ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Aspartoacylase deficiency (Canavan disease)

Aspartoacylase deficiency (Canavan disease)
Literature review current through: Jan 2024.
This topic last updated: Jan 19, 2024.

INTRODUCTION AND BACKGROUND — Aspartoacylase deficiency (Canavan disease; MIM #271900) is an autosomal recessive spongiform leukodystrophy that is prevalent in, but not restricted to, Ashkenazi Jewish individuals. The disease typically begins in infancy and is marked by relentless progression.

Canavan disease was first described in the early 20th century as spongy degeneration of central nervous system myelin in infancy. The discovery of N-acetylaspartic aciduria due to aspartoacylase deficiency in a leukodystrophy was made in 1987 [1]. The realization that this leukodystrophy was actually Canavan disease with abnormal N-acetylaspartic acid (NAA) metabolism was not made until 1988 [2]. Elevated urinary NAA excretion and aspartoacylase deficiency had been reported two years previously, but without a direct association to spongy degeneration of myelin in infancy [1]. The aspartoacylase gene was cloned in 1993, and numerous pathogenic variants have been identified since then [3-6].

Although the Canavan disease eponym is widely used, aspartoacylase deficiency is preferable. Of interest, Canavan's own report was linked to Schilder disease or Krabbe disease (galactocerebrosidase deficiency) and not spongy degeneration.

ETIOLOGY — Aspartoacylase deficiency is caused by pathogenic variants in the ASPA gene that encodes the enzyme aspartoacylase. The resulting deficiency of aspartoacylase leads to accumulation of N-acetylaspartic acid (NAA) in the brain and to oligodendrocyte dysfunction, spongiform changes, and absence of myelin. However, the precise mechanisms causing spongiform degeneration are uncertain.

Genetics — Aspartoacylase deficiency is transmitted in an autosomal recessive fashion. The ASPA gene encoding aspartoacylase is located on chromosome 17pter-p13 [4]. Several pathogenic variants have been defined in ASPA, but just four of them account for >99 percent of aspartoacylase deficiency cases in Ashkenazi Jews [3,5,7-10].

In non-Ashkenazi individuals, a broad range of distinctly different pathogenic variants, including large deletions, have been identified [8-15]. Unique ASPA pathogenic variants were found in the Japanese and Scandinavian populations [16,17], as well as in Northern and Southern Indian populations [18,19]. A novel homozygous pathogenic variant, associated with a milder juvenile-onset form of aspartoacylase deficiency, was reported in the Telugu Devanga Chettiar community from southern India [20].

Biochemistry — NAA, formed from acetyl-CoA and aspartic acid, is the second most prevalent free amino acid in the brain. It is localized to neurons, where it is synthesized within mitochondria and transferred to oligodendrocytes via axoglial contact zones between the innermost oligodendrocyte plasma membrane and the axonal membrane [21]. In neurons, NAA is converted to N-acetylaspartylglutamate (NAAG) and is taken up by astrocytes, where it is hydrolyzed to NAA and glutamate. The NAA is then taken up by oligodendrocytes, the primary location of aspartoacylase [22-24].

The synthesis of NAA in neuronal mitochondria is mediated by aspartate N-acetyltransferase (ANAT), which is encoded by NAT8L. The export of NAA from neurons may occur via anion channels and transporters along the very high intracellular-extracellular gradient. The high-affinity sodium-dependent dicarboxylate cotransporter 3, or NaDC3, which is encoded by SLC13A3, has been proposed to be involved in the uptake of NAA into astrocytes and oligodendrocytes. Astrocyte-oligodendrocyte gap junctions may also be involved in NAA transport [25].

Aspartoacylase catalyzes the conversion of NAA to aspartate (aspartic acid) and acetate. By mechanisms that are largely unknown, deficiency of aspartoacylase leads to oligodendrocyte dysfunction, the prominent spongiform changes, and absence of myelin [26,27]. Levels of NAA are markedly increased in plasma, urine, and cerebrospinal fluid [28,29].

Loss-of-function mutations of the ASPA gene lead to structural changes, including decreased thermal or conformational stability, which result in diminished or nearly absent enzymatic activity [30,31].

Pathogenesis — Abnormal myelination, with associated prominence of swollen, vacuolated astrocytes, is a fundamental hallmark of aspartoacylase deficiency. However, the specific role of NAA in the pathogenesis of this disease is unknown [32-34]. Multiple hypotheses for the pathogenic mechanisms underlying these findings have been proposed:

One suggestion is that a reduction in aspartic acid as a result of aspartoacylase deficiency adversely affects the recycling of aspartate for NAA synthesis and the availability of aspartate for intercellular signaling [35].

Another plausible hypothesis is that of a cytotoxic mechanism involving NAA or its metabolic product, NAAG [36,37].

It has been proposed that the extracellular concentration of NAA increases up to 1000-fold secondary to absent or markedly decreased aspartoacylase activity, resulting in disruption of the oligodendrocyte-axon interface and interruption of myelination [38-40]. However, NAA concentration in the brain of patients with aspartoacylase deficiency is less than twofold elevated [41]. In an animal model, deficient aspartoacylase expression resulted in altered oligodendrocyte maturation, markedly reduced myelination, and increased levels of GFAP protein, a marker of gliosis [42]. In addition, NAA is taken up by astrocytes, producing the prominent vacuolization within their cytoplasm, and leading to macrocephaly. In this model, NAA is cytotoxic for astrocytes and not oligodendrocytes. The turnover of NAA is regarded as highly dynamic [43]. As such, failure of NAA degradation could provide a profound osmotic force, resulting in the observed vacuolar changes in the brain [44,45]. Disrupted osmoregulation between neurons, astrocytes, and oligodendrocytes in the presence of elevated NAA may result in increased extracellular hydrostatic pressure within myelin lamellae, leading to intramyelinic splitting [25]. The disruption of mitochondria in astrocytes may be a secondary phenomenon, but a primary alteration in mitochondrial metabolism has been proposed as potentially related to the prominent cellular alterations. N-acetylaspartate synthase deficiency corrected the myelin and neuronal phenotype in a mouse model of Canavan disease [46,47]. These findings strongly support the notion that elevation of NAA is the major disease mechanism.

A further potential pathogenetic mechanism involves NAAG as a trigger of glutamate excitotoxicity [48]. NAAG is synthesized from NAA and glutamate and is increased in patients with aspartoacylase deficiency. It has been proposed that glutamate excitotoxicity may arise both directly from elevated NAA and indirectly via the catabolism of NAAG to NAA and glutamate [25].

The possibility of an acetate deficit secondary to low or absent aspartoacylase activity (NAA is hydrolyzed to acetate and L-aspartate) also has been suggested [49,50]. This hypothesis suggests that a deficiency in NAA-derived acetate within oligodendrocytes may affect postnatal myelin lipid synthesis and may impact histone acetylation and epigenetic gene regulation [25].

Finally, oxidative stress may play a role in the pathogenesis of aspartoacylase deficiency, based upon evidence that NAA produces lipid peroxidation and protein oxidation and reduces antioxidants in rat brain [51,52]. NAA-derived acetyl groups may be a means by which oligodendrocytes preserve energetic resources during myelination by uncoupling fatty acid synthesis from oxidative metabolism [53]. Further support for this hypothesis was found in the salutary effect of anaplerotic therapy in a Canavan disease mouse model [54].

These hypotheses are not mutually exclusive, and multiple mechanisms are likely involved in the pathogenesis of aspartoacylase deficiency [25,55].

EPIDEMIOLOGY — Aspartoacylase deficiency is most prevalent among Ashkenazi Jewish individuals [29,32]. It has been described in other populations as well, including a large series from Saudi Arabia [56]. The carrier frequency among the Ashkenazi ranges from 1:37 to 1:57, yielding a range of approximate prevalence rates between 1:6000 and 1:14,000 [57,58]. The disorder is much less common in non-Ashkenazi populations [7-10]. However, sufficient data are not available to calculate a prevalence rate in groups other than Jewish people.

CLINICAL FEATURES

Neonatal/infantile (severe) form — The initial presentation of aspartoacylase deficiency, generally at approximately age three months, features lethargy and listlessness, weak cry and suck, poor head control, and hypotonia with a paucity of extremity movement [59]. However, poor feeding, irritability, and visual inattention have been described in neonates.

Macrocephaly becomes prominent by three to six months, and thereafter hypotonia progresses to spasticity, hyperreflexia, extensor plantar responses, and tonic extensor spasms. The extensor spasms may occur in response to noise, but infants with aspartoacylase deficiency do not exhibit the hyperacusis noted in Tay-Sachs disease.

By age six months, neurologic abnormalities are invariant in those with the typical form of aspartoacylase deficiency. Little subsequent development is noted, although visual fixation may be acquired later only to be lost.

Blindness in association with optic atrophy occurs between 6 and 18 months. Seizures, usually generalized tonic-clonic, are noted in approximately one-third of patients in the first year of life; the prevalence of seizures increases with age, such that nearly all patients age >10 years have seizures [60]. Unlike most leukodystrophies, the cerebrospinal fluid protein is usually normal. Fair complexion has been described, but this is of dubious significance.

Pseudobulbar signs and decerebrate posturing dominate the end stage of aspartoacylase deficiency. Feeding is a major issue with prominent swallowing dysfunction and gastroesophageal reflux. Death may occur in childhood, although survival into the teens is common [60], as the result of enhanced medical management, particularly alternative feeding strategies.

Juvenile (mild) form — The presence of a potential juvenile-onset form, beginning after age five and characterized by spasticity, cerebellar dysfunction, and blindness, was first described prior to the availability of biochemical and molecular diagnosis [61]. At that time, the diagnosis was based upon the characteristic clinical features and demonstration of spongiform degeneration of the brain. However, this is a nonspecific pathology shared by a number of other unrelated conditions [62], and it was unclear whether this represented a true case of aspartoacylase deficiency.

The disease typically begins in infancy but progresses at a highly variable rate, with no correlation of genotype or residual enzyme activity to clinical presentation [63], further expanding the conversation regarding juvenile-onset forms of the condition.

One study reported 22 Ashkenazi infants with Canavan disease and typical onset in infancy [62]. There were 14 patients younger than age 6 who were clinically stable. Two patients died before the age of 5. Survival beyond age 5 was noted in six patients, with ages at death ranging from 6 to 17 years.

Another series of 60 children, mainly Ashkenazi, with onset prior to 10 months of life demonstrated high variability in survival, including significantly different longevity in two siblings from two families [59].

The advent of readily available molecular testing for aspartoacylase deficiency has helped to clarify this question. In subsequent reports, milder disease expression has been associated with specific ASPA gene mutations that may produce a relatively benign phenotype with variable features, including mild developmental delay without regression, hypotonia, normocephaly, tremor, and mild coordination difficulties [20,64-69].

Neuroimaging — In patients with aspartoacylase deficiency, cranial imaging by computed tomography (CT) and magnetic resonance imaging (MRI) typically shows diffuse and symmetrical white matter involvement. CT shows marked reduction of white matter [70-74]. MRI demonstrates white matter that is hypointense on T1 and hyperintense on T2 (image 1 and image 2) [71-75]. The white matter may show multiple round or oval cystic changes with a honeycomb or cribriform appearance that may represent dilatation of Virchow-Robin spaces related to spongiform degeneration [75,76]. On diffusion-weighted MRI sequences, restricted diffusion in deep white matter and brainstem corresponding to cytotoxic brain edema is frequently present early in Canavan disease [77]. In individuals with delayed onset and slower progression, brain MRI may show less prominent white matter changes and increased signal intensity in basal ganglia [63,78,79]. In a case report of a girl with incidentally discovered Canavan disease who had a benign clinical course, there were atypical MRI findings of diffuse T2 and fluid-attenuated inversion recovery (FLAIR) signal hyperintensity and restricted diffusion involving the cortex and sparing the white matter [67].

Magnetic resonance spectroscopy (MRS) typically shows markedly increased levels of N-acetylaspartic acid (NAA) (image 3) [73,74,80,81], regardless of echo time utilized [82]. MRI without any white matter abnormalities but with elevated NAA peak on MRS was reported in a five-year-old child with the insidious onset of mild motor and speech delay [83].

MRS is also being studied to track the natural progression of metabolite changes and could serve as a measure for monitoring therapeutic interventions [84]. In some instances, NAA levels are normal, but other metabolites are reduced, yielding elevated ratios of NAA/choline and NAA/creatine [85].

Neuropathology — Brain weight is increased significantly in aspartoacylase deficiency, reflecting the prominent macrocephaly noted in early infancy [61]. However, by age 30 months, brain weight may be normal, reflecting the progressive white matter loss [61].

The gross pathologic findings are dominated by spongy degeneration of deep cortex, subcortical white matter, and cerebellum. The spongiform changes reflect vacuolated astrocytes in deeper cortical layers and in adjacent subcortical white matter. Grossly enlarged Alzheimer type II astrocytes are seen in gray matter.

Ultrastructural studies reveal disruption of mitochondria in astrocytes. Myelin is markedly reduced and may be virtually absent in some areas. Myelin lamellae within subcortical white matter are separated by vacuoles. With prolonged survival, white matter is depleted and vacuolization is present throughout the cortex.

Unlike galactosylceramide lipidosis (Krabbe disease) and sulfatide lipidosis (metachromatic leukodystrophy), peripheral nerves are generally uninvolved in aspartoacylase deficiency [61].

Pregnancy — As affected individuals are unlikely to reproduce, no information is available regarding pregnancy in patients with aspartoacylase deficiency.

Findings from the knockout mouse model suggest reduced reproductive health in the homozygous affected females and heightened fetal lethality (50 percent fewer pups/litter) [86].

DIAGNOSIS — The clinical diagnosis of aspartoacylase deficiency is relatively straightforward. In symptomatic infants with compatible clinical features (eg, hypotonia, poor head control, macrocephaly) and neuroimaging findings, the diagnosis is supported by elevated levels of urine N-acetylaspartic acid (NAA) and/or biallelic pathogenic variants in ASPA identified by molecular genetic testing [29].

Laboratory studies — In patients with clinical features and cranial imaging findings suggesting aspartoacylase deficiency, one should request measurement of NAA in urine using gas chromatography/mass spectrometry [87].

Urine levels of NAA are increased up to 200 times normal in aspartoacylase deficiency [28]. This measurement is determined by gas chromatography/mass spectrometry. Occasional patients with aspartoacylase deficiency have lower levels of urine NAA excretion, but the levels are still approximately five- to ten-fold higher than normal [88].

In affected infants, levels of NAA are also increased in plasma and in cerebrospinal fluid, but elevated urine NAA is sufficient to support the diagnosis [29].

Canavan disease may be suspected when the NAA peak is elevated in MR spectroscopy [80,81,83-85]. (See 'Neuroimaging' above.)

Aspartoacylase activity can be measured reliably in cultured skin fibroblasts [2,89,90]. This test is used to confirm the chemical diagnosis and to rule out false positives, particularly for patients who have an elevation of urine NAA that is lower than usual for aspartoacylase deficiency.

Molecular genetic testing — For patients diagnosed with aspartoacylase deficiency by elevated urine NAA levels, targeted molecular genetic testing of the patient and family members can be useful for purposes of genetic counseling, particularly in Ashkenazi children where the number of possible mutations is small, or in families with a previously affected sibling whose mutation is known. In non-Ashkenazi families with their first affected child and no previously identified mutation, complete molecular testing may be required to provide genetic counseling, especially when the family may wish to consider a future pregnancy. Genetic testing is also useful for molecular confirmation of aspartoacylase deficiency, which may be needed for participation in treatment trials.

Targeted mutation analysis can identify specific alleles that cause most cases of aspartoacylase deficiency [29]:

In Ashkenazi Jewish cases, two pathogenic variants, p.Glu285Ala and p.Tyr231X, are found in approximately 98 percent of disease-causing alleles [12]. Two other pathogenic variants, p.Ala305Glu and c.433-2A>G, each account for approximately one percent of alleles [8].

In non-Jewish patients of European origin, the p.Ala305Glu pathogenic variant accounts for 40 to 60 percent of disease-causing alleles [8,10].

Sequence analysis of the ASPA coding region can be performed first for individuals of non-Ashkenazi Jewish ancestry [29]. Complete or partial deletions of the entire ASPA gene in aspartoacylase deficiency are rare [14,91].

Prenatal diagnosis — Prenatal diagnosis is possible once pathogenic variants of ASPA have been identified in an affected member of the family [29].

For pregnancies at 25 percent risk (ie, the parents are each carriers of a pathogenic variant in the ASPA gene, or the pathogenic variant is known from a previously affected offspring), mutation analysis can be performed from cells obtained using chorionic villous sampling at approximately 10 to 12 weeks gestation or amniocentesis at approximately 15 to 38 weeks gestation [92,93]. Analysis of cell-free DNA in the mother's circulation is likely to become more available in the coming years [94].

Preimplantation diagnosis using single cell molecular methodologies has been accomplished successfully in one of two families evaluated [95].

DIFFERENTIAL DIAGNOSIS — The differential diagnosis of aspartoacylase deficiency includes other progressive white matter diseases of infancy, particularly the following:

Krabbe disease (galactosylceramide lipidosis) (see "Krabbe disease")

Metachromatic leukodystrophy (sulfatide lipidosis) (see "Metachromatic leukodystrophy")

Vanishing white matter disease (see "Vanishing white matter disease")

Alexander disease which, like Canavan disease, is associated with macrocephaly (see "Alexander disease")

Spongiform degeneration of the brain is a nonspecific pathology shared by several other unrelated conditions [62]. These include certain mitochondrial disorders (eg, Leigh syndrome), metabolic diseases (eg, glycine encephalopathy), and viral infections [29].

The subsequent clinical course, cranial imaging abnormalities, and biochemical studies should differentiate aspartoacylase deficiency from the others.

MANAGEMENT — No effective disease-modifying therapy is available for aspartoacylase deficiency.

Supportive care — Management of aspartoacylase deficiency is supportive and aimed at maintaining nutrition and hydration, protecting the airway, preventing seizures, minimizing contractures, and treating infections [29]. Assessment of nutritional and developmental status is recommended to guide management.

A feeding gastrostomy tube is typically needed to maintain adequate nutrition and hydration in the presence of dysphagia. A gastrostomy tube can also reduce the risk of aspiration. (See "Poor weight gain in children younger than two years in resource-abundant settings: Management", section on 'Estimation of energy requirements'.)

Seizures are treated with standard antiseizure medications. (See "Seizures and epilepsy in children: Initial treatment and monitoring".)

Exercise (physical therapy) and position changes are helpful to reduce the risk of contractures and decubitus ulcers, and to improve sitting posture.

Special education programs and interventions to enhance communication skills may be helpful, particularly for those with a less severe clinical course.

Botulinum toxin injections can be used to treat spasticity.

Investigational therapies — Gene transfer, enzyme replacement therapy, antisense oligonucleotide therapy, aspartate N-acetyltransferase (ANAT) inhibition, modification of SLC13A3, acetate supplementation, anaplerotic therapy, and lithium are being studied as possible treatments for aspartoacylase deficiency.

Several preliminary studies have evaluated gene transfer:

The combined delivery of liposome enclosed enzyme and viral vector-mediated gene transduction was employed in two children [96,97]. While successful delivery could be demonstrated, the results did not demonstrate substantial efficacy.

Adeno-associated viral vector (AAV) mediated gene transfer of ASPA was studied in 10 children with aspartoacylase deficiency [98,99]. No significant adverse immune response was noted, and no statement on efficacy was available.

A knock-out mouse model has been developed, paving the way for additional gene therapy studies [86,100-105], and a spontaneously arising rat model with aspartoacylase deficiency also has been developed [106]. Effective gene transfer using an adeno-associated viral vector has been demonstrated in each model [107-109].

A novel AAV variant, AAV/Olig001, engineered with preferential tropism for the oligodendrocyte cell surface, demonstrated widespread oligodendrocyte transduction in adult Canavan mice (except for the cerebellum) after intracerebroventricular administration to the CSF. This treatment resulted in an improvement in aspartoacylase activity, motor function, and vacuolation [110].

Dual function AAV gene therapy utilizing an AAV vector designed to deliver ASPA and to inhibit ANAT activity was injected bilaterally directly into the striatum, thalamus, and cerebellum of adult Canavan mice, resulting in improvement in clinical parameters and reversal of pathologic findings [111].

AAV-ASPA gene therapy was performed in a two-year-old boy with Canavan disease using simultaneous intravenous and intracerebroventricular administration and preventive immunomodulation. This treatment resulted in increased brain myelination, improved vision, and improved motor function at more than two years posttreatment [112].

Neural progenitor cells have been transplanted successfully in the mouse knockout model and have differentiated into myelin-forming oligodendrocytes [113]. Cell-based therapy utilizing human-induced pluripotent stem cell (iPSC)-derived neural progenitor cells (NPCs) and oligodendrocyte progenitor cells (OPCs) may provide an alternative to AAV-mediated gene therapy [114,115].

Modified aspartoacylase has been shown to reach the central nervous system following intraperitoneal injection, resulting in a reduction of N-acetylaspartic acid (NAA) levels [116]. This finding suggests that enzyme replacement approaches may be examined in this model.

Intracisternal administration of locked nucleic acid antisense oligonucleotide (LNA ASO) to knock down the expression of neuronal NAA synthesizing ANAT in adult Canavan mice resulted in a transient decrease in NAA and improvement in ataxia, but these effects were not sustained, suggesting that repeated doses may be necessary [117].

Decreasing NAA by lowering its synthesis may be a useful approach [46,47]. Bisubstrate analog inhibitors of ANAT have been developed [118], and drug discovery screening techniques have identified potential ANAT inhibitors with drug-like chemical properties [119,120]. The use of ANAT inhibitors alone and in combination with other treatments may be complementary.

In Canavan disease mice, knockout of the high-affinity sodium-dependent dicarboxylate cotransporter 3 (NaDC3), which is encoded by SLC13A3, resulted in decreased brain NAA accumulation, improved motor function, and decreased brain vacuolation [121]. However, SLC13A3 pathogenic variants have been linked to fever-induced acute reversible leukoencephalopathy with increased urinary excretion of alpha-ketoglutarate (OMIM#618384) in a small number of patients [122], which suggests that caution is indicated when considering the development of SLC13A3 inhibitors for clinical use.

Acetate supplementation has been proposed as potentially beneficial for myelin formation by oligodendrocytes based upon the presumption that aspartoacylase deficiency and resultant acetate deficiency in these cells could be responsible for abnormal myelination [49]. The acetate precursor glyceryl triacetate was evaluated in low doses (up to 25 mg/kg daily) in two children with aspartoacylase deficiency and in higher doses (up to 5.8 mg/kg daily) in an animal model [123]. No evidence of toxicity was noted with either dosing schedule. In addition, the two children demonstrated no further clinical deterioration, providing a rationale for evaluating a larger dose in affected children.

Anaplerotic therapy using triheptanoin shows promise based upon a study in an animal model [54]. (See 'Pathogenesis' above.)

Intraperitoneal lithium chloride injection produced reduction, albeit modest, of brain NAA levels in the spontaneous rat model of aspartoacylase deficiency [124]. In a study of six children with Canavan disease, lithium citrate administration was associated with a small but statistically significant decline in brain NAA detected by magnetic resonance spectroscopy [125].

SUMMARY

Description – Aspartoacylase deficiency (Canavan disease; MIM #271900) is an autosomal recessive spongiform leukodystrophy that is prevalent in the Ashkenazi Jewish population. The disease typically begins in infancy and is marked by relentless progression. (See 'Introduction and background' above.)

Etiology – Aspartoacylase deficiency is caused by pathogenic variants in the ASPA gene that encodes the enzyme aspartoacylase. The resulting deficiency of aspartoacylase leads to accumulation of N-acetylaspartic acid (NAA) in brain and to oligodendrocyte dysfunction, spongiform changes, and absence of myelin. However, the precise mechanisms causing spongiform degeneration are uncertain. (See 'Etiology' above.)

Epidemiology – Aspartoacylase deficiency is highly prevalent among Ashkenazi Jewish individuals, but is much less common in other populations. (See 'Epidemiology' above.)

Neonatal/infantile form – Aspartoacylase deficiency typically presents at approximately age three months with lethargy and listlessness, weak cry and suck, poor head control, and hypotonia with a paucity of extremity movement. Macrocephaly becomes prominent by three to six months. Thereafter hypotonia progresses to spasticity and tonic extensor spasms. By age six months, neurologic abnormalities are invariant. Little subsequent development is noted. Blindness from optic atrophy occurs between 6 and 18 months. Seizures are noted in approximately one-third of patients in the first year of life, and seizure prevalence increases with age. Pseudobulbar signs and decerebrate posturing dominate the end stage. (See 'Neonatal/infantile (severe) form' above.)

Juvenile form – Juvenile forms of aspartoacylase deficiency are rare but may be underdiagnosed. Measurement of NAA in urine or by magnetic resonance spectroscopy (MRS) should be included as part of the evaluation of leukoencephalopathy at all ages. (See 'Juvenile (mild) form' above.)

Imaging – Brain imaging by computed tomography (CT) and magnetic resonance imaging (MRI) reveals diffuse and symmetrical white matter involvement. MRS reveals markedly increased levels of NAA. (See 'Neuroimaging' above.)

Pathology – The gross pathologic findings are dominated by spongy degeneration of deep cortex, subcortical white matter, and cerebellum. The spongiform changes reflect vacuolated astrocytes in deeper cortical layers and in adjacent subcortical white matter. (See 'Neuropathology' above.)

Diagnosis – In symptomatic infants with compatible clinical features (eg, hypotonia, poor head control, macrocephaly) and neuroimaging findings, the diagnosis of aspartoacylase deficiency is supported by elevated levels of urine NAA and/or biallelic pathogenic variants in ASPA identified by molecular genetic testing. (See 'Diagnosis' above.)

Differential diagnosis – The differential diagnosis of aspartoacylase deficiency includes other progressive white matter diseases of infancy, particularly Krabbe disease, metachromatic leukodystrophy, early-onset adrenoleukodystrophy, Alexander disease, and demyelinating disorders. (See 'Differential diagnosis' above.)

Management – No effective disease-modifying therapy is available for aspartoacylase deficiency. Management is supportive and aimed at maintaining nutrition and hydration, protecting the airway, preventing seizures, minimizing contractures, and treating infections. Investigational therapies include gene transfer, enzyme replacement therapy, antisense oligonucleotide therapy, reduction in NAA synthesis by ANAT inhibition, modification of SLC13A3, acetate supplementation, anaplerotic therapy, and lithium. (See 'Management' above.)

ACKNOWLEDGMENTS — The editorial staff at UpToDate acknowledge Alan K Percy, MD, and Raphael Schiffman, MD, MHSc, FAAN, who contributed to earlier versions of this topic review.

  1. Hagenfeldt L, Bollgren I, Venizelos N. N-acetylaspartic aciduria due to aspartoacylase deficiency--a new aetiology of childhood leukodystrophy. J Inherit Metab Dis 1987; 10:135.
  2. Matalon R, Michals K, Sebesta D, et al. Aspartoacylase deficiency and N-acetylaspartic aciduria in patients with Canavan disease. Am J Med Genet 1988; 29:463.
  3. Kaul R, Gao GP, Balamurugan K, Matalon R. Cloning of the human aspartoacylase cDNA and a common missense mutation in Canavan disease. Nat Genet 1993; 5:118.
  4. Kaul R, Balamurugan K, Gao GP, Matalon R. Canavan disease: genomic organization and localization of human ASPA to 17p13-ter and conservation of the ASPA gene during evolution. Genomics 1994; 21:364.
  5. Kaul R, Gao GP, Balamurugan K, Matalon R. Canavan disease: molecular basis of aspartoacylase deficiency. J Inherit Metab Dis 1994; 17:295.
  6. Matalon R, Michals-Matalon K. Biochemistry and molecular biology of Canavan disease. Neurochem Res 1999; 24:507.
  7. Elpeleg ON, Anikster Y, Barash V, et al. The frequency of the C854 mutation in the aspartoacylase gene in Ashkenazi Jews in Israel. Am J Hum Genet 1994; 55:287.
  8. Kaul R, Gao GP, Aloya M, et al. Canavan disease: mutations among Jewish and non-Jewish patients. Am J Hum Genet 1994; 55:34.
  9. Shaag A, Anikster Y, Christensen E, et al. The molecular basis of canavan (aspartoacylase deficiency) disease in European non-Jewish patients. Am J Hum Genet 1995; 57:572.
  10. Elpeleg ON, Shaag A. The spectrum of mutations of the aspartoacylase gene in Canavan disease in non-Jewish patients. J Inherit Metab Dis 1999; 22:531.
  11. Kaul R, Gao GP, Matalon R, et al. Identification and expression of eight novel mutations among non-Jewish patients with Canavan disease. Am J Hum Genet 1996; 59:95.
  12. Sistermans EA, de Coo RF, van Beerendonk HM, et al. Mutation detection in the aspartoacylase gene in 17 patients with Canavan disease: four new mutations in the non-Jewish population. Eur J Hum Genet 2000; 8:557.
  13. Zeng BJ, Wang ZH, Ribeiro LA, et al. Identification and characterization of novel mutations of the aspartoacylase gene in non-Jewish patients with Canavan disease. J Inherit Metab Dis 2002; 25:557.
  14. Zeng BJ, Wang ZH, Torres PA, et al. Rapid detection of three large novel deletions of the aspartoacylase gene in non-Jewish patients with Canavan disease. Mol Genet Metab 2006; 89:156.
  15. Zeng BJ, Pastores GM, Leone P, et al. Mutation analysis of the aspartoacylase gene in non-Jewish patients with Canavan disease. Adv Exp Med Biol 2006; 576:165.
  16. Kobayashi K, Tsujino S, Ezoe T, et al. Missense mutation (I143T) in a Japanese patient with Canavan disease. Hum Mutat 1998; Suppl 1:S308.
  17. Olsen TR, Tranebjaerg L, Kvittingen EA, et al. Two novel aspartoacylase gene (ASPA) missense mutations specific to Norwegian and Swedish patients with Canavan disease. J Med Genet 2002; 39:e55.
  18. Bijarnia S, Kohli S, Puri RD, et al. Molecular characterisation and prenatal diagnosis of Asparto-acylase deficiency (Canavan disease)--report of two novel and two known mutations from the Indian subcontinent. Indian J Pediatr 2013; 80:26.
  19. Gowda VK, Bharathi NK, Bettaiah J, et al. Canavan Disease: Clinical and Laboratory Profile from Southern Part of India. Ann Indian Acad Neurol 2021; 24:347.
  20. Kotambail A, Selvam P, Muthusamy K, et al. Clustering of Juvenile Canavan disease in an Indian community due to population bottleneck and isolation: genomic signatures of a founder event. Eur J Hum Genet 2023; 31:73.
  21. Moffett JR, Arun P, Ariyannur PS, Namboodiri AM. N-Acetylaspartate reductions in brain injury: impact on post-injury neuroenergetics, lipid synthesis, and protein acetylation. Front Neuroenergetics 2013; 5:11.
  22. Baslow MH. N-acetylaspartate in the vertebrate brain: metabolism and function. Neurochem Res 2003; 28:941.
  23. Baslow MH, Suckow RF, Sapirstein V, Hungund BL. Expression of aspartoacylase activity in cultured rat macroglial cells is limited to oligodendrocytes. J Mol Neurosci 1999; 13:47.
  24. Kirmani BF, Jacobowitz DM, Kallarakal AT, Namboodiri MA. Aspartoacylase is restricted primarily to myelin synthesizing cells in the CNS: therapeutic implications for Canavan disease. Brain Res Mol Brain Res 2002; 107:176.
  25. Wei H, Moffett JR, Amanat M, et al. The pathogenesis of, and pharmacological treatment for, Canavan disease. Drug Discov Today 2022; 27:2467.
  26. Chakraborty G, Mekala P, Yahya D, et al. Intraneuronal N-acetylaspartate supplies acetyl groups for myelin lipid synthesis: evidence for myelin-associated aspartoacylase. J Neurochem 2001; 78:736.
  27. Kirmani BF, Jacobowitz DM, Namboodiri MA. Developmental increase of aspartoacylase in oligodendrocytes parallels CNS myelination. Brain Res Dev Brain Res 2003; 140:105.
  28. Burlina AP, Corazza A, Ferrari V, et al. Detection of increased urinary N-acetylaspartylglutamate in Canavan disease. Eur J Pediatr 1994; 153:538.
  29. Matalon R, Delgado L, Michals-Matalon K. Canavan Disease. In: GeneReviews, Adam MP, Ardinger HH, Pagon RA, et al. (Eds), University of Washington, Seattle (WA) 2019. https://www.ncbi.nlm.nih.gov/books/NBK1234/ (Accessed on July 29, 2019).
  30. Hershfield JR, Pattabiraman N, Madhavarao CN, Namboodiri MA. Mutational analysis of aspartoacylase: implications for Canavan disease. Brain Res 2007; 1148:1.
  31. Zano S, Wijayasinghe YS, Malik R, et al. Relationship between enzyme properties and disease progression in Canavan disease. J Inherit Metab Dis 2013; 36:1.
  32. Kumar S, Mattan NS, de Vellis J. Canavan disease: a white matter disorder. Ment Retard Dev Disabil Res Rev 2006; 12:157.
  33. Baslow MH. Canavan's spongiform leukodystrophy: a clinical anatomy of a genetic metabolic CNS disease. J Mol Neurosci 2000; 15:61.
  34. Moffett JR, Ross B, Arun P, et al. N-Acetylaspartate in the CNS: from neurodiagnostics to neurobiology. Prog Neurobiol 2007; 81:89.
  35. Baslow MH, Resnik TR. Canavan disease. Analysis of the nature of the metabolic lesions responsible for development of the observed clinical symptoms. J Mol Neurosci 1997; 9:109.
  36. Kolodziejczyk K, Hamilton NB, Wade A, et al. The effect of N-acetyl-aspartyl-glutamate and N-acetyl-aspartate on white matter oligodendrocytes. Brain 2009; 132:1496.
  37. Kumar S, Biancotti JC, Matalon R, de Vellis J. Lack of aspartoacylase activity disrupts survival and differentiation of neural progenitors and oligodendrocytes in a mouse model of Canavan disease. J Neurosci Res 2009; 87:3415.
  38. Namboodiri AM, Peethambaran A, Mathew R, et al. Canavan disease and the role of N-acetylaspartate in myelin synthesis. Mol Cell Endocrinol 2006; 252:216.
  39. Namboodiri AM, Moffett JR, Arun P, et al. Defective myelin lipid synthesis as a pathogenic mechanism of Canavan disease. Adv Exp Med Biol 2006; 576:145.
  40. Kumar S, Sowmyalakshmi R, Daniels SL, et al. Does ASPA gene mutation in Canavan disease alter oligodendrocyte development? A tissue culture study of ASPA KO mice brain. Adv Exp Med Biol 2006; 576:175.
  41. Blüml S. In vivo quantitation of cerebral metabolite concentrations using natural abundance 13C MRS at 1.5 T. J Magn Reson 1999; 136:219.
  42. Mattan NS, Ghiani CA, Lloyd M, et al. Aspartoacylase deficiency affects early postnatal development of oligodendrocytes and myelination. Neurobiol Dis 2010; 40:432.
  43. Baslow MH. Brain N-acetylaspartate as a molecular water pump and its role in the etiology of Canavan disease: a mechanistic explanation. J Mol Neurosci 2003; 21:185.
  44. Baslow MH, Guilfoyle DN. Are astrocytes the missing link between lack of brain aspartoacylase activity and the spongiform leukodystrophy in Canavan disease? Neurochem Res 2009; 34:1523.
  45. Baslow MH, Guilfoyle DN. Canavan disease, a rare early-onset human spongiform leukodystrophy: insights into its genesis and possible clinical interventions. Biochimie 2013; 95:946.
  46. Maier H, Wang-Eckhardt L, Hartmann D, et al. N-Acetylaspartate Synthase Deficiency Corrects the Myelin Phenotype in a Canavan Disease Mouse Model But Does Not Affect Survival Time. J Neurosci 2015; 35:14501.
  47. Sohn J, Bannerman P, Guo F, et al. Suppressing N-Acetyl-l-Aspartate Synthesis Prevents Loss of Neurons in a Murine Model of Canavan Leukodystrophy. J Neurosci 2017; 37:413.
  48. Burlina AP, Ferrari V, Divry P, et al. N-acetylaspartylglutamate in Canavan disease: an adverse effector? Eur J Pediatr 1999; 158:406.
  49. Mathew R, Arun P, Madhavarao CN, et al. Progress toward acetate supplementation therapy for Canavan disease: glyceryl triacetate administration increases acetate, but not N-acetylaspartate, levels in brain. J Pharmacol Exp Ther 2005; 315:297.
  50. Madhavarao CN, Arun P, Moffett JR, et al. Defective N-acetylaspartate catabolism reduces brain acetate levels and myelin lipid synthesis in Canavan's disease. Proc Natl Acad Sci U S A 2005; 102:5221.
  51. Pederzolli CD, Mescka CP, Scapin F, et al. N-acetylaspartic acid promotes oxidative stress in cerebral cortex of rats. Int J Dev Neurosci 2007; 25:317.
  52. Pederzolli CD, Rockenbach FJ, Zanin FR, et al. Intracerebroventricular administration of N-acetylaspartic acid impairs antioxidant defenses and promotes protein oxidation in cerebral cortex of rats. Metab Brain Dis 2009; 24:283.
  53. Francis JS, Strande L, Markov V, Leone P. Aspartoacylase supports oxidative energy metabolism during myelination. J Cereb Blood Flow Metab 2012; 32:1725.
  54. Francis JS, Markov V, Leone P. Dietary triheptanoin rescues oligodendrocyte loss, dysmyelination and motor function in the nur7 mouse model of Canavan disease. J Inherit Metab Dis 2014; 37:369.
  55. Pleasure D, Guo F, Chechneva O, et al. Pathophysiology and Treatment of Canavan Disease. Neurochem Res 2020; 45:561.
  56. Gascon GG, Ozand PT, Mahdi A, et al. Infantile CNS spongy degeneration--14 cases: clinical update. Neurology 1990; 40:1876.
  57. Kronn D, Oddoux C, Phillips J, Ostrer H. Prevalence of Canavan disease heterozygotes in the New York metropolitan Ashkenazi Jewish population. Am J Hum Genet 1995; 57:1250.
  58. Feigenbaum A, Moore R, Clarke J, et al. Canavan disease: carrier-frequency determination in the Ashkenazi Jewish population and development of a novel molecular diagnostic assay. Am J Med Genet A 2004; 124A:142.
  59. Traeger EC, Rapin I. The clinical course of Canavan disease. Pediatr Neurol 1998; 18:207.
  60. Bley A, Denecke J, Kohlschütter A, et al. The natural history of Canavan disease: 23 new cases and comparison with patients from literature. Orphanet J Rare Dis 2021; 16:227.
  61. Adachi M, Schneck L, Cara J, Volk BW. Spongy degeneration of the central nervous system (van Bogaert and Bertrand type; Canavan's disease). A review. Hum Pathol 1973; 4:331.
  62. Zelnik N, Luder AS, Elpeleg ON, et al. Protracted clinical course for patients with Canavan disease. Dev Med Child Neurol 1993; 35:355.
  63. Sarret C, Boespflug-Tanguy O, Rodriguez D. Atypical clinical and radiological course of a patient with Canavan disease. Metab Brain Dis 2016; 31:475.
  64. Zafeiriou DI, Kleijer WJ, Maroupoulos G, et al. Protracted course of N-acetylaspartic aciduria in two non-Jewish siblings: identical clinical and magnetic resonance imaging findings. Brain Dev 1999; 21:205.
  65. Tacke U, Olbrich H, Sass JO, et al. Possible genotype-phenotype correlations in children with mild clinical course of Canavan disease. Neuropediatrics 2005; 36:252.
  66. Janson CG, Kolodny EH, Zeng BJ, et al. Mild-onset presentation of Canavan's disease associated with novel G212A point mutation in aspartoacylase gene. Ann Neurol 2006; 59:428.
  67. Nguyen HV, Ishak GE. Canavan disease - unusual imaging features in a child with mild clinical presentation. Pediatr Radiol 2015; 45:457.
  68. Velinov M, Zellers N, Styles J, Wisniewski K. Homozygosity for mutation G212A of the gene for aspartoacylase is associated with atypical form of Canavan's disease. Clin Genet 2008; 73:288.
  69. Çakar NE, Aksu Uzunhan T. A case of juvenile Canavan disease with distinct pons involvement. Brain Dev 2020; 42:222.
  70. Rushton AR, Shaywitz BA, Duncan CC, et al. Computed tomography in the diagnosis of Canavan's disease. Ann Neurol 1981; 10:57.
  71. McAdams HP, Geyer CA, Done SL, et al. CT and MR imaging of Canavan disease. AJNR Am J Neuroradiol 1990; 11:397.
  72. Brismar J, Brismar G, Gascon G, Ozand P. Canavan disease: CT and MR imaging of the brain. AJNR Am J Neuroradiol 1990; 11:805.
  73. Marks HG, Caro PA, Wang ZY, et al. Use of computed tomography, magnetic resonance imaging, and localized 1H magnetic resonance spectroscopy in Canavan's disease: a case report. Ann Neurol 1991; 30:106.
  74. Michel SJ, Given CA 2nd. Case 99: Canavan disease. Radiology 2006; 241:310.
  75. Pradhan S, Goyal G. Teaching NeuroImages: honeycomb appearance of the brain in a patient with Canavan disease. Neurology 2011; 76:e68.
  76. Bhat MD, Manjunath N, Kumari R, et al. Cribriform Appearance of White Matter in Canavan Disease Associated with Novel Mutations of ASPA Gene. J Pediatr Genet 2022; 11:267.
  77. Merrill ST, Nelson GR, Longo N, Bonkowsky JL. Cytotoxic edema and diffusion restriction as an early pathoradiologic marker in canavan disease: case report and review of the literature. Orphanet J Rare Dis 2016; 11:169.
  78. Toft PB, Geiss-Holtorff R, Rolland MO, et al. Magnetic resonance imaging in juvenile Canavan disease. Eur J Pediatr 1993; 152:750.
  79. Yalcinkaya C, Benbir G, Salomons GS, et al. Atypical MRI findings in Canavan disease: a patient with a mild course. Neuropediatrics 2005; 36:336.
  80. Wittsack HJ, Kugel H, Roth B, Heindel W. Quantitative measurements with localized 1H MR spectroscopy in children with Canavan's disease. J Magn Reson Imaging 1996; 6:889.
  81. Galanaud D, Nicoli F, Confort-Gouny S, et al. [Brain magnetic resonance spectroscopy]. J Radiol 2007; 88:483.
  82. Cecil KM. Use of MRS in inborn errors of metabolism: Canavan's disease and MRS in differential diagnosis. In: Magnetic Resonance Spectroscopy: Tools for Neuroscience Research and Emerging Clinical Applications, 1st ed, Stagg C, Rothman DL (Eds), Academic Press, 2013. p.196.
  83. Jauhari P, Saini L, Chakrabarty B, et al. Juvenile Canavan Disease: A Leukodystrophy without White Matter Changes. Neuropediatrics 2018; 49:420.
  84. Janson CG, McPhee SW, Francis J, et al. Natural history of Canavan disease revealed by proton magnetic resonance spectroscopy (1H-MRS) and diffusion-weighted MRI. Neuropediatrics 2006; 37:209.
  85. Barker PB, Bryan RN, Kumar AJ, Naidu S. Proton NMR spectroscopy of Canavan's disease. Neuropediatrics 1992; 23:263.
  86. Surendran S, Ezell EL, Quast MJ, et al. Aspartoacylase deficiency does not affect N-acetylaspartylglutamate level or glutamate carboxypeptidase II activity in the knockout mouse brain. Brain Res 2004; 1016:268.
  87. Inoue Y, Kuhara T. Rapid and sensitive screening for and chemical diagnosis of Canavan disease by gas chromatography-mass spectrometry. J Chromatogr B Analyt Technol Biomed Life Sci 2004; 806:33.
  88. Matalon R, Kaul R, Michals K. Canavan disease: biochemical and molecular studies. J Inherit Metab Dis 1993; 16:744.
  89. Matalon R, Kaul R, Casanova J, et al. SSIEM Award. Aspartoacylase deficiency: the enzyme defect in Canavan disease. J Inherit Metab Dis 1989; 12 Suppl 2:329.
  90. Ozand PT, Gascon GG, Dhalla M. Aspartoacylase deficiency and Canavan disease in Saudi Arabia. Am J Med Genet 1990; 35:266.
  91. Kaya N, Imtiaz F, Colak D, et al. Genome-wide gene expression profiling and mutation analysis of Saudi patients with Canavan disease. Genet Med 2008; 10:675.
  92. Matalon R, Kaul R, Gao GP, et al. Prenatal diagnosis for Canavan disease: the use of DNA markers. J Inherit Metab Dis 1995; 18:215.
  93. Elpeleg ON, Shaag A, Anikster Y, Jakobs C. Prenatal detection of Canavan disease (aspartoacylase deficiency) by DNA analysis. J Inherit Metab Dis 1994; 17:664.
  94. Carlson LM, Vora NL. Prenatal Diagnosis: Screening and Diagnostic Tools. Obstet Gynecol Clin North Am 2017; 44:245.
  95. Yaron Y, Schwartz T, Mey-Raz N, et al. Preimplantation genetic diagnosis of Canavan disease. Fetal Diagn Ther 2005; 20:465.
  96. Leone P, Janson CG, McPhee SJ, During MJ. Global CNS gene transfer for a childhood neurogenetic enzyme deficiency: Canavan disease. Curr Opin Mol Ther 1999; 1:487.
  97. Leone P, Janson CG, Bilaniuk L, et al. Aspartoacylase gene transfer to the mammalian central nervous system with therapeutic implications for Canavan disease. Ann Neurol 2000; 48:27.
  98. Janson C, McPhee S, Bilaniuk L, et al. Clinical protocol. Gene therapy of Canavan disease: AAV-2 vector for neurosurgical delivery of aspartoacylase gene (ASPA) to the human brain. Hum Gene Ther 2002; 13:1391.
  99. McPhee SW, Janson CG, Li C, et al. Immune responses to AAV in a phase I study for Canavan disease. J Gene Med 2006; 8:577.
  100. Matalon R, Rady PL, Platt KA, et al. Knock-out mouse for Canavan disease: a model for gene transfer to the central nervous system. J Gene Med 2000; 2:165.
  101. Surendran S, Rady PL, Michals-Matalon K, et al. Expression of glutamate transporter, GABRA6, serine proteinase inhibitor 2 and low levels of glutamate and GABA in the brain of knock-out mouse for Canavan disease. Brain Res Bull 2003; 61:427.
  102. Surendran S, Matalon KM, Szucs S, et al. Metabolic changes in the knockout mouse for Canavan's disease: implications for patients with Canavan's disease. J Child Neurol 2003; 18:611.
  103. Surendran S, Ezell EL, Quast MJ, et al. Mental retardation and hypotonia seen in the knock out mouse for Canavan disease is not due to succinate semialdehyde dehydrogenase deficiency. Neurosci Lett 2004; 358:29.
  104. Surendran S, Szucs S, Tyring SK, Matalon R. Aspartoacylase gene knockout in the mouse: impact on reproduction. Reprod Toxicol 2005; 20:281.
  105. Matalon R, Michals-Matalon K, Surendran S, Tyring SK. Canavan disease: studies on the knockout mouse. Adv Exp Med Biol 2006; 576:77.
  106. Kitada K, Akimitsu T, Shigematsu Y, et al. Accumulation of N-acetyl-L-aspartate in the brain of the tremor rat, a mutant exhibiting absence-like seizure and spongiform degeneration in the central nervous system. J Neurochem 2000; 74:2512.
  107. Matalon R, Surendran S, Rady PL, et al. Adeno-associated virus-mediated aspartoacylase gene transfer to the brain of knockout mouse for canavan disease. Mol Ther 2003; 7:580.
  108. McPhee SW, Francis J, Janson CG, et al. Effects of AAV-2-mediated aspartoacylase gene transfer in the tremor rat model of Canavan disease. Brain Res Mol Brain Res 2005; 135:112.
  109. Ahmed SS, Li H, Cao C, et al. A single intravenous rAAV injection as late as P20 achieves efficacious and sustained CNS Gene therapy in Canavan mice. Mol Ther 2013; 21:2136.
  110. Francis JS, Markov V, Wojtas ID, et al. Preclinical biodistribution, tropism, and efficacy of oligotropic AAV/Olig001 in a mouse model of congenital white matter disease. Mol Ther Methods Clin Dev 2021; 20:520.
  111. Fröhlich D, Kalotay E, von Jonquieres G, et al. Dual-function AAV gene therapy reverses late-stage Canavan disease pathology in mice. Front Mol Neurosci 2022; 15:1061257.
  112. Corti M, Byrne BJ, Gessler DJ, et al. Adeno-associated virus-mediated gene therapy in a patient with Canavan disease using dual routes of administration and immune modulation. Mol Ther Methods Clin Dev 2023; 30:303.
  113. Surendran S, Shihabuddin LS, Clarke J, et al. Mouse neural progenitor cells differentiate into oligodendrocytes in the brain of a knockout mouse model of Canavan disease. Brain Res Dev Brain Res 2004; 153:19.
  114. Chao J, Feng L, Ye P, et al. Therapeutic development for Canavan disease using patient iPSCs introduced with the wild-type ASPA gene. iScience 2022; 25:104391.
  115. Feng L, Chao J, Tian E, et al. Cell-Based Therapy for Canavan Disease Using Human iPSC-Derived NPCs and OPCs. Adv Sci (Weinh) 2020; 7:2002155.
  116. Zano S, Malik R, Szucs S, et al. Modification of aspartoacylase for potential use in enzyme replacement therapy for the treatment of Canavan disease. Mol Genet Metab 2011; 102:176.
  117. Hull V, Wang Y, Burns T, et al. Antisense Oligonucleotide Reverses Leukodystrophy in Canavan Disease Mice. Ann Neurol 2020; 87:480.
  118. Mutthamsetty V, Dahal GP, Wang Q, Viola RE. Development of bisubstrate analog inhibitors of aspartate N-acetyltransferase, a critical brain enzyme. Chem Biol Drug Des 2020; 95:48.
  119. Stecula A, Hussain MS, Viola RE. Discovery of Novel Inhibitors of a Critical Brain Enzyme Using a Homology Model and a Deep Convolutional Neural Network. J Med Chem 2020; 63:8867.
  120. Nešuta O, Thomas AG, Alt J, et al. High Throughput Screening Cascade To Identify Human Aspartate N-Acetyltransferase (ANAT) Inhibitors for Canavan Disease. ACS Chem Neurosci 2021; 12:3445.
  121. Wang Y, Hull V, Sternbach S, et al. Ablating the Transporter Sodium-Dependent Dicarboxylate Transporter 3 Prevents Leukodystrophy in Canavan Disease Mice. Ann Neurol 2021; 90:845.
  122. Dewulf JP, Wiame E, Dorboz I, et al. SLC13A3 variants cause acute reversible leukoencephalopathy and α-ketoglutarate accumulation. Ann Neurol 2019; 85:385.
  123. Madhavarao CN, Arun P, Anikster Y, et al. Glyceryl triacetate for Canavan disease: a low-dose trial in infants and evaluation of a higher dose for toxicity in the tremor rat model. J Inherit Metab Dis 2009; 32:640.
  124. Baslow MH, Kitada K, Suckow RF, et al. The effects of lithium chloride and other substances on levels of brain N-acetyl-L-aspartic acid in Canavan disease-like rats. Neurochem Res 2002; 27:403.
  125. Assadi M, Janson C, Wang DJ, et al. Lithium citrate reduces excessive intra-cerebral N-acetyl aspartate in Canavan disease. Eur J Paediatr Neurol 2010; 14:354.
Topic 1702 Version 25.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟