGoogle ScholarPubMed
Baizabal-Carvallo, J. F. and Jankovic, J. (2012). “Movement disorders in autoimmune diseases,” Mov. Disord. 27(8): 935–46.CrossRefGoogle ScholarPubMed
Berkovic, S. F. (2000). “Paroxysmal movement disorders and epilepsy: links across the channel,” Neurology 55(2): 169–70.CrossRefGoogle ScholarPubMed
Berkovic, S. F., Heron, S. E., Giordano, L., Marini, C., Guerrini, R., Kaplan, R. E., et al. (2004). “Benign familial neonatal-infantile seizures: characterization of a new sodium channelopathy,” Ann. Neurol. 55: 550–7.CrossRefGoogle ScholarPubMed
Berkovic, S. F., Mulley, J. C., Scheffer, I. E., and Petrou, S. (2006). “Human epilepsies: interaction of genetic and acquired factors,” Trends Neurosci. 29(7): 391–7.CrossRefGoogle ScholarPubMed
Bernard, G. and Shevell, M. I. (2008). “Channelopathies: a review,” Pediatr. Neurol. 38(2): 73–85.CrossRefGoogle ScholarPubMed
Biraben, A., Semah, F., Ribeiro, M. J., Douaud, G., Remy, P., and Depaulis, A. (2004). “PET evidence for a role of the basal ganglia in patients with ring chromosome 20 epilepsy,” Neurology 63(1): 73–7.CrossRefGoogle ScholarPubMed
Bouilleret, V., Semah, F., Chassoux, F., Mantzaridez, M., Biraben, A., Trebossen, R., et al. (2008). “Basal ganglia involvement in temporal lobe epilepsy: a functional and morphologic study,” Neurology 70(3): 177–84.CrossRefGoogle ScholarPubMed
Bozzi, Y., Dunleavy, M., and Henshall, D. C. (2011). “Cell signaling underlying epileptic behavior,” Front. Behav. Neurosci. 5: 45.CrossRefGoogle ScholarPubMed
Browner, N., Azher, S. N., and Jankovic, J. (2006). “Botulinum toxin treatment of facial myoclonus in suspected Rasmussen encephalitis,” Mov. Disord. 21: 1500–2.CrossRefGoogle ScholarPubMed
Caviness, J. N. (2009). “Pathophysiology and treatment of myoclonus,” Neurol. Clin. 27(3): 757–77.CrossRefGoogle ScholarPubMed
Chabardès, S., Kahane, P., Minotti, L., Koudsie, A., Hirsch, E., and Benabid, A. L. (2002). “Deep brain stimulation in epilepsy with particular reference to the subthalamic nucleus,” Epileptic Disord. 4(Suppl. 3): S83–93.Google ScholarPubMed
Chadwick, D., Hallett, M., Harris, R., Jenner, P., Reynolds, E. H., and Marsden, C. D. (1977). “Clinical, biochemical, and physiological features distinguishing myoclonus responsive to 5-hydroxytryptophan, tryptophan with a monoamine oxidase inhibitor, and clonazepam,” Brain 100(3): 455–87.CrossRefGoogle ScholarPubMed
Chauvel, P., Trottier, S., Vignal, J. P., and Bancaud, J. (1992). “Somatomotor seizures of frontal lobe origin,” Adv. Neurol. 57: 185–232.Google ScholarPubMed
Chkhenkeli, S. A. and Chkhenkeli, I. S. (1997). “Effects of therapeutic stimulation of nucleus caudatus on epileptic electrical activity of brain in patients with intractable epilepsy,” Stereotact. Funct. Neurosurg. 69(1–4 Pt. 2): 221–4.CrossRefGoogle ScholarPubMed
Chkhenkeli, S. A., Sramka, M., Lortkipanidze, G. S., Rakviashvili, T. N., Bregvadze, E. Sh., Magalashvili, G. E., et al. (2004). “Electrophysiological effects and clinical results of direct brain stimulation for intractable epilepsy,” Clin. Neurol. Neurosurg. 106(4): 318–29.CrossRefGoogle ScholarPubMed
Ciumas, C., Wahlin, T. B., Jucaite, A., Lindstrom, P., Halldin, C., and Savic, I. (2008). “Reduced dopamine transporter binding in patients with juvenile myoclonic epilepsy,” Neurology 71(11): 788–94.CrossRefGoogle ScholarPubMed
Cokar, O., Gelisse, P., Livet, M. O., Bureau, M., Habib, M., and Genton, P. (2001). “Startle response: epileptic or non-epileptic? The case for ‘flash’ SMA reflex seizures,” Epileptic Disord. 3(1): 7–12.Google ScholarPubMed
Comella, C. L., Tanner, C. M., and Ristanovic, R. K. (1993). “Polysomnographic sleep measures in Parkinson’s disease patients with treatment-induced hallucinations,” Ann. Neurol. 34(5): 710–14.CrossRefGoogle ScholarPubMed
Crompton, D. E. and Berkovic, S. F. (2009). “The borderland of epilepsy: clinical and molecular features of phenomena that mimic epileptic seizures,” Lancet Neurol. 8(4): 370–81.CrossRefGoogle ScholarPubMed
Crompton, D. E., Sadleir, L. G., Bromhead, C. J., Bahlo, M., Bellows, S. T., Arsov, T., et al. (2012). “Familial adult myoclonic epilepsy: recognition of mild phenotypes and refinement of the 2q locus,” Arch. Neurol. 69(4): 474–81.Google ScholarPubMed
Depienne, C., Magnin, E., Bouteiller, D., Stevanin, G., Saint-Martin, C., Vidailhet, M., Apartis, E., Hirsch, E., LeGuern, E., Labauge, P., and Rumbach, L. (2010). “Familial cortical myoclonic tremor with epilepsy: the third locus (FCMTE3) maps to 5p,” Neurology 74(24): 2000–3.CrossRefGoogle Scholar
Deppe, M., Kellinghaus, C., Duning, T., Möddel, G., Mohammadi, S., Deppe, K., et al. (2008). “Nerve fiber impairment of anterior thalamocortical circuitry in juvenile myoclonic epilepsy,” Neurology 71(24): 1981–5.CrossRefGoogle ScholarPubMed
Deprez, L., Weckhuysen, S., Holmgren, P., Suls, A., Van Dyck, T., Goossens, D., et al. (2010). “Clinical spectrum of early-onset epileptic encephalopathies associated with STXBP1 mutations,” Neurology 75(13): 1159–65.CrossRefGoogle ScholarPubMed
Dichgans, M., Freilinger, T., Eckstein, G., Babini, E., Lorenz-Depiereux, B., Biskup, S., et al. (2005). “Mutation in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine,” Lancet Neurol. 366(9483): 371–7.CrossRefGoogle ScholarPubMed
Edwards, M. J., Lang, A. E., and Bhatia, K. P. (2012). “Stereotypies: a critical appraisal and suggestion of a clinically useful definition,” Mov. Disord. 27(2): 179–85.CrossRefGoogle ScholarPubMed
Eisensehr, I., Lindeiner, H., Jäger, M., and Noachtar, S. (2001). “REM sleep behavior disorder in sleep-disordered patients with versus without Parkinson’s disease: is there a need for polysomnography?J. Neurol. Sci. 186(1–2): 7–11.CrossRefGoogle Scholar
Espeche, A. and Cersosimo, R. (2011). “Caraballo RH. Benign infantile seizures and paroxysmal dyskinesia: a well-defined familial syndrome,” Seizure 20(9): 686–91.CrossRefGoogle ScholarPubMed
Feddersen, B., Remi, J., Kilian, M., Vercueil, L., Deransart, C., Depaulis, A., et al. (2012). “Is ictal dystonia associated with an inhibitory effect on seizure propagation in focal epilepsies?Epilepsy Res. 99(3): 274–80.CrossRefGoogle ScholarPubMed
Fedi, M., Berkovic, S. F., Scheffer, I. E., O’Keefe, G., Marini, C., Mulligan, R., et al. (2008). “Reduced striatal D1 receptor binding in autosomal dominant nocturnal frontal lobe epilepsy,” Neurology 71(11): 795–8.CrossRefGoogle ScholarPubMed
Fisher, R., Salanova, V., Witt, T., Worth, R., Henry, T., Gross, R., et al. (SANTE Study Group) (2010). “Electrical stimulation of the anterior nucleus of thalamus for treatment of refractory epilepsy,” Epilepsia 51(5): 899–908.CrossRefGoogle ScholarPubMed
Franzini, A., Messina, G., Marras, C., Villani, F., Cordella, R., and Broggi, G. (2008). “Deep brain stimulation of two unconventional targets in refractory non-resectable epilepsy,” Stereotact. Funct. Neurosurg. 86(6): 373–81.CrossRefGoogle ScholarPubMed
Frucht, S. (2002). “Dystonia, athetosis, and epilepsia partialis continua in a patient with late-onset Rasmussen’s encephalitis,” Mov. Disord. 17(3): 609–12.CrossRefGoogle Scholar
Fulp, C. T., Cho, G., Marsh, E. D., Nasrallah, I. M., Labosky, P. A., and Golden, J. A. (2008). “Identification of Arx transcriptional targets in the developing basal forebrain,” Hum. Mol. Genet. 17(23): 3740–60.CrossRefGoogle ScholarPubMed
Gabasvili, V. M., Chkhenkeli, S. A., and Sramka, M. (1988). “The treatment of status epilepticus by electrostimulation of deep brain structures,” presented at: Modern Trends in Neurology and Neurological Emergencies, The First European Congress of Neurology (Prague).
Gambardella, A., Andermann, F., Shorvon, S., Le Piane, E., and Aguglia, U. (2008). “Limited chronic focal encephalitis: another variant of Rasmussen syndrome?Neurology 70(5): 374–7.CrossRefGoogle ScholarPubMed
Gangarossa, G., Di Benedetto, M., O’Sullivan, G. J., Dunleavy, M., Alcacer, C., Bonito-Oliva, A., et al. (2011). “Convulsant doses of a dopamine D1 receptor agonist result in Erk-dependent increases in Zif268 and Arc/Arg3.1 expression in mouse dentate gyrus,” PLoS One 6(5): e19415.CrossRefGoogle ScholarPubMed
Geis, C., Weishaupt, A., Hallermann, S., Grünewald, B., Wessig, C., Wultsch, T., et al. (2010). “Stiff person syndrome-associated autoantibodies to amphiphysin mediate reduced GABAergic inhibition,” Brain 133(11): 3166–80.CrossRefGoogle ScholarPubMed
Grant, R. and Graus, F. (2009). “Paraneoplastic movement disorders,” Mov. Disord. 24(12): 1715–24.CrossRefGoogle ScholarPubMed
Guerrini, R., Moro, F., Kato, M., Barkovich, A. J., Shiihara, T., McShane, M. A., et al. (2007). “Expansion of the first PolyA tract of ARX causes infantile spasms and status dystonicus,” Neurology 69(5): 427–33.CrossRefGoogle ScholarPubMed
Hallett, M., Chadwick, D., and Marsden, C. D. (1979). “Cortical reflex myoclonus,” Neurology 29(8): 1107–25.CrossRefGoogle ScholarPubMed
Hanna, P. A. and Walters, A. (2003). “Benign neonatal sleep myoclonus” in Chokroverty, S., Hening, W. A., and Walters, A., Sleep and Movement Disorders (Philadelphia, PA: Butterworth-Heinemann).Google Scholar
Haut, S. R. and Albin, R. L. (2008). “Dopamine and epilepsy: hints of complex subcortical roles,” Neurology 71(11): 784–5.CrossRefGoogle ScholarPubMed
Homanics, G. E., DeLorey, T. M., Firestone, L. L., Quinlan, J. J., Handforth, A., Harrison, N. L., et al. (1997). “Mice devoid of gamma-aminobutyrate type A receptor beta3 subunit have epilepsy, cleft palate, and hypersensitive behavior,” Proc. Natl. Acad. Sci. USA 94: 4143–8.CrossRefGoogle ScholarPubMed
Ikeda, A. and Kurihara, H. (2005). “Possible anticipation in BAFME: three generations examined in a Japanese family,” Mov. Disord. 8: 1076–89.CrossRefGoogle Scholar
Inoue, S. (1951). “On a pedigree of the hereditary tremor with epileptiform seizures,” J. Neuro. Psychiatry 53: 33–7.Google Scholar
Irani, S. R. and Vincent, A. (2012). “The expanding spectrum of clinically-distinctive, immunotherapy-responsive autoimmune encephalopathies,” Arq. Neuropsiquiatr. 70(4): 300–4.CrossRefGoogle ScholarPubMed
Kase, D., Inoue, T., and Imoto, K. (2012). “Roles of the subthalamic nucleus and subthalamic HCN channels in absence seizures,” J. Neurophysiol. 107(1): 393–406.CrossRefGoogle ScholarPubMed
Kortüm, F., Das, S., Flindt, M., Morris-Rosendahl, D. J., Stefanova, I., Goldstein, A., et al. (2011). “The core FOXG1 syndrome phenotype consists of postnatal microcephaly, severe mental retardation, absent language, dyskinesia, and corpus callosum hypogenesis,” J. Med. Genet. 48(6): 396–406.CrossRefGoogle ScholarPubMed
Hahn, C. D. (2007). “The ARX story: a new twist,” Neurology 69(5): 421–2.CrossRefGoogle ScholarPubMed
Harvey, R. J., Topf, M., Harvey, K., and Rees, M. I. (2008). “The genetics of hyperekplexia: more than startle!Trends Genet. 24(9): 439–47.CrossRefGoogle ScholarPubMed
Husain, A. M. and Sinha, S. R. (2011). “Nocturnal epilepsy in adults,” J. Clin. Neurophysiol. 28(2): 141–5.CrossRefGoogle ScholarPubMed
Korn-Lubetzki, I., Bien, C. G., Bauer, J., Gomori, M., Wiendl, H., Trajo, L., et al. (2004). “Rasmussen encephalitis with active inflammation and delayed seizures onset,” Neurology 62(6): 984–6.CrossRefGoogle ScholarPubMed
Krauss, G. L. and Mathews, G. C. (2003). “Similarities in mechanisms and treatments for epileptic and non-epileptic myoclonus,” Epilepsy Curr. 3(1): 19–21.CrossRefGoogle Scholar
Kurian, M., Fluss, J., and Korff, C. (2012). “Anti-NMDA receptor encephalitis: the importance of early diagnosis and aggressive immunotherapy in tumor negative pediatric patients,” Eur. J. Paediatr. Neurol. 16(6): 764–5.CrossRefGoogle ScholarPubMed
Kurian, M. A., Gissen, P., Smith, M., Heales, S., and Clayton, P. T. (2011). “The monoamine neurotransmitter disorders: an expanding range of neurological syndromes,” Lancet Neurol. 10(8): 721–33.CrossRefGoogle ScholarPubMed
Lee, H. Y., Huang, Y., Bruneau, N., Roll, P., Roberson, E. D., Hermann, M., et al. (2012). “Mutations in the novel protein PRRT2 cause paroxysmal kinesigenic dyskinesia with infantile convulsions,” Cell Rep. 1(1): 2–12.CrossRefGoogle Scholar
Lee, H. Y., Xu, Y., Huang, Y., Ahn, A. H., Auburger, G. W., Pandolfo, M., et al. (2004). “The gene for paroxysmal non-kinesigenic dyskinesia encodes an enzyme in a stress response pathway,” Hum. Mol. Genet. 13(24): 3161–70.CrossRefGoogle Scholar
Le Meur, N., Holder-Espinasse, M., Jaillard, S., Goldenberg, A., Joriot, S., Amati-Bonneau, P., et al. (2010). “MEF2C haploinsufficiency caused by either microdeletion of the 5q14.3 region or mutation is responsible for severe mental retardation with stereotypic movements, epilepsy and/or cerebral malformations,” J. Med. Genet. 47(1): 22–9.CrossRefGoogle ScholarPubMed
Mastrangelo, M. and Leuzzi, V. (2012). “Genes of early-onset epileptic encephalopathies: from genotype to phenotype,” Pediatr. Neurol. 46(1): 24–31.CrossRefGoogle ScholarPubMed
Matsumoto, R. R., Truong, D. D., Nguyen, K. D., Dang, A. T., Hoang, T. T., Vo, P. Q., et al. (2000). “Involvement of GABA(A) receptors in myoclonus,” Mov. Disord. 15(Suppl. 1): 47–52.CrossRefGoogle ScholarPubMed
Mima, T., Nagamine, T., Ikeda, A., Yazawa, S., Kimura, J., and Shibasaki, H. (1998). “Pathogenesis of cortical myoclonus studied by magnetoencephalography,” Ann. Neurol. 43: 598–607.CrossRefGoogle ScholarPubMed
Mima, T., Nagamine, T., Nishitani, N., Mikuni, N., Ikeda, A., Fukuyama, H., et al. (1998). “Cortical myoclonus. Sensorimotor hyperexcitability,” Neurology 50: 933–42.CrossRefGoogle ScholarPubMed
Mullen, S. A., Suls, A., De Jonghe, P., Berkovic, S. F., and Scheffer, I. E. (2010). “Absence epilepsies with widely variable onset are a key feature of familial GLUT1 deficiency,” Neurology 75(5): 432–40.CrossRefGoogle ScholarPubMed
Nickels, K. and Wirrell, E. (2010). “GLUT1-ous maximus epilepticus: the expanding phenotype of GLUT-1 mutations and epilepsy,” Neurology 75(5): 390–1.CrossRefGoogle ScholarPubMed
Nishi, A., Snyder, G. L., and Greengard, P. (1997). “Bidirectional regulation of DARPP-32 phosphorylation by dopamine,” J. Neurosci. 17(21): 8147–55.CrossRefGoogle ScholarPubMed
Oakley, J. C. and Ojemann, G. A. (1982). “Effects of chronic stimulation of the caudate nucleus on a pre-existing alumina seizure focus,” Exp. Neurol. 75(2): 360–7.CrossRefGoogle Scholar
O’Sullivan, G. J., Dunleavy, M., Hakansson, K., Clementi, M., Kinsella, A., Croke, D. T., et al. (2008). “Dopamine D1 vs D5 receptor-dependent induction of seizures in relation to DARPP-32, ERK1/2 and GluR1-AMPA signalling,” Neuropharmacology 54(7): 1051–61.CrossRefGoogle ScholarPubMed
Paciorkowski, A. R., Thio, L. L., and Dobyns, W. B. (2011). “Genetic and biologic classification of infantile spasms,” Pediatr. Neurol. 45(6): 355–67.CrossRefGoogle ScholarPubMed
Pons, R., Cuenca-León, E., Miravet, E., Pons, M., Xaidara, A., Youroukos, S., et al. (2012). “Paroxysmal non-kinesigenic dyskinesia due to a PNKD recurrent mutation: report of two Southern European families,” Eur. J. Paediatr. Neurol. 16(1): 86–9.CrossRefGoogle ScholarPubMed
Ridley, A., Kennard, C., Scholtz, C. L., Büttner-Ennever, J. A., Summers, B., and Turnbull, A. (1987). “Omnipause neurons in two cases of opsoclonus associated with oat cell carcinoma of the lung,” Brain 110 (Pt. 6): 1699–709.CrossRefGoogle ScholarPubMed
Rochette, J., Roll, P., Fu, Y. H., Lemoing, A. G., Royer, B., Roubertie, A., et al. (2010). “Novel familial cases of ICCA (infantile convulsions with paroxysmal choreoathetosis) syndrome,” Epileptic Disord. 12(3): 199–204.Google ScholarPubMed
Sasaki, M., Sakuma, H., Fukushima, A., Yamada, K., Ohnishi, T., and Matsuda, H. (2009). “Abnormal cerebral glucose metabolism in alternating hemiplegia of childhood,” Brain Dev. 31(1): 20–6.CrossRefGoogle ScholarPubMed
Seebohm, G. (2005). “Activators of cation channels: potential in treatment of channelopathies,” Mol. Pharmacol. 67(3): 585–8.CrossRefGoogle ScholarPubMed
Shen, Y., Lee, H. Y., Rawson, J., Ojha, S., Babbitt, P., Fu, Y. H., et al. (2011). “Mutations in PNKD causing paroxysmal dyskinesia alters protein cleavage and stability,” Hum. Mol. Genet. 20(12): 2322–32.CrossRefGoogle ScholarPubMed
Shibasaki, H. and Hallett, M. (2005). “Electrophysiological studies of myoclonus,” Muscle Nerve 31(2): 157–74.CrossRefGoogle ScholarPubMed
Sramka, M. and Chkhenkeli, S. A. (1990). “Clinical experience in intraoperational determination of brain inhibitory structures and application of implanted neurostimulators in epilepsy,” Stereotact. Funct. Neurosurg. 54–5: 56–9.CrossRefGoogle ScholarPubMed
Sramka, M., Fritz, G., Galanda, M., and Nadvornik, P. (1976). “Some observations in treatment stimulation of epilepsy,” Acta Neurochir. (Wien) (Suppl. 23): 257–62.
Szepetowski, P., Rochette, J., Berquin, P., Piussan, C., Lathrop, G. M., and Monaco, A. P. (1997). “Familial infantile convulsions and paroxysmal choreoathetosis: a new neurological syndrome linked to the pericentromeric region of human chromosome 16,” Am. J. Hum. Genet. 61(4): 889–98.CrossRefGoogle ScholarPubMed
Tassinari, C. A., Rubboli, G., Gardella, E., Cantalupo, G., Calandra-Buonaura, G., Vedovello, M., et al. (2005). “Central pattern generators for a common semiology in fronto-limbic seizures and in parasomnias. A neuroethologic approach,” Neurol. Sci. 26(Suppl. 3): s225–32.CrossRefGoogle ScholarPubMed
Tenney, J. R. and Schapiro, M. B. (2010). “Child neurology: alternating hemiplegia of childhood,” Neurology 74(14): e57–9.CrossRefGoogle ScholarPubMed
Terada, K., Ikeda, A., Mima, T., Kimura, M., Nagahama, Y., Kamioka, Y., et al. (1997). “Familial cortical myoclonic tremor as a unique form of cortical reflex myoclonus,” Mov. Disord. 12: 370–7.CrossRefGoogle ScholarPubMed
Tripathi, P. P., Santorufo, G., Brilli, E., Borrelli, E., and Bozzi, Y. (2010). “Kainic acid-induced seizures activate GSK-3β in the hippocampus of D2R-/- mice,” Neuroreport 21(12): 846–50.CrossRefGoogle ScholarPubMed
Turner, G., Partington, M., Kerr, B., Mangelsdorf, M., and Gecz, J. (2002). “Variable expression of mental retardation, autism, seizures, and dystonic hand movements in two families with an identical ARX gene mutation,” Am. J. Med. Genet. 112(4): 405–11.CrossRefGoogle ScholarPubMed
Uesaka, Y., Terao, Y., Ugawa, Y., Yumoto, M., Hanajima, R., and Kanazawa, I. (1996). “Magnetoencephalographic analysis of cortical myoclonic jerks,” Electroencephalogr. Clin. Neurophysiol. 99: 141–8.CrossRefGoogle ScholarPubMed
Uncini, A., Basciani, M., Di Muzio, A., Antonini, D., and Onofrj, M. (1990). “Methyl bromide myoclonus: an electrophysiological study,” Acta Neurol. Scand. 81(2): 159–64.CrossRefGoogle Scholar
Velísková, J. and Moshé, S. L. (2006). “Update on the role of substantia nigra pars reticulata in the regulation of seizures,” Epilepsy Curr. 6(3): 83–7.CrossRefGoogle ScholarPubMed
Vincent, A., Bien, C. G., Irani, S. R., and Waters, P. (2011). “Autoantibodies associated with diseases of the CNS: new developments and future challenges,” Lancet Neurol. 10(8): 759–72.CrossRefGoogle ScholarPubMed
Wang, J. L., Cao, L., Li, X. H., Hu, Z. M., Li, J. D., Zhang, J. G., et al. (2011). “Identification of PRRT2 as the causative gene of paroxysmal kinesigenic dyskinesias,” Brain 134(Pt. 12): 3493–501.CrossRefGoogle ScholarPubMed
Wetter, T. C., Brunner, H., Högl, B., Yassouridis, A., Trenkwalder, C., and Friess, E. (2001). “Increased alpha activity in REM sleep in de novo patients with Parkinson’s disease,” Mov. Disord. 16(5): 928–33.CrossRefGoogle ScholarPubMed
Zafeiriou, D., Vargiami, E., and Kontopoulos, E. (2003). “Reflex myoclonic epilepsy in infancy: a benign age-dependent idiopathic startle epilepsy,” Epileptic Disord. 5(2): 121–2.Google ScholarPubMed
Zeng, L. H., Rensing, N. R., and Wong, M. (2009). “Developing antiepileptogenic drugs for acquired epilepsy: targeting the mammalian target of rapamycin (mTOR) pathway,” Mol. Cell. Pharmacol. 1(3): 124–9.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×