Document Type : Review

Authors

1 Iranian Child Neurology Center of Excellence, Pediatric Neurology Department, Mofid Children’s Hospital, School of Medicine, Shahid Beheshti University of Medical Sciences (SBMU), Tehran, Iran

2 Pediatric Neurology Research Center, Research Institute for Children Health, Shahid Beheshti University of Medical Sciences, Tehran, Iran.

Abstract

Epilepsy is a common neurological disorder in childhood with prominent neurological manifestations, signs, and symptoms in inherited neurometabolic disorders. Accurate diagnosis of neurometabolic disorders in epileptic patients increases the possibility of a specific treatment to improve epilepsy. Therefore, early diagnosis is essential in potentially treatable epileptic disorders. Various seizure types occur in neurometabolic disorders, which are often refractory to antiepileptic drugs (without the treatment of the underlying neurometabolic disorders). Patients with underlying disorders have severe clinical presentations, such as refractory seizures. In addition, they do not respond to antiepileptic drugs in many cases. In the epileptic patients with developmental delay and/or regression, neurometabolic disorders should be considered in the presence of abnormal neurological examination and brain imaging with specific patterns. Some of these disorders are potentially treatable. Therefore, neurologists should determine the etiology of epilepsy, especially in pediatric patients, and the treatment should not be restricted to symptomatic therapy. The present study aimed to introduce some of the treatable causes of epilepsy in pediatric patients.

Keywords

Introduction
Epilepsy is a common neurological disorder in childhood with the global prevalence rate of 0.5-1% (1). Inherited metabolic disorders are individually rare, with an estimated prevalence of 1:1000 (2), and are associated with significant morbidity. More than 100 inherited neurometabolic disorders are presented with epilepsy (3). Therefore, attempts should be made for the accurate diagnosis of these potentially treatable disorders.
Treatable inherited neurometabolic disorders are of 10-15 types (4). The present study aimed to identify the neurological diseases that meet three criteria: 1) eminently treatable diseases; 2) clinical presentation of prominent epilepsy and 3) association of prognosis with early diagnosis and treatment intervention. The current investigation focused on the most common treatable epilepsies and other conditions in which early diagnosis and intervention could lead to better prognosis in children. Of note, the emphasis has been placed on the conditions in which early detection and treatment are crucial.
This review was conducted based on literature search, clinical experience, and expert opinions.

Literature Review
1-Pyridoxine-DependentSeizures/Pyridoxal Phosphate- and Folinic Acid-Responsive Seizures
Pyridoxine-dependent or pyridoxal phosphate (PLP)-responsive epilepsies must be considered in the neonates with unexplained, refractory seizures, beginning before or shortly after birth (1,2). Pyridoxine-dependent epilepsy is caused by mutations in the ALDH7A1 gene encoding the antiquitin (ATQ) protein, which is involved in lysine catabolism in the central nervous system (CNS) (3,4). ATQ deficiency leads to increased alpha-aminoadipic semialdehyde (α-AASA) and piperideine-6-carboxylic acid (P6C) (4,5). P6C has been shown to inactivate PLP, thereby leading to a secondary deficiency (4,5). The final dysfunctional pathway is brain gamma-aminobutyric acid (GABA) deficiency, which causes an imbalance between excitatory and inhibitory activities, reducing the epileptic threshold.

Clinical Features
Classically, pyridoxal-dependent epilepsy is presented as neonatal seizures, which are refractory to some conventional antiepileptic drugs (7). ATQ deficiency is characterized by the early onset of epileptic encephalopathy. Despite seizure control, most of these patients develop intellectual disabilities (2). These prolonged seizures are a fracture of status epilepticus or early myoclonic encephalopathy (3). Folinic acid-responsive seizures have been reported in a few newborns with encephalopathy and apneas within five days after birth (8).

Diagnosis
Both α-AASA and pipecolic acid are the diagnostic markers of ATQ deficiency. Furthermore, elevated α-AASA in the urine, plasma, and cerebrospinal fluid (CSF) is pathognomonic for this disorder (2).

Management
Intravenous administration of pyridoxine (100 mg) induces the cessation of epileptic seizures and electroencephalographic discharges within minutes. However, seizures may relapse, and pyridoxine administration may be repeated (total of 500 mg) within 24 hours or continued at 30/mg/kg for seven days in the case of a partial response (3).
Chronic therapy with oral PLP (30-50 mg/kg/day) may also induce seizure recovery (3). Furthermore, folinic acid-responsive seizures could be ceased with an additional dose of enteral folinic acid (3-5 mg/kg/day). Since mutations in the ALDH7A1 gene have been reported in these patients, administration of the adequate doses of pyridoxine has been proposed (6-8).

2-Glucose Transporter Type-1 Deficiency
Glucose is the most essential substrate for brain energy metabolism (10,11). At rest, the brain of infants and children consumes up to 80% of the total glucose supply in the body (12). Transport of glucose from the blood-brain barrier is performed by glucose transporter 1 (GLUT1), which is encoded by the GLUT1 gene on chromosome 1p34.2 (SLC2A1 gene) (13).

Clinical Features
Classically, patients with GLUT1 deficiency syndrome (Glut1-DS) present with epileptic encephalopathy, as well as various seizure types, developmental delay, acquired microcephaly, and movement disorders (14). Seizures are myoclonic (involuntary limb jerking with head bobbing), hypotonic, and unresponsive, along with eye-rolling and staring (14). The spectrum of epilepsy syndromes includes juvenile myoclonic epilepsy, childhood absence epilepsy, juvenile absence epilepsy, early-onset absence epilepsy, and focal epilepsy (15).

Diagnosis
In the patients with Glut1-DS, CSF glucose concentration is extremely low (16). Moreover, hypoglycorrhachia is observed in meningitis, meningeal carcinomatosis, subarachnoid hemorrhage, prolonged seizures or status epilepticus, and mitochondrial diseases (17,18). In suspected cases, the CSF/serum ratio of 0.33 to 0.37 is considered diagnostic for Glut1-DS (19). CSF lactate level of less than 1.4 mmol/l is another marker of Glut1-DS (20,21), and the results of a study in this regard showed that CSF lactate never elevates in Glut1-DS (22).

Management
Epilepsy in the patients with Glut1-DS could be effectively treated with a ketogenic diet (22). This syndrome mimics the metabolic state of fasting, while maintaining ketosis and utilizing nutritional fat rather than the body fat. Prolonged hypoglycorrhachia may lead to irreversible brain damage (15). Ketones serve as an alternative fuel to the brain in the presence of hypoglycorrhachia (22). A ketogenic diet could improve myelination and prevent brain damage in these patients (23). Additionally, alpha lipoic acid and triheptanoin are among the potential supplementary Glut1-DS therapies (23).

3-Hyperinsulinism/Ammonemia (HI/HA)
This syndrome presents as serum hypoglycemia and hyperammonemia. With autosomal dominant inheritance, this syndrome is provoked by fasting or high-protein diets (24,25). The disorder is caused by the mutations in glutamate dehydrogenase (GDH), which is a mitochondrial enzyme (26).

Clinical Features
One of the major clinical manifestations of HI/HA in recurrent hypoglycemia, which could be accompanied by hyperammonemia (27,28). Fasting causes hypoglycemic attacks, and the most common neurological manifestations in the patients include cognitive decline and seizures due to hypoglycemia or its complications (29,30).

Diagnosis
Elevated plasma ammonia concentrations (>35 µmol⁄lL) and glycemic response to glucagon in the presence of hypoglycemia (29,31,32) are noted in the patients with HI/HA. Genetic or enzyme assays are essential to confirming the diagnosis of the hyperinsulinism/hyperammonemia syndrome. Biochemical studies have included the evaluation of GDH activity and GTP inhibition of GDH activity in lymphoblast homogenates (31).

Management
HI/HA could be controlled by diazoxide treatment at the dosage of 10-15 mg/kg/day (33).

4-Developmental Delay, Epilepsy, Neonatal Diabetes (DEND)
Potassium channels are important regulators of tissue excitation, and their activation damps down electrical activity and causes neuron membrane repolarization (34). Mutations in the KATP channel are associated with several metabolic syndromes (35). Severe gain-of-function mutations lead to developmental delay, epilepsy, and neonatal diabetes, also known as the DEND syndrome.

Clinical Features
Neonatal diabetes mellitus (NDM) is defined as insulin-requiring hyperglycemia, which is often diagnosed within the first three months of life. NDM is a rare disorder with an incidence of 1:300,000-500,000 live births (36). DEND syndrome is a form of permanent-NDM (PNDM) associated with developmental delay, epilepsy, and muscle weakness (37). It is an autosomal recessive disorder caused by the activated mutations in the KATP channel subunit Kir6.2 (potassium inverse rectifying channel 6.2, encoded by KCNJ11) or SUR1 gene (sulfonylurea receptor1, encoded by ABCC8) located on chromosome 11 (38). Some of the dysmorphic features in DEND include prominent metopic suture, bilateral ptosis, downturned mouth, and contractures (39).

Diagnosis
Diagnosis of DEND could be confirmed by genetic mutation testing (38).

Management
Glibenclamide treatment could enhance the mental function, motor function, and glucose homeostasis of the patients with DEND (40).

5-Hyperekplexia, Startle Disease, and Stiff Infant Syndrome
Hyperekplexia is an autosomal dominant disorder defined by an exaggerated startle reaction in response to unexpected, sudden tactile or auditory stimuli (41-43).

Clinical Features
Symptoms may occur due to severe opisthotonic posturing and hypertonia in the neonatal period. In addition, head retraction may be observed in older children (44). The stiffness disappears spontaneously during infancy, and most children become normal by the age of three. In adolescence, stiffness attacks may recur in response to cold exposure and startle reactions (45).

Diagnosis
Family history of startle disease is considered essential to diagnosis. Unlike startle-provoked epilepsy, electroencephalography is normal in stiff infant syndrome (45). The disorder is associated with mutations in the α-1 subunit of the glycine receptor (46).

Management
Clonazepam, levetiracetam, and sodium valproate are the most effective agent to reduce these attacks (45). In addition, the combination of clonazepam and clobazam could be effective in the ambulation and elimination of the falls (47).

6-Creatine Synthesis Disorders
Creatine kinase converts creatine into creatine phosphate (48). Creatine is an essential material for the brain. In this regard, one syndrome is focused on creatine transport, and two other syndromes are focused on creatine synthesis. Creatine synthesis disorders are caused by the defects in the guanidinoacetic acid methyltransferase (GAMT) or arginine glycine acyl transferase (AGAT) enzymes (48,49). Creatine deficiency in the brain is caused by the deficiency in creatine transporter gene (SLC6A8) (50).

Clinical Features
Seizures, mental retardation, and speech delay are the main symptoms associated with this syndrome (49). Moreover, global developmental delay is observed prior to 12 months of age. Speech delay and receptive language deficiencies are the other severe symptoms in this syndrome (48). Severe types of this syndrome are associated with complications such as the abnormal signal changes of the basal ganglia, extrapyramidal movement disorders, early global developmental delay, and refractory seizures (49).

Diagnosis
Cerebral creatine deficiency syndrome could be diagnosed based on the abnormal guanidine-acetic acid (GAA) levels in the urine or plasma (51,55). Magnetic resonance spectroscopy reveals severely reduced creatine peak (52).

Management
Cerebral creatine deficiency syndrome responds to the supplementations of creatine and ornithine with arginine restriction (51). Creatine doses in the patients diagnosed with AGAT deficiency have been determined at 100-800 mg/kg/day (3-4 doses daily) (53). Early treatment with 100-800 mg/kg/day creatine supplementation could prevent the cognitive deficits associated with GAMT disorder (48,54). In addition, sodium benzoate (100 mg/kg/day) and ornithine (100-700 mg/kg/day) are recommended for GAA reduction (54,55).

7-Serine Biosynthesis Defects
Serine is an amino acid that is absorbed from dietary protein (56). It is synthesized by serine hydroxymethyltransferase from the conversion of glycine (57-59).

Clinical Features
Patients with serine deficiency present with severe neurological symptoms, including intractable seizures, psychomotor retardation, and congenital microcephaly (56). Furthermore, serine deficient children have neurodevelopmental delay and variable clinical seizure patterns (60). Seizures begin as flexor spasms with West syndrome or generalized tonic-clonic seizures (61). In some cases, bursts of inappropriate laughing suggest that gelastic seizures are present as well (62).

Diagnosis
Low plasma level of serine and low CSF concentrations of serine, glycine, and 5-methyltetrahydrofolate (5-MTHF) are the diagnostic markers for this type of seizures (60,63).

Management
High doses of serine (200-600 mg/kg/day) are recommended for the patients with serine deficiency (64). Serine has been shown to prevent epilepsy and cause normal neurodevelopment in the siblings of these patients (60). Moreover, treatment with serine (500 mg/kg/d) and glycine (200 mg/kg/d) could normalize the plasma and CSF concentrations (64).

8-Biotinidase Deficiency
Biotinidase deficiency is an autosomal recessive disease with the prevalence of 1:60,000 in normal populations (65,66). The main cause of biotinidase deficiency is problematic biotin absorption, which was previously known as late-onset multiple holocarboxylase deficiency (67,68).

Clinical Features
Major clinical phenotypes of biotinidase deficiency include developmental delay, hypotonia, seizures, ataxia, alopecia, perioral rashes, sensorineural hearing loss, vision loss, optic atrophy, lactic academia (51,65,66), myopathy/peripheral neuropathy, and cell immunity disorders (66).

Diagnosis
Patients with biotinidase deficiency present with ketosis, lactic acidosis, hypoglycemia, and hyperammonia (69). In severe cases, enzyme activity is less than 10%, and in partial enzyme deficiency, biothinidase activity is 10-30% (70).

Management
Children with severe biotinidase deficiency are treated with the pharmacologic doses of biotin (5-20 mg/day) (71). Early diagnosis and treatment could significantly reduce the associated complications, resulting in excellent prognosis (65).

9-Cerebral Folate Deficiency
Infants with cerebral folate deficiency normally present with progressive neurological regression at the age of 4-6 months (72,73).

Clinical Features
Cerebral folate deficiency is associated with severe irritability, reduction of the head circumference, and developmental regression, which lead to profound cognitive decline, blindness, ataxia, and spastic quadriplegia (69).

Diagnosis
Diagnosis of cerebral folate deficiency is confirmed by the low levels of folate metabolite and methyltetrahydrofolate (5 MTHF) in the CSF despite normal serum folate (72-74).

Management
Treatment with oral folinic acid (0.5-1 mg/kg/day) for a minimum of one year has been shown to normalize the levels of 5 MTHF in the CSF and result in neurological recovery (69,75).

10-Biopterin Synthesis Disorders
Tetrahydrobiopterin (BH4) synthesis deficiency is a malignant type of hyperphenylalaninemia (76). BH4 deficiency is associated with hyperphenylalaninemia in 1-3% of the patients (69). In addition, BH4 is known as the phenylalanine-hydroxylase cofactor (77).

Clinical Features
Major phenotypes of BH4 disorders include intellectual disability, myoclonic seizures, muscular rigidity, dystonia, and microcephaly (51).

Diagnosis
This condition could be diagnosed in the neonatal period based on the elevated serum phenylalanine level (69,76,77). Other diagnostic measures in this regard include the measurement of dihydropteridine reductase in dried blood spots, neurotransmitters and pterins in the CSF, pterin metabolites in urine, and reduced enzyme activity (77).

Management
BH4 does of 5-10 mg/kg is the significant dose for the correction of peripheral hyperphenylalaninemia, and 10-20 mg/kg/day is the required dose for the correction of classical phenylketonuria (69,76). Furthermore, lifelong supplementation with L-DOPA, 5-hydroxytryptophan, and carbidopa is considered essential for the treatment of hyperphenylalaninemia (69,78-81). Intracranial calcifications in these patients may be reversible with folinic acid treatment (51).

11-Methylmalonic Acidemia (MMA)
Methylmalonic acidemia (MMA) in neonates is presented with the rapid deterioration of the newborn after a short symptom-free interval (8,82). In such cases, seizures may complicate acute metabolic decompensation (83).

Clinical Features
Children with MMA may appear normal at birth. The symptoms often manifest during the first week of birth in 80% of the cases with complete mutase deficiency (67). After protein feeding, the infant may have recurrent vomiting, dehydration, hypotonia, lethargy, respiratory distress, and failure to thrive. More than 50% of the patients with MMA present with anemia, leukopenia, and thrombocytopenia (67, 69). In addition, intracranial hemorrhage has been reported in MMA (84).

Diagnosis
Analysis of urinary organic acids is essential to the diagnosis of MMA (69). Neuroimaging in MMA may show swelling, delayed myelin maturation, focal necrosis of the globus pallidus, calcification of the basal ganglia, and volume loss (83).

Management
The aim of MMA treatment is to limit catabolism and restrict protein intake (0.5-l.5 g/kg/day) in the acute cases. Moreover, intramuscular hydroxocobalamin (1 mg/week), L-carnitine, betaine, and folate are typically used in the treatment of MMA (69,86-88).

12-Glutaric Aciduria Type 1 (GA1)
Glutaric aciduria type 1 (GA1) is an inborn error of lysine, hydroxylysine, and tryptophan catabolism (8,89).

Clinical Features
In late infancy, GA1 is mainly presented as acute encephalopathy with predominant dyskinesia and dystonia due to the necrosis of the basal ganglia, particularly the putamina (90). In GA1, neonates present with agitation, irritability, macrocephaly, and hypotonia. Seizures normally occur within the context of acute decompensation in association with the symptoms of rapid deterioration. Dyskinetic movements may often be misdiagnosed for seizures (91).

Diagnosis
Urinary organic acids profile shows increased 3-OH-glutaric acid and glutarylcarnitine as the major peak in GA1 (92). Neuroimaging typically shows enlarged frontotemporal CSF spaces, wide Sylvian fissures, and occasional subdural hematoma (Figure 1) (93,94).

 

Figure 1. A patient with glutaric aciduria type 1. Axial T1WI MRI shows frontotemporal atrophy, resulting in wide sylvian fissures with right subdural hematoma.

Management
The basic treatments for GA1 involve dietary protein supplementation, lysine restriction, and carnitine and riboflavin supplementation (69,92).

13-Homocystinuria
Increased excretion of homocysteine is noted in several inborn errors of methionine metabolism, which is mainly caused by cystathionine β synthase deficiency (95).

Clinical Features
Clinical features of homocystinuria include vascular occlusive disease, malar flush, osteoporosis, genu valgum, pes cavus, marfanoid appearance, mental retardation, seizures, typical strokes, dystonia, and psychiatric abnormalities (96).

Diagnosis
Diagnosis of homocystinuria is based on the accumulation of serum homocysteine and methionine (67,96).

Management
All the patients with homocystinuria need pyridoxine (vitamin B6) prior to treatment (67). Pyridoxine, folic acid, and vitamin B12 have been used in pyridoxine non-responders as the cofactors of methionine metabolism (69). Homocysteine concentrations may also decrease in B6-unresponsive patients through betaine treatment (86-280 mg/kg) (96).

14-Ketogenic Diet-Responsive Epilepsies
GLUT1 deficiency has been described in detail in the previous sections of the article (11-23).

Pyruvate Dehydrogenase Complex Deficiency
The clinical signs associated with pyruvate dehydrogenase complex deficiency include hypotonia/hypertonia, seizures, microcephaly, ataxia, respiratory distress, facial dysmorphism, spasticity, peripheral neuropathy, optic atrophy, nystagmus, ptosis, and strabismus (97-100). Patients are normally treated with a high-fat (>55%), low-carbohydrate (ketogenic) diet and thiamine administration (100-600 mg/day) (67).

Conclusion
Epilepsy is among the most common clinical features in various neurogenetic and neurometabolic disorders. Etiological findings play a pivotal role in the treatment of epilepsy, especially in pediatric epileptic patients. In the present study, some treatable causes of epilepsy in pediatric patients were explored. In the majority of these epileptic disorders, treatment of the underlying causes of the disease could effectively control epilepsy. Therefore, it is recommended that pediatric neurologists investigate the etiologies in pediatric epileptic patients, while symptomatic therapy is not considered to be a viable option in this regard.

Acknowledgements
We appreciate our patients, who gave us a grate opportunity to perceive more scientific information about neurometabolic diseases due to their clinical and paraclinical examinations.

Conflict of Interest
The authors declare no conflict of interest.

  1. Hunt ad JR, Stokes J Jr, McCRORY WW, et al. Pyridoxine dependency: report of a case of intractable convulsions in an infant controlled by pyridoxine. Pediatrics. 1954;13:140-145.
  2. Stockler S, Plecko B, Gospe SM Jr, et al. Pyridoxine dependent epilepsy and antiquitin deficiency: clinical and molecular characteristics and recommendations for diagnosis, treatment and follow-up. Mol Genet Metab. 2011;104:48-60.
  3. Fernandes J, Saudubray JM, Van den Berghe G. Inborn Metabolic Diseases: Diagnosis and Treatment. Berlin: Springer; 2006.
  4. Mills PB, Struys E, Jakobs C, et al. Mutations in antiquitin in individuals with pyridoxine-dependent seizures. Nat Med. 2006;12:307-309.
  5. Plecko B, Stöckler-Ipsiroglu S, Paschke E, et al. Pipecolic acid elevation in plasma and cerebrospinal fluid of two patients with pyridoxine-dependent epilepsy. Ann Neurol. 2000;48:121-125.
  6. Gallagher RC, Van Hove JL, Scharer G, et al. Folinic acid-responsive seizures are identical to pyridoxine dependent epilepsy. Ann Neurol. 2009;65:550-556.
  7. Gospe SM. Neonatal Vitamin-responsive Epileptic Encephalopathies. Chang Gung Med J. 2010;33:1-12.
  8. Parisi E, Nicotera A, Alagna A, et al. Neonatal Seizures and Inborn Errors of Metabolism: An Update. Int J Pediatr Neonat Care. 2015; 1: 111.
  9. Laoprasert P. Atlas of Pediatric EEG.1st ed. New York:Mc Graw Hill; 2011.
  10. Wolf NI, Bast T, Surtees R, et al. Epilepsy in inborn errors of metabolism. Epileptic Disord. 2005;7:67-81.
  11. Klepper J. Glucose transporter deficiency syndrome (GLUT1DS) and the ketogenic diet.Epilepsia. 2008;49:46-49.
  12. Cremer JE. Substrate utilization and brain development. J Cereb Blood Flow Metab. 1982;2:394-407.
  13. Pong AW, Geary BR, Engelstad KM, et al. Glucose transporter type I deficiency syndrome: Epilepsy phenotypes and outcomes. Epilepsia. 2012;53:1503-1510.
  14. De Giorgis V, Veggiotti P. GLUT1 deficiency syndrome 2013: Current state of the art. Seizure. 2013;22:803-811.
  15. Verrotti A, D’Egidio C, Agostinelli S, et al. Glut1 deficiency: When to suspect and how to diagnose? Eur J Paediatr Neurol. 2012;16:3-9.
  16. Rostein M, Englestad K, Yang H, et al. Glut1 deficiency: inheritance pattern determined by haploinsufficiency. Ann Neurol. 2010;68:955-958.
  17. Silver TS, Todd JK. Hypoglycorrhachia in pediatric patients. Pediatrics. 1976;58:67-71.
  18. Huang HR, Chen HL, Chu SM. Clinical spectrum of meningococcal infection in infants younger than six months of age. Chang Gung Med J. 2006;29:107-113.
  19. Aktas D, Utine EG, Mrasek K, et al. Derivative chromosome 1 and GLUT1 deficiency syndrome in a sibiling pair. Mol Cytogenet. 2010;3:10.
  20. De Vivo DC, Trifiletti RR, Jacobson RI, et al. Defective glucose transport across the blood–brain barrier as a cause of persistent hypoglycorrhachia, seizures, and developmental delay. N Engl J Med. 1991;325:703-709.
  21. De Vivo DC, Leary L, Wang D. Glucose transporter 1 deficiency syndrome and other glycolytic defects. J Child Neurol. 2002;17:3S15-23.
  22. Klepper J. GLUT1 deficiency syndrome in clinical practice. Epilepsy Res. 2012;100:272-277.
  23. Klepper J, Engelbrecht V, Scheier H. GLUT1 deficiency with delayed myelination responding to ketogenic diet. Pediatr Neurol. 2007;37:130-133.
  24. Corrêa-Giannella ML, Freire DS, Cavaleiro AM, et al. Hyperinsulinism/hyperammonemia (HI/HA) syndrome due to a mutation in the glutamate dehydrogenase gene.Arq Bras Endocrinol Metabol. 2012 Nov;56(8):485-9.
  25. Cochrane WA, Payne WW, Simpkiss MJ, et al. Familial hypoglycemia precipitated by amino acids. J Clin Invest. 1956;35:411-422.
  26. Stanley CA. Hyperinsulinism/hyperammonemia syndrome: insights into the regulatory role of glutamate dehydrogenase in ammonia metabolism. Mol Genet Metab. 2004;81 Suppl 1:S45-51.
  27. Stanley CA. The hyperinsulinism-hyperammonemia syndrome: gain of function mutations of glutamate dehydrogenase. In: Dunger DB, editor. Genetic insights in paediatric endocrinology and metabolism. Bristol: BioScientifica; 2000. p. 23–30.
  28. MacMullen C, Fang J, Hsu BYL, et al. Hyperinsulinism/Hyperammonemia syndrome in children with regulatory mutations in the inhibitory GTP binding domain of glutamate dehydrogenase. J Clin Endocrinol Metab. 2001;86:1782-1787.
  29. Hsu BY, Kelly A, Thornton PS, et al. Protein-sensitive and fasting hypoglycemia in children with the hyperinsulinism/hyperammonemia syndrome. J Pediatr. 2001;138:383-389.
  30. Bahi-Buisson N, Roze E, Dionisi C, et al. Neurological aspects of hyperinsulinism– hyperammonaemia syndrome. Dev Med Child Neurol. 2008;50:945-949.
  31. Stanley CA, Lieu YK, HsuBYL, et al. Hyperinsulinism and hyperammonemia in infants with regulatory mutations of the glutamate dehydrogenase gene. N Engl J Med. 1998;338:1352-1357.
  32. De Lonlay P, Benelli C, Fouque F, et al. Hyperinsulinism and hyperammonemia syndrome: report of twelve unrelated patients. Pediatr Res. 2001;50:353-357.
  33. Palladino AA, Stanley CA. The hyperinsulinism/hyperammonemia syndrome. Rev Endocr Metab Disord. 2010;11:171-178.
  34. Lahmann C, Ashcroft F. DEND Syndrome: Developmental Delay, Epilepsy, and Neonatal Diabetes, a Potassium Channelopathy. In: Pearl PL, editor. Inherited Metabolic Epilepsies. New York, NY, USA: Demos Medical; 2013.p. 189-200.
  35. Hattersley AT, Ashcroft FM. Activating mutations in Kir6.2 and neonatal diabetes: new clinical syndromes, new scientific insights, and new therapy. Diabetes. 2005;54:2503-2513.
  36. PolakM, Shield J. Neonataldiabetesmellitus-genetic aspects. Pediatr Endocrinol Rev. 2004;2:193-198.
  37. Singh P, Rao SC, Parikh R. Neonatal Diabetes with Intractable Epilepsy: DEND Syndrome. Indian J Pediatr. 2014;81:1387-1388.
  38. Edghill EL, Gloyn AL, Gillespie KM, et al. Activating mutations in the KCNJ11 gene encoding the ATP-sensitive K+ channel subunit Kir6.2 are rare in clinically defined type 1 diabetes diagnosed before 2 years. Diabetes. 2004;53:2998-3001.
  39. Gloyn AL, Pearson ER, Antcliff JF, et al. Activating Mutations in the Gene Encoding the ATP-Sensitive Potassium-Channel Subunit Kir6.2 and Permanent Neonatal Diabetes. N Engl J Med. 2004;350:1838-1849.
  40. Mlynarski W, Tarasov AI, Gach A, et al. Sulfonylurea improves CNS function in a case of intermediate DEND syndrome caused by a mutation in KCNJ11. Nat Clin Pract Neurol. 2007;3:640-645.
  41. Ryan SG, Dixon MJ, Nigro MA, et al. Genetic and Radiation Hybrid Mapping of the Hyperekplexia Region on Chromosome 5q. Am J Hum Genet. 1992;51:1334-1343.
  42. Andermann F, Keene DL, Andermann E, et al. Startle disease or hyperekplexia: future delineation of the syndrome. Brain. 1980 ;103:985-997.
  43. Praveen V, Patole SK, Whitehall JS. Hyperekplexia in neonates. Postgrad Med J. 2001;77:570-572.
  44. Swaiman KF, Ashwal S, Ferriero DM, et al. Swaiman’s Pediatric Neurology Principles and Practice. 5th ed. London: Elsevier Saunders; 2012.
  45. Piña-Garza JE. Fenichel’s Clinical Pediatric Neurology A Signs and Symptoms Approach. 7th ed. London: ElsevierSaunders; 2013.
  46. Shiang R, Ryan SG, Zhu YZ, et al. Mutations in the alpha1 subunit of the inhibitory glycine receptor cause the dominant neurologic disorder, hyperekplexia. Nat Genet. 1993;5:351-358.
  47. McAbee GN. Clobazam-Clonazepam Combination Effective for Stimulus-Induced Falling in Hyperekplexia. J Child Neurol. 2015;30:91-92.
  48. Clark JF, Cecil KM. Diagnostic methods and recommendations for the cerebral creatine deficiency syndromes. Pediatr Res. 2015;77:398-405.
  49. Sykut-Cegielska J, Gradowska W, Mercimek-Mahmutoglu S, et al. Biochemical and clinical characteristics of creatine deficiency syndromes. Acta Biochim Pol. 2004;51:875-882.
  50. Salomons GS, van Dooren SJ, Verhoeven NM, et al. X-linked creatine transporter gene (SLC6A8) defect: a new creatine-deficiency syndrome. Am J Hum Genet. 2001;68:1497-14500.
  51. Pearl PL. Inherited Metabolic Epilepsies: The Top 10 Diagnoses You Cannot Afford to Miss. Demos Medical Publishing 2013.
  52. Barkovich AJ, Raybaud C. Pediatric neuroimaging. 5th ed. Lippincott Williams & Wilkins: Philadelphia; 2012.
  53. Battini R, Alessandrì MG, Leuzzi V, et al. Arginine:glycineamidinotransferase (AGAT) deficiency in a newborn: early treatment can prevent phenotypic expression of the disease. J Pediatr. 2006;148:828-830.
  54. Stockler-Ipsiroglu S, van Karnebeek C, Longo N, et al. Guanidinoacetate methyltransferase (GAMT) deficiency: outcomes in 48 individuals and recommendations for diagnosis, treatment and monitoring. Mol Genet Metab. 2014;111:16-25.
  55. Mercimek-Mahmutoglu S, Stoeckler-Ipsiroglu S, Adami A, et al. GAMT deficiency: features, treatment, and outcome in an inborn error of creatine synthesis. Neurology. 2006;67:480-484.
  56. Tabatabaie L, Klomp LW, Berger R, et al. L-Serine synthesis in the central nervous system: A review on serine deficiency disorders. Mol Genet Metab. 2010;99:256-262.
  57. Ichihara A, Greenberg DM. Pathway of serine formation from carbohydrate in rat liver. ProProc Natl Acad Sci U S A. 1955;41:605-609.
  58. Snell K. Enzymes of serine metabolism in normal, developing and neoplastic rat tissues. Adv Enzyme Regul. 1984;22:325-400.
  59. Van der Crabben SN, Verhoeven-Duif NM, Brilstra EH, et al. An update on serine deficiency disorders.J Inherit Metab Dis. 2013;36:613-619.
  60. Swaiman KF, Ashwal S, Ferriero DM, et al. Swaiman’s Pediatric Neurology Principles and Practice. 5th ed. London: Elsevier Saunders;2012.
  61. de Koning TJ, Klomp LWJ. Serine-deficiency syndromes.Curr Opin Neurol. 2004;17:197-204.
  62. de Koning TJ, Poll-The BT, Jaeken J. Continuing education in neurometabolic disorders: serine deficiency disorders. Neuropediatrics.1999;30:1-4.
  63. Moat S, Carling R, Nix A, et al. Multicentre age-related reference intervals for cerebrospinal fluid serine concentrations: Implications for the diagnosis and follow-up of serine biosynthesis disorders. Mol Genet Metab. 2010;101:149-152.
  64. Hart CE, Race V, Achouri Y, et al. Phosphoserine Aminotransferase Deficiency: A Novel Disorder of the Serine Biosynthesis Pathway. Am J Hum Genet. 2007;80:931-937.
  65. Akhondian J, Ashrafzadeh A, BeiraghiM, et al. A Treatable Refractory Epilepsy: A Case Report. Int J Pediatr. 2014:2;93-96.
  66. Wolf B. The neurology of biotinidase deficiency. Mol Genet Metab. 2011;104:27-34.
  67. Piña-Garza JE. Fenichel’s Clinical Pediatric Neurology A Signs and Symptoms Approach. 7th ed. London: ElsevierSaunders;2013.
  68. Wolf B. Disorders of biotin metabolism. In: Scriver CR, Beaudet AL, Sly WS, Valle D, editors. The Metabolic and Molecular Bases of Inherited Disease. McGraw-Hill: New York; 2001.p. 3935–3962.
  69. Swaiman K, Ashwal S, Ferriero DM, et al. Swaiman’s Pediatric Neurology Principles and Practice. 5th ed. London: Elsevier Saunders; 2012.
  70. Cowan TM, Blitzer MG, Wolf B. Technical standards and guidelines for the diagnosis of biotinidase deficiency. Genet Med. 2010;12:464-470.
  71. Wolf B. Clinical issues and frequent questions about biotinidase deficiency. Mol Genet Metab. 2010;100:6-13.
  72. Ramaekers VT, Rothenberg SP, Sequeira JM, et al. Autoantibodies to Folate Receptors in the Cerebral Folate Deficiency Syndrome. N Engl J Med. 2005;352:1985-1991.
  73. Perez-Duenas B, Ormazbal A, Toma C, et al. Cerebral Folate Deficiency Syndromes in Childhood Clinical, Analytical, and Etiologic Aspects. Arch Neurol. 2011;68:615-621.
  74. RamaekersV ,Sequeira JM, Quadros EV. Clinical recognition and aspects of the cerebral folate deficiency syndromes.Clin Chem Lab Med. 2013;51:497-511.
  75. Gordon N. Cerebral folate deficiency. Dev Med Child Neurol. 2009;51:180-182.
  76. Piña-Garza JE. Fenichel’s Clinical Pediatric Neurology A Signs and Symptoms Approach. 7th ed. London:Elsevier Saunders;2013.
  77. Longo N. Disorders of biopterin metabolism. J Inherit Metab Dis. 2009;32:333-342.
  78. Dulac O, Plecko B, Gataullina S, et al. Occasional seizures, epilepsy, and inborn errors of metabolism. Lancet Neurol. 2014;13:727-739.
  79. Blau N, Bélanger-Quintana A, Demirkol M, et al. Optimizing the use of sapropterin (BH4) in the management of phenylketonuria. Mol Genet Metab. 2009;96:158-163.
  80. Leuret O, Barth M, Kuster A, et al. Efficacy and safety of BH4 before the age of 4 years in patients with mild phenylketonuria. J Inherit Metab Dis. 2012;35:975-981.
  81. MenkesJH, Sarnat HB, Maria BL.Child neurology. 7th ed. Philadelphia:Lippincott Williams &Wilkins;2006.
  82. Van Gosen L. Organic acidemias: a methylmalonic and propionic focus. J Pediatr Nurs. 2008;23:225-233.
  83. Poretti A, Blaser SI, Lequin MH, et al. Neonatal neuroimaging findings in inborn errors of metabolism. J Magn Reson Imaging. 2013;37:294-312.
  84. Dave P, Curless RG, Steinman L. Cerebellar hemorrhage complicating methylmalonic and propionic acidemia. Arch Neurol. 1984;41:1293-1296.
  85. Ogier de Baulny H, Gérard M, Saudubray JM, ET AL. Remethylation defects: guidelines for clinical diagnosis and treatment. Eur J Pediatr. 1998;157 Suppl 2:S77-83.
  86. Roe CR, Hoppel CL, Stacey TE, et al. Metabolic response to carnitine in methylmalonic aciduria. An effective strategy for elimination of propionyl groups. Arch Dis Child. 1983;58:916-20.
  87. Rosenblatt DS, Thomas IT, Watkins D, Cooper BA, et al. Vitamin B12 responsive homocystinuria and megaloblastic anemia: Heterogeneity in methylcobalamin deficiency. Am J Med Genet. 1987;26:377-383.
  88. Thompson GN, Walter JH, Bresson JL, et al. Sources of Propionate in Inborn Errors of Propionate Metabolism. Metabolism. 1990;39:1133-1137.
  89. Kölker S, Christensen E, Leonard JV, et al. Guideline for the diagnosis and management of glutaryl-CoA dehydrogenase deficiency (glutaric aciduria type I). J Inherit Metab Dis. 2007;30:5-22.
  90. Hoffmann GF, Athanassopoulos S, Burlina AB, et al. Clinical course, early diagnosis, treatment, and prevention of disease in glutaryl-CoA dehydrogenase deficiency. Neuropediatrics. 1996;27:115-123.
  91. Hunt AD Jr, Stokes J Jr, Mc Crory WW ,et al. Pyridoxine dependency: report of a case of intractable convulsions in an infant controlled by pyridoxine. Pediatrics. 1954;13:140-145.
  92. Fernandes J, Saudubray JM, Berghe Van den G. Inborn Metabolic Diseases: Diagnosis and Treatment. 4thed. New York: Springer-Verlag Berlin Heidelberg; 2006.
  93. Poretti A, Blaser SI, Lequin MH, et al. Neonatal neuroimaging findings in inborn errors of metabolism. J Magn Reson Imaging. 2013;37:294-312.
  94. Barkovich AJ, Koch BL, Moore KR. Diagnostic imaging: pediatric neuroradiology. 2nd ed. Philadelphia: Elsevier;2015.
  95. Sacharow SJ, Picker JD, Levy HL. Homocystinuria caused by cystathioninebeta-synthase deficiency. In: Pagon RA, Bird TD,Dolan CR, et al, editors. GeneReviews™. Seattle: University of Washington. Seattle; 1993-2018. PMID: 20301697.
  96. Nyhan WL, Barshop BA,Ozand PT. Atlas of Metabolic Diseases. 2nd ed. CRC Press: Oxford University Press;2005.
  97. Dulac O, Plecko B, Gataullina S, et al. Occasional seizures, epilepsy, and inborn errors of metabolism. Lancet Neurol. 2014;13:727-739.
  98. Prasad C, Rupar T, Prasad AN.Pyruvate dehydrogenase deficiency and epilepsy. Brain Dev. 2011;33:856-865.
  99. DeBrosseSD, Okajima K, Zhang S, Nakouzet al. Spectrum of neurological and survival outcomes in pyruvate dehydrogenase complex(PDC) deficiency: Lack of correlation with genotype. Molecular Genetics and Metabolism 2012: 107; 394–402. Mol Genet Metab. 2012;107:394-402.
  100. Patel KP, O’Brien TW, Subramony SH, et al. The spectrum of pyruvate dehydrogenase complex deficiency: Clinical, biochemicaland genetic features in 371 patients.Mol Genet Metab. 2012;105:34-43.