From Wikipedia, the free encyclopedia
Hyperekplexia
Other namesExaggerated surprise, exaggerated startle response, startle disease [1]
Mutations of the neuroreceptor glycine receptor subunit alpha-1 (GLRA1) can cause hyperekplexia.
Pronunciation
  • /ˌhaɪ.pɚ.ɛkˈplɛk.si.ə/
Specialty Neurology, genetics
SymptomsIncreased startle response to sudden auditory, visual, or tactile stimulation
ComplicationsIncreased alcohol and drug use
DurationChronic
CausesMutation of either the GLRA1 gene, GLRB gene, SLC6A5 gene, X-linked ( ARHGEF9) gene, or GPHN gene [2]
Differential diagnosis Jumping Frenchmen of Maine syndrome
Medication Clonazepam or phenobarbital, carbamazepine, 5-hydroxytryptophan, phenytoin, sodium valproate, diazepam, or piracetam [2]
Frequency1 in 40,000 [2]

Hyperekplexia ( /ˌh.pər.ɛkˈplɛk.si.ə/; "exaggerated surprise") is a very rare neurologic disorder, classically characterised by a pronounced startle responses to tactile or acoustic stimuli and an ensuing period of hypertonia. The hypertonia may be predominantly truncal, attenuated during sleep, or less prominent after one year of age.

Classic hyperekplexia is caused by genetic mutations in a number of different genes, all of which play an important role in glycine neurotransmission. Glycine is used by the central nervous system as an inhibitory neurotransmitter. Hyperekplexia is generally classified as a genetic disease; [3] some disorders can mimic the exaggerated startle of hyperekplexia. [4]

Signs and symptoms

The three main signs of hyperekplexia are generalized stiffness, excessive startle beginning at birth and nocturnal myoclonus. [5] Affected individuals are fully conscious during episodes of stiffness, which consist of forced closure of the eyes and an extension of the extremities followed by a period of generalised stiffness and uncontrolled falling at times. [6] Initially, the disease was classified into a "major" and a "minor" form, with the minor form being characterized by an excessive startle reflex, but lacking stiffness. [6] Genetic evidence has only been found for the major form of the condition. [6]

Other signs and symptoms of hyperekplexia may include episodic neonatal apnea, excessive movement during sleep and the head-retraction reflex. The link to some cases of Sudden Infant Death remains controversial. [3]

Genetics

Hyperekplexia is known to be caused by a variety of genes, encoding both pre- and postsynaptic proteins. The symptoms displayed, as well as the types of inheritance, vary, based on the affected gene.[ citation needed]

GLRA1

The first gene linked conclusively to hyperekplexia was GLRA1. [6] The GLRA1 gene encodes the glycine receptor, alpha 1 subunit, which, together with the glycine receptor beta subunit, forms synaptic glycine receptors. Inhibitory glycine receptors are ligand-gated chloride channels that facilitate fast responses in the brainstem and spinal-cord. Homomeric glycine receptors composed exclusively of alpha-1 subunits exhibit normal ion channel electrophysiology, but are not sequestered at the synaptic junction. [7] Wild-type glycine receptors are thus presumed to be pentameric heteromers of the alpha-1 and beta subunits, in either a 3:2 or 2:3 ratio. [7]

Within these heteromers, it is believed that the alpha-1 subunits bind glycine and undergo a conformational change, inducing a conformational change in the pentamer, causing the ion-channel to open. Although autosomal dominant [6] inheritance was initially reported, there are at least as many cases described with autosomal recessive inheritance. [8] Thus far, the general rule is that mutations causing structurally normal proteins that cannot bind glycine or cannot properly undergo a required conformational change in response to glycine will result in a dominant form of the disease, while mutations that result in truncated or greatly malformed subunits that cannot be integrated into a receptor protein will result in a recessive form. [8]

GLRB

The GLRB gene encodes the beta subunit of the glycine receptor. Homomeric glycine receptors composed of beta subunits do not open in response to glycine stimulation, [9] however, the beta subunit is essential for proper receptor localization, through its interactions with gephyrin, which results in receptor clustering at the synaptic cleft. [10] As such, the defects within the GLRB gene show autosomal recessive inheritance. [11]

SLC6A5

The SLC6A5 gene encodes the GlyT2 transporter, a neuronal pre-synaptic glycine re-uptake transporter. In comparison to the GlyT1 transporter, found mostly in glial cells, GlyT2 helps maintain a high concentration of glycine within the axon terminal of glycinergic neurons. [12] Mutations of the SLC6A5 gene have been associated with hyperekplexia in an autosomal recessive inheritance pattern. [13] Defects within this gene are hypothesized either to affect the incorporation of the transporter into the cellular membrane or to affect its affinity for the molecules it transports: sodium ions, chloride ions and glycine. [13] Any of these actions would drastically reduce the pre-synaptic cell's ability to produce the high vesicular concentrations of glycine necessary for proper glycine neurotransmission. GPHN and ARHGEF9 are often included in lists of genetic causes of hyperekplexia - but, in fact, they produce a much more complex phenotype, very distinct from classical hyperekplexia. As such they are no longer considered to be causative genes.[ citation needed]

GPHN

Gephyrin, an integral membrane protein believed to coordinate glycine receptors, is coded by the gene GPHN. A heterozygous mutation in this gene has been identified in one sporadic case of hyperekplexia, though experimental data is inconclusive as to whether the mutation itself is, in fact, pathogenic. [14] Gephyrin is essential for glycine receptor clustering at synaptic junctions, through its action of binding both the glycine receptor beta subunit and internal cellular microtubule structures. [10] Gephyrin also assists in clustering GABA receptors at synapses and molybdenum cofactor synthesis. [15] Because of gephyrin's multi-functional nature, in mutated form it is not presumed to be a common genetic source of hyperekplexia. [14]

ARHGEF9

A defect within the gene coding for collybistin, ARHGEF9, has been shown to cause hyperekplexia occurring with epilepsy. [16] Since the ARHGEF9 gene is on the X chromosome, this gene displays X-linked recessive heritance. The collybistin protein is responsible for proper gephyrin targeting, which is crucial for the proper localization of glycine and GABA receptors. Deficiencies in collybistin function would result in a lack of glycine and GABA receptors at the synaptic cleft. [16]

Diagnosis

There are three signs used to diagnose if an infant has hereditary hyperekplexia: if the child's body is stiff all over as soon as they are born, if they overreact to noises and other stimuli, and if the reaction to stimuli is followed by an overall stiffness where the child is unable to make any voluntary movements. [17] A combination of electroencephalogram and an electromyogram may help diagnose this condition in patients who have not displayed symptoms as children. The electroencephalogram will not show abnormal activity other than a spike in wakefulness or alertness, while the electromyogram will show rapid muscular responses and hyperreflexia. Otherwise, genetic testing is the only definitive diagnosis. [17] MRIs and CT scans will be normal unless other conditions are present. [17]

Treatment

The most commonly-used effective treatment is clonazepam, which leads to the increased efficacy of another inhibitory neurotransmitter, GABA. [3] There are anecdotal reports of the use of levetiracetam in genetic and acquired[ clarification needed] hyperekplexia. [18] During attacks of hypertonia and apnea, the limbs and head may be forcibly manipulated towards the trunk in order to resolve the symptoms. This is referred to as the "Vigevano maneuver'. [19]

History

The disorder was first described in 1958 by Kirstein and Silfverskiold, who reported a family with 'drop seizures'. [20] In 1962 Drs. Kok and Bruyn reported an unidentified hereditary syndrome, which initially presented as hypertonia in infants. [21] Genetic analysis within this large Dutch pedigree revealed a mutation within the GLRA1 gene, the first gene to be implicated in hyperekplexia. [6]

See also

References

  1. ^ Beers MH (2006). The Merck Manual (16th ed.). Whitehouse Station, NJ: Merck Research Laboratories. p. 1764. ISBN  0911910-18-2.
  2. ^ a b c Kerkar Pramod, M.D., FFARCSI, DA (22 December 2015). "Exaggerated Startle Response: Causes, Symptoms, Treatment, Recovery, Yoga". PainAssist. Retrieved 19 May 2020.{{ cite web}}: CS1 maint: multiple names: authors list ( link)
  3. ^ a b c Bakker MJ, van Dijk JG, van den Maagdenberg AM, Tijssen MA (2006-05-19). "Startle Syndromes". Lancet Neurology. 5 (6): 513–524. doi: 10.1016/S1474-4422(06)70470-7. PMID  16713923. S2CID  24056686.
  4. ^ van de Warrenburg BP, C. Cordivari, P. Brown, K. P. Bhatia (2007-04-05). "Persisting Hyperekplexia After Idiopathic, Self-Limiting Brainstem Encephalopathy". Movement Disorders. 22 (7): 1017–20. doi: 10.1002/mds.21411. PMID  17415799. S2CID  30137238.
  5. ^ Koning-Tijssen M, O.F. Brouwer (2000-04-27). "Hyperekplexia in the Neonate". Movement Disorders. 15 (6): 1293–6. doi: 10.1002/1531-8257(200011)15:6<1293::aid-mds1047>3.0.co;2-k. PMID  11104232. S2CID  29366280.
  6. ^ a b c d e f Tijssen M, R. Shiang, J. van Deutekom, R. H. Boerman, J. Wasmuth, L. A. Sandkuijl, R. R. Frants, G. W. Padberg (1995-06-01). "Molecular Genetic Reevaluation of the Dutch Hyperekplexia Family" (PDF). Archives of Neurology. 52 (6): 578–582. doi: 10.1001/archneur.1995.00540300052012. hdl: 2066/20657. PMID  7763205. S2CID  14067463.
  7. ^ a b Lynch JW (2008-08-03). "Native glycine receptor subtypes and their physiological roles". Neuropharmacology. 56 (1): 303–9. doi: 10.1016/j.neuropharm.2008.07.034. PMID  18721822. S2CID  43613876.
  8. ^ a b Villmann C, Oertel J, Melzer N, Becker CM (2009). "Recessive hyperekplexia mutations of the glycine receptor [alpha]-1 subunit affect cell surface integration and stability". Journal of Neurochemistry. 111 (3): 837–847. doi: 10.1111/j.1471-4159.2009.06372.x. PMID  19732286. S2CID  35256060.
  9. ^ Bormann J, N. Rundstrom, H. Betz, D. Langosch (1993). "Residues within transmembrane segment M2 determine chloride conductance of glycine receptor homo- and hetero-oligomers". EMBO Journal. 12 (10): 3729–37. doi: 10.1002/j.1460-2075.1993.tb06050.x. PMC  413654. PMID  8404844.
  10. ^ a b Meyer G, J. Kirsch, H. Betz, D. Langosch (1995). "Identification of a Gephyrin Binding Motif on the Glycine Receptor Beta Subunit". Neuron. 15 (3): 563–572. doi: 10.1016/0896-6273(95)90145-0. PMID  7546736. S2CID  10164739.
  11. ^ Rees MI, T. M. Lewis, J. B. Kwok, G. R. Mortier, P. Govaert, R. G. Snell, P. R. Schofield, M. J. Owen (2002-04-01). "Hyperekplexia associated with compound heterozygote mutations in the beta-subunit of the human inhibitory glycine receptor (GLRB)". Human Molecular Genetics. 11 (7): 853–860. doi: 10.1093/hmg/11.7.853. PMID  11929858.
  12. ^ Rousseau F, K. R. Aubrey, S. Supplisson (2008-09-24). "The Glycine Transporter GlyT2 Controls the Dynamics of Synaptic Vesicle Refilling in Inhibitory Spinal Cord Neurons". Journal of Neuroscience. 28 (39): 9755–68. doi: 10.1523/JNEUROSCI.0509-08.2008. PMC  6671229. PMID  18815261.
  13. ^ a b Rees MI, Harvey K, Pearce BR, Chung SK, Duguid IC, Thomas P, Beatty S, Graham GE, Armstrong L, Shiang R, Abbott KJ, Zuberi SM, Stephenson JB, Owen MJ, Tijssen MA, van den Maagdenberg AM, Smart TG, Supplisson S, Harvey RJ (2006). "Mutations in the gene encoding GlyT2 (SLC6A5) define a presynaptic component of human startle disease". Nature Genetics. 38 (7): 801–806. doi: 10.1038/ng1814. PMC  3204411. PMID  16751771.
  14. ^ a b Rees MI, Harvey K, Ward H, White JH, Evans L, Duguid IC, Hsu CC, Coleman SL, Miller J, Baer K, Waldvogel HJ, Gibbon F, Smart TG, Owen MJ, Harvey RJ, Snell RG (2003-04-08). "Isoform Heterogeneity of the Human Gephyrin Gene (GPHN), Binding Domains to the Glycine Receptor, and Mutation Analysis in Hyperekplexia". Journal of Biological Chemistry. 278 (27): 24688–96. doi: 10.1074/jbc.M301070200. PMID  12684523.
  15. ^ Fritschy JM, R. J. Harvey, G. Schwarz (2008). "Gephyrin: where do we stand, where do we go?". Trends in Neurosciences. 31 (5): 257–264. doi: 10.1016/j.tins.2008.02.006. PMID  18403029. S2CID  6885626.
  16. ^ a b Harvey K, Duguid IC, Alldred MJ, Beatty SE, Ward H, Keep NH, Lingenfelter SE, Pearce BR, Lundgren J, Owen MJ, Smart TG, Lüscher B, Rees MI, Harvey RJ (2004). "The GDP-GTP Exchange Factor Collybistin: An Essential Determinant of Neuronal Gephyrin Clustering" (PDF). Journal of Neuroscience. 24 (25): 5816–26. doi: 10.1523/JNEUROSCI.1184-04.2004. PMC  6729214. PMID  15215304.
  17. ^ a b c Tijssen MA, Rees MI (1993). "Hyperekplexia". In Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJ, Stephens K, Amemiya A (eds.). GeneReviews®. Seattle (WA): University of Washington, Seattle. PMID  20301437.
  18. ^ Luef GJ, W. N. Loescher (June 2007). "The effect of levetiracetam in startle disease". Journal of Neurology. 254 (6): 808–9. doi: 10.1007/s00415-006-0437-z. PMID  17401745. S2CID  358799.
  19. ^ Vigevano F, M. Di Capua, B. Dalla Bernardina (1989). "Startle disease: an avoidable cause of sudden infant death". Lancet. 1 (8631): 216. doi: 10.1016/s0140-6736(89)91226-9. PMID  2563117. S2CID  32077413.
  20. ^ Kirstein L, B. P. Silfverskiold (1958). "A Family with Emotionally Precipitated Drop Seizures". Acta Psychiatrica Scandinavica. 33 (4): 471–6. doi: 10.1111/j.1600-0447.1958.tb03533.x. PMID  13594585. S2CID  143799581.
  21. ^ Kok O, G. W. Bruyn (1962). "An Unidentified Hereditary Disease". Lancet. 279 (7243): 1359. doi: 10.1016/S0140-6736(62)92475-3.

External links