Download PDF
Review  |  Open Access  |  7 Apr 2022

Lysosomal storage disorders with neurological manifestations

Views: 2338 |  Downloads: 1009 |  Cited:   0
Rare Dis Orphan Drugs J 2022;1:6.
10.20517/rdodj.2021.05 |  © The Author(s) 2022.
Author Information
Article Notes
Cite This Article

Abstract

Lysosomal storage disorders (LSDs) constitute a large group of rare, multisystemic, progressive, inherited disorders of metabolism. The aberrant metabolic processes often lead to the cellular accumulation of incompletely metabolized macromolecules or their metabolic byproducts. Most of the patients affected by LSD can experience a variety of neurological presentations including, but not limited to, psychiatric complications, seizures, and/or developmental delays. The onset of symptoms can range from birth to adulthood, and disease severity can vary. Since there is significant overlap in the symptomatology of LSDs, diagnosis is typically confirmed through biochemical and molecular assays. There are currently no approved cures for any LSDs; however, in most cases, treatment of symptoms can lead to better outcomes and improvements in quality of life. The use of hematopoietic stem cell transplantation, enzyme replacement or substrate reduction therapy, and viral vector gene transfer is the subject of many ongoing and completed clinical trials. In this mini review, we provide an overview of LSDs with neurological manifestations, describe the current endeavors in alleviating peripheral symptoms and discuss effective therapeutics strategies.

Keywords

Lysosomal storage disorders, hematopoietic stem cell transplantation, enzyme replacement therapy, gene therapy, AAV viral vectors

OVERVIEW OF LYSOSOMAL STORAGE DISORDERS

Lysosomal storage disorders (LSD) are characterized by the accumulation of toxic substrates in the cells due to mutations in genes encoding lysosomal proteins[1]. There are over seventy distinct disorders classified as LSDs, many of them having an affected lysosomal hydrolase enzyme which is directly implicated in the aberrant processing and degradation of the accumulating substrates[1,2]. Others are due to mutations on lysosomal membrane proteins or proteins involved in the transport of lysosomal enzymes[1]. In addition to the specific mutation involved, oxidative stress and inflammation are also key in LSD pathophysiology and define its progression. Autoimmune responses to specific accumulated macromolecules or metabolites, as well as neuroinflammation-driven microglial activation and astrogliosis, cause activation of apoptotic mechanisms leading to irreversible cellular damage and organ dysfunction[3,4].

Most LSDs are monogenic, autosomal recessive disorders, although three of them (Fabry, Danon, and Hunter syndromes) are X-linked recessive. A study of the Australian population estimated that collectively, LSDs have a combined prevalence of 1 per 7700 live births, but individually are rare-to-ultra-rare; Gaucher disease is the most common LSD, with a prevalence of 1 per 57,000 while sialidosis has the lowest estimated prevalence with 1 in 4.2 million[5]. Despite the global distribution of LSDs, some syndromes cluster in high-risk populations. For example, the Ashkenazi Jewish population has a higher prevalence of Gaucher, Tay-Sachs, and Niemann-Pick diseases[6], Scandinavian and Russian populations have an increased frequency of Hurler syndrome, and Finnish populations have a higher incidence of aspartylglucosaminuria or neuronal ceroid lipofuscinosis[7,8]. However, these prevalence statistics most likely represent an underestimation of their true magnitude due to the frequent misdiagnosis or underdiagnosis that occurs in LSDs[9]. This, in combination with the clinical heterogenicity even within the same disorder, is a significant complication for epidemiological studies in LSD, leading to poor or incomplete data that do not reflect the real-world impact of these diseases in the community.

A few LSDs do have treatment options, including both enzyme replacement therapies (ERTs), which can alleviate some symptoms by providing the missing enzyme[10-12], and substrate reduction therapy (SRT), which can reduce the amount of accumulated toxic substrate by decreasing its biosynthesis[13]. Although ERT is often a viable option for the treatment of peripheral organs of non-neurological LSDs, it is not an option for LSDs with neurological involvement because of the inability of most of the enzyhmes to cross the blood-brain-barrier (BBB). Cerebrospinal fluid (CSF) delivery of the enzymes has emerged as a solution to circumvent the BBB[14], and several clinical trials have been initiated to test safety and efficacy. Current data show some degree of effect in MSP and CLN clinical trials, although efficacy is limited if not treated in very early stages. An alternative to rescue or restore the loss-of-function in LSDs with neurological involvement is gene therapy. Adeno-associated viruses (AAVs) are the most common vector used to direct the expression of the therapeutic transgene into the central nervous system (CNS)[15], and several clinical trials are in progress to probe the safety and efficacy of rescuing the enzymatic activity[16]. In addition to bypassing the BBB, one of the advantages of the CNS delivery of viral vectors is axonal transport, the capacity of the viral particles to be transported from the injection site to distal interconnected structures[17-19]. Axonal transport facilitates broader distribution of the therapeutic into cortical and subcortical structures of brain compared to ERT injections into the CSF of brain parenchyma[20].

CLINICAL FEATURES AND NEUROLOGICAL MANIFESTATIONS

The inability of lysosomes to perform their physiological function because of an enzyme deficiency is the common feature of all LSDs. There is a great degree of variability in age of onset and clinical manifestations, depending on the accumulating substrate involved, which complicates efforts to standardize the clinical features [Table 1]. Common clinical features include physical development delay, difficulty of breathing, “cherry-red spot”, deafness, bone deformities, hydrops, seizures, and ataxia[21]. The onset of symptoms varies greatly, from birth to late adolescence and adulthood[21]. Many patients with LSDs are born with no apparent pathological phenotype for several months after birth, until hepatosplenomegaly or cardiomyopathy is evident, generally in non-neuropathic types[21]. Types with predominant neuropathic involvement are often associated with cognitive and motor developmental delay[22], leading to mental retardation, progressive neurodegeneration, and premature death[23]. Even then, these distinctions are not always clear. Gaucher’s disease, for example, exemplifies the wide array of clinical manifestations associated with LSDs. Type I patients have no neurological symptoms, while type II and III patients can experience ataxia, myoclonic seizures, deafness, dementia, and psychosis[21,24]. This variability in clinical features is related to the particular defect of the malfunctioning enzyme and the accumulating substrate. In the case of Gaucher, glucosylceramides (GCS) accumulate due to deficiency in glucocerebrosidase activity[25], and different origins of GCS cause different clinical manifestations and disease progression. In the brain, GCS are derived primarily from gangliosides, while in peripheral organs, GCS are derived from the breakdown of blood cells. Type I maintains sufficient residual enzyme activity in the brain to avoid ganglioside GCS accumulation, while types II and III patients show lower residual enzymatic activity in the brain and GCS accumulation inside the neurons leading to aberrant inflammatory and apoptotic responses[24]. GM1 and GM2 gangliosidosis and metachromatic leukodystrophy (MLD) are typically associated with more severe neurologic outcomes, including visual loss, cerebellar atrophy associated with ataxia, seizures, and peripheral neuropathy[21]. Type A and Type B Niemann-Pick disease (NPD-A and NPD-B, respectively) are another notable example of a pair of LSDs that share a common deficient enzyme, acid sphingomyelinase (ASM), which is crucial in sphingolipid homeostasis and membrane turnover, but presents with different symptomatology[26]. Individuals with both NPD-A and NPD-B exhibit hepatosplenomegaly, progressive lung disease, and failure to gain weight[27,28], and develop large, lipid-laden foam cells accumulate in affected organs as a result of sphingolipid backup[29]. Involvement of the central nervous system in enzyme deficiency is the primary distinguishing feature between NPD-A and NPD-B; NPD-A is also characterized by rapidly progressive neurodegeneration. Most infants with NPD-A are diagnosed early in their first year of life, do not meet critical developmental milestones, and typically do not survive past their third year of life[27,30]. In stark contrast, patients with NPD-B display no signs of central nervous system involvement and often live into adulthood with medical management of peripheral symptoms. Small deletions or nonsense mutations in the ASM polypeptide, as well as missense mutations in the gene encoding the production of the acid sphingomyelinase, cause Type A. In Type B, a separate missense mutation produces defective acid sphingomyelinase with minimal residual activity, leading to the nonneuronopathic presentation of the disease[31]. While results from clinical trials have demonstrated that ERT can address the peripheral pathology of NPD[32,33], potential gene therapy treatments to address the CNS pathology seen in NPD-A are in the pre-clinical stages[34].

Table 1

A collection of neuropathic lysosomal storage disorders

DiseaseDefective proteinNeurological clinical signsGene (locus)OMIM
AspartylglucosaminuriaAspartylglucosaminidaseES, DDAGA (4q34.3)208400
Fabry diseaseAlpha-galactosidaseCC, CAT, LK, DF, CV, DP, PNGLA (Xq22.1)301500
Farber diseaseAcid ceramidaseDDASAH1 (8p22)228000
Fucosidosisα-L-fucosidaseMS, ES, DDFUCA1 (1p36.11)230000
GalactosialidosisProtective protein cathepsin AOA, MS, DF, AX, DDCTSA (20q13.12)256540
Gaucher disease Type IIGlucocerebrosidaseMS, DF, DDGBA (1q22)230900
231000
Gaucher disease Types IIIGlucocerebrosidaseOA, MS, AX, ES, DP, DD
GBA (1q22)230900
231000
Glycogenosis II/Pompe diseaseα-GlucosidaseDDGAA (17q25.3)232300
GM1 gangliosidosis (Types I, II, III)GM1-β-galactosidaseOA, CA, AX, ES, DP, DDGLB1 (3p22.3)230500
230600
230650
GM2 gangliosidosis, AB variantGM2 activator proteinMS, CA, AX, ES, DPGM2A (5q33.1)272750
GM2 gangliosidosis, Sandhoff variantβ-Hexoaminidase A + BOA, MS, MC, CA, AX, ES, DPHEXB (5q13.3)268800
GM2 gangliosidosis, Tay-Sachs variantβ-Hexosaminidase AOA, MS, MC, CA, AX, ES, DP, DDHEXA (15q23)272800
Krabbe diseaseGalactocerebrosidaseLK, MC, PN, AX, ES, DP, DDGALC (14q31.3)245200
Metachromatic leukodystrophyArylsulfatase ALK, DF, PN, CA, AX, DP, DDARSA (22q13.33)250100
MPS I (Hurler, Scheie, Hurler-Scheie)α-L-iduronidaseCC, DF, DDIDUA (4p16.3)607014
607015
607016
MPS II (Hunter)Iduronate-2-sulfataseDP, DD, DF, OA, MS, MCIDS (Xq28)309900
MPS III A (Sanfilippo)heparan N-sulfataseDP, DD, DF, OA, MS, MCSGSH (17q25.3)252900
MPS III Bα-N-Ac-glucosaminidaseDP, DD, DF, OA, MS, MCNAGLU (17q21.2)252920
MPS III Cheparan-α-glucosaminide N-acetyltransferaseDP, DD, DF, OA, MS, MCHGSNAT (8p11.21-p11.1)252930
MPS III DN-acetylglucosamine 6-sulfataseDP, DD, DF, OA, MS, MCGNS (12q14.3)252940
MPS IV A (Morquio A)N-acetylgalactosamine 6-sulfataseOA, CC, DF, DDGALNS (16q24.3)253000
MPS IV BBeta-galactosidaseOA, CC, DDGLB1 (3p22.3)253010
MPS VI (Maroteaux-Lamy)Arylsulfatase BCC, DDARSB (5q14.1)253200
MPS VII (Sly)Beta-glucuronidaseOA, DDGUSB (7q11.21)253220
Mucolipidosis IIUDP-N-Ac-glucosaminyl phosphotransferaseOA, CC, DF, CA, DDGNPTAB (12q23.2)252500
Mucolipidosis IVMucolipin 1CC, AX, DDMCOLN1 (19p13.2)252650
Multiple sulfatase deficiencySulfatase modifier proteinPN, DDSUMF1 (3p26.1)272200
Neuronal Ceroid Lipofuscinosis 1Palmitoyl-protein thioesterase 1OA, RP, MS, CA, AX, DP, DDPPT1 (1p34.2)256730
Neuronal Ceroid Lipofuscinosis 2Tripeptidyl peptidase 1OA, RP, MS, CA, AX, DP, DDTPP1 (11p15.4)204500
Niemann-Pick type A (Neurovisceral Type)
SphingomyelinaseOA, DDSMPD1 (11p15.4)257200
Niemann-Pick type B (Visceral Type)SphingomyelinaseOA, DDSMPD1
(11P15.4)
607616
Niemann-Pick type C1NPC1OA, MS, AX, ES, DP, DDNPC1 (18q11.2)257220
Niemann-Pick type C2NPC2OA, MS, AX, ES, DP, DDNPC1 (18q11.2)607625
Saposin defect, Gaucher typeSaposin COA, MS, AX, ES, DP, DDPSAP (10q22.1)610539
Saposin defect, generalized typeProsaposinAX, CA, DDPSAP (10q22.1)611721
Saposin defect, MLD typeSaposin BLK, DF, PN, CA, AX, DP, DDPSAP (10q22.1)249900
Schindler diseaseα-N-acetylgalactosaminidaseOA, MS, ES, DDNAGA (22q13.2)609241
Sialic acid storage disease, infantile (ISSD), adult (Salla)SialinOA (ISSD), AX (Salla), DDSLC17A5 (6q13)604369
SialidosisSialidaseOA, CAT, MS, AX, ES, DDNEU1 (6p21.33)256550
Wolman disease/CESDAcid lipaseAX, DP, DDLIPA (10q23.31)278000
α-Mannosidosisα-MannosidaseCC, CAT, DF, ES, DDMAN2B1 (19p13.13)248500
β-Mannosidosisβ-MannosidaseDF, ES, DDMANBA (4q24)248510

Age-of-onset of the disease phenotype also has a significant impact on both severity and progression of the disease, generally led by the genomic background or the causal mutations involved in the disease. Fabry disease, for example, is an X-linked LSD caused by deficiency of the enzyme alpha-galactosidase, which results in the accumulation of globotriaosylceramide in the brain, liver, skin, heart, and kidneys[35]. In early-onset Fabry disease, patients typically show hallmark symptoms including angiokeratoma, hypohidrosis, corneal opacities, and neuropathic pain, often in the form of acroparesthesias[35,36], while late-onset patients may instead present severe renal, cardiac, and vascular manifestations as the disease progresses[37]. Especially in young/early-onset patients, LSDs are also often misdiagnosed as psychiatric disorders due to the presence of similar symptoms such as anxiety, restlessness, aggression, and increased sensitivity to touch[38]. As the disease progresses, other psychiatric symptoms start to be more evident, like hallucinations, psychosis, and dementia[38]. In patients with late-onset Tay Sachs Disease, up to 20%-40% of affected individuals will present with psychiatric symptoms years before the onset of motor deficits[39]. Infantile diagnoses of MLD present with motor symptoms such as gait instability and loss of motor developmental milestones, while adults with MLD demonstrate behavioral disturbances and dementia, which may be mistakenly diagnosed as psychosis[40]. Patients with Niemann-Pick Type C (NP-C) may develop neuropsychiatric symptoms at any point in their disease progression, including agitations, sleep disorders, and depression[41].

Despite advances in the identification and treatment of LSDs, affected individuals are often diagnosed with a psychiatric disorder without appropriate follow-up for their underlying disease. One study reported an average delay of 6.2 years between initial symptom presentation and diagnosis in patients with NP-C, as the majority of patients who presented with psychosis as their initial symptom of disease did not demonstrate any neurological exam abnormalities and were given a diagnosis of schizophrenia or other forms of psychosis[42]. Individuals who are being evaluated for a psychiatric disorder, particularly those who also present with motor deficits or who do not respond to standard psychiatric management, may greatly benefit from a rigorous metabolic work-up[38].

DIAGNOSIS, GENETIC TESTING AND METABOLOMICS

Antecedents of the disease in the family (proband), or positive screening of family members (carrier) can increase the identification of presymptomatic individuals through prenatal testing or newborn screening, but most often, the lack of symptoms, especially in the carriers, causes diagnostic confirmation to occur after multiple organs have been irreversibly damaged and therefore therapeutic interventions are less efficacious[43]. Early identification of positive cases at prodromal stages would facilitate the design and execution of a clinical management plan when the therapeutic intervention has a better chance of success[44]. However, the ethical dimension of early-in-life screening programs is still an active discussion among genetic counselors, health-care professionals, and health policymakers[45].

In general, the onset of symptoms may present anywhere between infancy and later adulthood and continue worsening over time, chronifying the disease or causing death[21]. Although mutational and enzymatic analyses stablish the diagnostic in all cases, biochemical markers in presymptomatic individuals can predict cases in some LSD. For example, the mucopolysaccharidoses group of LSDs, increased urinary excretion of glycosaminoglycans can be detected, while increased levels of oligosaccharides will be detected in the urine of patients in the oligosaccharidoses group[46]. In Pompe Disease, increased blood levels of creatine kinase can be found, while patients with Gaucher or Niemann-Pick C Disease may demonstrate elevated levels of chitotriosidase in serum[46]. Currently, measurement of lysosomal enzyme activity can reliably diagnose most LSDs by identifying a severe deficiency. Fluorometry[47,48], tandem mass spectrometry[49,50], and radioactive assays are the principal techniques for enzymatic activity assessment. However, these assays can be tissue-dependent, leading to false-negative results if the wrong specimen is used in the assay[51]. Also, it is important to know that the complete absence of lysosomal enzyme activity generally leads to an LSD diagnosis, but conversely, normal lysosomal enzyme activity is not enough to discard LSD diagnosis if there is a clinical phenotype suggestive of LSD[46]. In certain variant forms of GM2 gangliosidosis, Krabbe disease, MLD, and Gaucher disease, enzyme levels could be normal, but disease phenotype may surface due to a defect in saposins, a group of glycoproteins that activate lysosomal hydrolases related to sphingolipid metabolism[52].

Genetic confirmatory analysis is being utilized with increasing frequency as laboratory techniques and bioinformatic data management are advancing. Conventional methods for genetic confirmation of LSDs like MLPA, RFPL, or Sanger sequencing are limited by the fact that they analyze a single mutation at a time. Modern techniques like next-generation sequencing (NGS), including whole-genome sequencing and whole-exome sequencing, can identify causal genetic variants of the monogenic disease, including neurometabolic disease streamlining the mutation screening in the diagnostic process[53]. These high-throughput DNA sequencing technologies are time and cost-effective and can interrogate a large number of genes in a single reaction[54,55]. For example, in 2013, Hoffman and colleagues concluded that NGS can be superior in identifying Tay-Sachs Disease carriers compared to traditional enzyme and genotype analyses which can be limited by both false positives and negatives[56], and in 2016, Yoshida and colleagues demonstrated the application of NGS in prenatal diagnosis of Gaucher disease even in the absence of prior genetic information from the family[57].

New challenges have been introduced as these high-throughput technologies advance. Higher sophistication of laboratory techniques, including new reagents and new equipment, is needed, but the biggest challenge lies at the bioinformatic level, which includes a new requirement for ‘big data’ management and accessible data interpretation[58]. Certainly, new intersections between different specialties like informatics, statistics, and biology, as well as new emerging fields of science like data mining or artificial intelligence, are facilitating the accessibility of these new techniques to the clinic. Genetic counseling is also an important consideration of these new techniques, especially in LSDs[59]. Counseling for families with LSD generally include education of the disease, inheritance patterns, recurrence risks, and implications of the diagnosis, progression, and needs for ongoing medical and family management. Currently, no curative therapies are approved, with current disease management primarily focusing on the control of symptoms in LSD patients via ERT. Counseling and carrier screening are essential in the management and mitigation of the incidence and prevalence of theses genetic disorders. One successful example of management through genetic counseling is in the Ashkenazi Jewish population[60], where carrier screening in individuals with Ashkenazi ancestry resulted in a dramatic decrease in the incidence of some LSD, including Tay Sachs, Canavan, Gaucher, or Niemann-Pick[61,62].

AVAILABLE TREATMENT AND CLINICAL TRIALS FOR LSD WITH CNS INVOLVEMENT

Several symptomatic treatment options are available for LSDs. Enzymatic or protein deficiencies can be restored by cellular or recombinant proteins like hematopoietic stem cells transplantation (HSCT) or ERT, respectively, and the buildup of metabolites can be reduced through SRT. Biological treatments can be combined with physical and occupational therapy to improve outcomes[63].

Hematopoietic stem cells transplantation

In use for over 20 years, HSCT originally involved bone marrow transplantation and has advanced to using unrelated umbilical cord blood as a source of stem cells[64-66]. HSCT is primarily performed in patients with mucopolysaccharidoses and leukodystrophies and ultimately functions as an ERT. Healthy engrafted donor cells not only provide a continuous endogenous supply of enzyme in the extracellular space and into the blood circulation, but also microglia derived from the donor stem cells can migrate to the brain and secrete functional enzyme in the CNS improving neurocognitive function and quality of life[67]. HSCT is the gold standard treatment in young Hurler’s syndrome patients with mild to no cognitive impairment, showing a significant reduction in mortality[65,68,69]. Important limitations are associated with this technique, including graft failure and high morbidity and mortality rates. Leukocyte enzyme levels may also not reach target levels for several years, potentially limiting the effectiveness of this treatment modality in patients with rapidly progressing disease. Additionally, in treated Hurler’s syndrome patients, skeletal complications continue to develop, and the majority of the patient will require surgical intervention[65,66,70].

Enzyme replacement therapy

ERT has been shown to be an effective therapy in ameliorating the visceral, hematological, and biochemical manifestations for some LSDs, substantially improving the patient quality of life. Treatments for Gaucher, Fabry, MPS, and Pompe diseases are a clear example of successful ERT. Weekly or bimonthly intravenous injections provide broad biodistribution of the recombinant enzyme that is internalized by cells and conducted to the lysosomal compartment. This internalization is mediated by mannose or the mannose-6-phosphate receptor following the endocytic route and rescuing the lysosomal enzyme deficit. Despite the successes of ERT, the production of IgG antibodies against recombinant enzyme (anti-drug antibodies) can lead to diminished activity or cellular uptake[71]. However, recent studies reported the benefit of the immune tolerance induction in some patients[72]. This technique aims to diminish the development or minimize the consequences of the response against the ERT, especially in cross-reactive immunologic material -negative or null patients[73,74]. Another major limitation is the inability of intravenously administered functional enzyme to cross the BBB. This is especially problematic given that about two-thirds of LSDs present with neurologic symptoms[75]. Current research is focused on ensuring the accessibility of the CNS to treatment modalities, including modifications to the enzyme that allow passage through the BBB, delivering a high intravenous dose of enzyme, or intrathecal injections[75].

Substrate reduction therapy

SRT aims to mitigate the effects of substrate accumulation. SRT uses small molecules to inhibit the biosynthesis of compounds that are accumulating problematic levels in the absence of functional lysosomal enzyme[76,77]. In Gaucher disease, accumulation of glucosylceramide can be targeted with miglustat, an oral glucosylceramide synthase inhibitor. This decreased rate of accumulation allows residual enzyme activity to more effectively clear lysosomal stores of glucosylceramide, slowing disease progression[78]. Miglustat has also demonstrated similar efficacy in the treatment of NPD-C, with the potential for extending patients’ lives[79]. More recently, eliglustat, a more potent ceramide-mimetic inhibitor that demonstrated less severe side effects than miglustat, was developed. Unfortunately, eliglustat does not cross the BBB, limiting its efficacy in treating the neuropathic types of Gaucher disease[80,81]. SRT and ERT can be used in conjunction with other therapies, including each other. In Fabry disease mouse models, the therapeutic effect of the combination of SRT and ERT was found to be additive and complementary[82]. The combination treatment may allow for reduced frequency of ERT, potentially leading to an increase in quality of life due to decreased dependency on enzyme infusions[83]. This combination of treatments could be beneficial for patients who demonstrate a complete absence of enzymatic activity.

Gene therapy

Gene therapy has seen substantial improvement in recent years, especially for LSDs. With this approach, functional copies of defective genes are introduced into the human cells to treat or prevent diseases. This “genome editing” technique can be performed ex-vivo or in-vivo. Adeno-associated viral vectors (AAV) and retroviral vectors have been extensively used in pre-clinical studies of LSDs with CNS implications like mucopolysaccharidosis types I[84], III[85], and VII[86], Krabbe disease[87], metachromatic leukodystrophy[88], Niemann-Pick[34], ceroid lipofuscinosis[89], or GM1 gangliosidosis[90]. One major benefit of gene therapy over ERT is that gene therapy is administered in a single dose. During transfection, the new genetic material is introduced into the cell, engaging existing cellular machinery to produce the protein of intereste, with the goal of restoring missing functionality. These vectors can be introduced intravenously, although, for neuropathic LSD, parenchymal or intraventricular/intrathecal have been demonstrated to have better results[16,91,92]. As with other systemic approaches, the ability to cross the BBB is also an important limitation for vector distribution into the brain. Since AAV serotype 9 is the only vector capable of crossing the BBB, it is currently the only viable option for AAV-based systemically administered therapy[93,94]. To achieve significant and widespread effect in the CNS after systemic delivery, large volumes of the viral vector are required, and these high doses have been associated with sensory neuron toxicity in some cases[95,96]. In contrast, CSF delivery needs less total load of viral vector to achieve significant levels of transduction in the brain and spinal cord[94,97], but recently, dorsal root ganglia histopathology with no clinical signs has also been reported in a meta-analysis in non-human primates[98]. Parenchymal delivery, although technically more invasive, has so far shown no adverse effects related to transgene-overexpression and remains the platform of choice in multiple clinical trials. Once in the brain, there are two characteristics differentiate gene therapy from other therapeutic approaches: axonal transport and the bystander effect. Axonal transport is the property by which viral particles use the projections between neurons to travel to different brain structures, increasing the distribution beyond the injection site[17,99]. Convection enhanced delivery is a bulk-flow mechanism used to convey and distribute macromolecules into brain parenchymal tissue improving the local distribution into the injection site and reducing the off-target delivery[100,101]. Altogether, this efficient local transduction improves the distal transduction by reaching a larger number of projections. Because most affected enzymes in CNS-involved LSDs are secreted, treatments benefit from the cross-correction of bystander cells. Once some of the cells are transduced by the vectors, functional enzymes are produced then secreted and internalized by mannose-6-phosphate receptors-mediated endocytosis by other neighboring enzyme-deficient cells[102], enhancing the distribution of the therapeutic effect into cells that were not transduced[34,103]. Multiple examples define the proof-of-concept for this therapeutic approach in LSDs,[104-106] but without doubt, from the early mouse models for Mucopolysaccharidosis[107] to the current large animal models for LSDs[108-110], the existence of a large selection of animal models propelled the better understanding of the underlying disease mechanisms, accelerating the pre-clinical development[111]. However, species-specific differences in cellular metabolism are the basis of differences in the pathological phenotype that sometimes do not effectively recapitulate what is seen in patients, especially in CNS, where brain complexity can be challenging to model in animals.

Because of the poor efficacy in the CNS of the available therapies, including systemic ERT, most of the current clinical trials involve the use of gene therapy to correct neurological defects. Detailed information of the clinical trials is available through clincialtrials.gov, and the most recent active trials for lysosomal storage disorders with neurological deficits are summarized in Table 2.

Table 2

Active clinical trials for neuropathic lysosomal storage disorders

DiseaseVectorPhaseNCT NumberStatus
FabryAAVPhase 1/2NCT04455230Recruiting
FabryAAV2/6Phase 1/2NCT04046224Recruiting
FabryAAV 4D-310Phase 1/2NCT04519749Recruiting
FabryLentiviral-HSC GTPhase 1/2NCT03454893Recruiting
FabryLentiviral-HSC GTPhase 1NCT02800070Active, not recruiting
GM1 GangliosidosisAAVrh.10Phase1/2NCT04273269Recruiting
GM1 GangliosidosisAAVHu68Phase 1/2NCT04713475Recruiting
GM1 Gangliosidosis Type IIAav9Phase 1/2NCT03952637Recruiting
GM2 Gangliosidosis (Inftantile-Onset)AAV9Phase 1/2NCT04798235Recruiting
KrabbeHSC GT + AAVrh10Phase 1/2NCT04693598Recruiting
Krabbe (Early infantile)AAVHu68Phase 1/2NCT04771416Recruiting
Late Juvenile Metachromatic LeukodystrophyLentiviral-HSC GTPhase 3NCT04283227Recruiting
Metachromatic LeukodystrophyLentiviral-HSC GTPhase 2NCT03392987Active, not recruiting
Metachromatic LeukodystrophyLentiviral-HSC GTPhase 1/2NCT01560182Active, not recruiting
Metachromatic LeukodystrophyLentiviral-HSC GTPhase 1/2NCT02559830Recruiting
MPS Type IrAAV9Phase 1/2NCT03580083Recruiting
MPS Type ILentiviral-HSC GTPhase 1/2NCT03488394Active, not recruiting
MPS Type IZFN-AAVPhase 1/2NCT02702115Active, not recruiting
MPS Type IAutologous PlasmablastsPhase 1/2NCT04284254Not yet recruiting
MPS Type IIAAV9Phase 1/2NCT04573023Recruiting
MPS Type IIrAAV9Phase 1/2NCT03566043Recruiting
MPS Type IIaAAVrh10Phase 2/3NCT03612869Active, not recruiting
MPS Type IIIascAAV9Phase 1/2NCT04088734Recruiting
MPS Type IIIaLentiviral-HSC GTPhase 1/2NCT04201405Active, not recruiting
MPS Type IIIascAAV9Phase1/2NCT02716246Recruiting
MPS Type IIIbrAAV9Phase 1/2NCT03315182Active, not recruiting
MPS Type IVAAV2/8Phase 1/2NCT03173521Recruiting
Neuronal Ceroid Lipofuscinosis 3scAAV9Phase 1/2NCT03770572Active, not recruiting
Neuronal Ceroid Lipofuscinosis 6scAAV9Phase 1/2NCT02725580Active, not recruiting
Neuronal Ceroid Lipofuscinosis 7rAAV9Phase 1NCT04737460Recruiting
Pompe (Late-onset)AAVPhase 1/2NCT04093349Recruiting
Pompe (Late-Onset)AAV8Phase 1/2NCT04174105Recruiting
Pompe (Late-Onset)AAV9Phase 1NCT02240407Active, not recruiting
SandhoffAAVrh.8Phase 1NCT04669535Recruiting
Tay-SachsAAVrh.8Phase 1NCT04669535Recruiting

ONGOING DEVELOPMENT: NEW APPROACHES, NEW HOPES

A wide collection of studies continues to examine the utility of AAV and its related variants in the treatment of neuropathic LSD. New natural capsids are being discovered by high-throughput screening or isolated from other vertebrate species that are under investigation for potential therapeutic use. Novel capsids are also being engineered by rational design or directed evolution. One notable example is the discovery of AAV-PHP.B, an engineered capsid identified by Cre recombination-based AAV targeted evolution, a directed evolution method that uses Cre/lox technology[112], and is capable of crossing the BBB and transducing the CNS[113]. More recently, barcoded rational AAV vector evolution optimized a screening process used to identify the optimal capsid from multiple rounds to a single in vivo screening round[114].

While AAV-based technologies seek to circumvent an aberrant mutation, other investigational strategies aim to precisely manipulate the patient genome. Currently, meganucleases, zinc finger nucleases, TALENs, and CRISPR/Cas can modulate gene expression by modifying DNA, RNA, or transcription factors. Zinc finger nucleases (ZFNs), engineered proteins comprised of the non-specific cleavage domain of FokI endonuclease and zinc finger proteins[115], showed promising results in hemophilia animal models[116,117], and lead to the initiation of phase 1 clinical trial (NCT02695160). Following this development, ZFNs have been tested in models of Gaucher disease, Fabry disease, MPS type I, and MPS type II, with the administration of AAV8-ZFNs and AAV8 carrying the gene for the relevant lysosomal enzyme to animal models[118-120], and in MSP type I a first-in-human clinical trial (NCT02702115). Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR) has also garnered much attention in the field of gene editing. On the therapeutic level, this molecular technology utilizes the RNA-mediated adaptive immune system found in bacteria to modify, remove, or add genes in vivo[121,122]. An experiment utilizing an intravenously administered novel AAV8 system carrying CRISPR to insert a promoterless α-l-iduronidase (IDUA) cDNA sequence into the albumin locus of hepatocytes in MPS type I-affected neonatal and adult mice resulted in IDUA activity in the brain and reduction of pathology in peripheral organs, as well as notable behavioral improvements, learning ability, and restored memory functionality[123]. A similar strategy was used in models of Tay-Sachs and Sandhoff diseases[124], and in NPC1 mouse model to assess its treatment potential in Niemann-Pick disease type C. The study showed robust base editing in brain, liver, heart, retina, and skeletal muscle cells, leading to a 9.2% increase in lifespan and higher Purkinje neuron survival[125].

Crossing the BBB remains a difficult challenge in searching for effective pharmacological candidates for CNS therapy. Currently, new low molecular weight therapeutic drugs are under development to enhance their interaction with luminal receptors to cross the BBB, like the IgG-IDS fusion protein that showed high penetrance into the brain and spinal cord after intravenously injected in the MSP Type I mouse model[126,127]. Other pharmacological strategies to efficiently access the CNS include nanotechnologies like nanoparticles-based polymers and lipid nanoparticles. Nanoparticles formulated with polylactide-coglycolide (PLGA) showed efficient BBB crossing in MPS type I model and suggest a possible intervention with PLGA loaded with rhGAA (human acid α-glucosidase) [128,129].

As prenatal genetic diagnostics continue to advance in safety and accessibility, the possibility of early diagnosis and intervention of LSDs is becoming a reality. Several studies have now examined in utero ERT[130], in utero hematopoietic stem cell therapy[130-132], early gene therapy[133-136], and gene editing[137] for the treatment of various lysosomal storage disorders. Benefits of targeting the disease in utero include small fetal size, the tolerant fetal immune system, the presence of highly proliferative and developing stem/progenitor cells, and the potential to treat diseases in which devastating pathology begins prior to birth. The studies tend to target LSDs that have multi-organ dysfunction and in utero or perinatal morbidity and mortality, such as MPS type I, MPS type VII, and neuronopathic Gaucher disease. Results from these studies have contributed to the first-in-human fetal ERT clinical trial for the treatment of MPS types I, II, IVa, VI, VII, infantile-onset Pompe disease, Neuronopathic Gaucher Disease (types 2 and 3), and Wolman disease (NCT04532047).

CONCLUSION

Overall, LSDs are multisystemic diseases with a very heterogeneous clinical etiopathology in which progression, complexity, and severity are defined by very diverse factors like genetic background or residual levels of enzymatic activity. New developments, especially in gene therapy, have the potential to improve the efficacy of these interventions by increasing levels of enzymatic expression, both systemically and cellularly, but early diagnosis and pre-symptomatic intervention are essential to ensure therapeutic benefits for those with neurological involvement before irreversible neurological consequences appear. This underscores the need for cost-effective, universal newborn screening by which patients can be rapidly identified and treated.

DECLARATIONS

Authors’ contributions

Wrote the manuscript: Munjal V, Thompson-Clarke M, Vignolles-Jeong J, Valencia JA, Samaranch L

Edited the manuscript: Travis M, Samaranch L

Availability of data and materials

Not applicable.

Financial support and sponsorship

None.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2022.

REFERENCES

1. Marques ARA, Saftig P. Lysosomal storage disorders - challenges, concepts and avenues for therapy: beyond rare diseases. J Cell Sci 2019;132:jcs221739.

2. Platt FM, d’Azzo A, Davidson BL, Neufeld EF, Tifft CJ. Lysosomal storage diseases. Nat Rev Dis Primers 2018;4:27.

3. Bosch ME, Kielian T. Neuroinflammatory paradigms in lysosomal storage diseases. Front Neurosci 2015;9:417.

4. Donida B, Jacques CED, Mescka CP, et al. Oxidative damage and redox in lysosomal storage disorders: biochemical markers. Clin Chim Acta 2017;466:46-53.

5. Meikle PJ, Hopwood JJ, Clague AE, Carey WF. Prevalence of lysosomal storage disorders. JAMA 1999;281:249-54.

6. Vallance H, Ford J. Carrier testing for autosomal-recessive disorders. Crit Rev Clin Lab Sci 2003;40:473-97.

7. Santavuori P. Neuronal ceroid-lipofuscinoses in childhood. Brain and Development 1988;10:80-3.

8. Arvio M, Autio S, Louhiala P. Early clinical symptoms and incidence of aspartylglucosaminuria in Finland. Acta Paediatr 1993;82:587-9.

9. Giugliani R, Federhen A, Michelin-Tirelli K, Riegel M, Burin M. Relative frequency and estimated minimal frequency of lysosomal storage diseases in brazil: report from a reference laboratory. Genet Mol Biol 2017;40:31-9.

10. Brady RO. Enzyme replacement for lysosomal diseases. Annu Rev Med 2006;57:283-96.

11. Beck M. Treatment strategies for lysosomal storage disorders. Dev Med Child Neurol 2018;60:13-8.

12. Thomas R, Kermode AR. Enzyme enhancement therapeutics for lysosomal storage diseases: Current status and perspective. Mol Genet Metab 2019;126:83-97.

13. Coutinho MF, Santos JI, Alves S. Less is more: substrate reduction therapy for lysosomal storage disorders. Int J Mol Sci 2016;17:1065.

14. Bellettato CM, Scarpa M. Possible strategies to cross the blood-brain barrier. Ital J Pediatr 2018;44:131.

15. Kantor B, Bailey RM, Wimberly K, Kalburgi SN, Gray SJ. Methods for gene transfer to the central nervous system. Adv Genet 2014;87:125-97.

16. Sevin C, Deiva K. Clinical trials for gene therapy in lysosomal diseases with cns involvement. Front Mol Biosci 2021;8:624988.

17. Salegio EA, Samaranch L, Kells AP, et al. Axonal transport of adeno-associated viral vectors is serotype-dependent. Gene Ther 2013;20:348-52.

18. Castle MJ, Perlson E, Holzbaur EL, Wolfe JH. Long-distance axonal transport of AAV9 is driven by dynein and kinesin-2 and is trafficked in a highly motile Rab7-positive compartment. Mol Ther 2014;22:554-66.

19. Green F, Samaranch L, Zhang HS, et al. Axonal transport of AAV9 in nonhuman primate brain. Gene Ther 2016;23:520-6.

20. Ziegler RJ, Salegio EA, Dodge JC, et al. Distribution of acid sphingomyelinase in rodent and non-human primate brain after intracerebroventricular infusion. Exp Neurol 2011;231:261-71.

21. Pastores GM, Maegawa GH. Clinical neurogenetics: neuropathic lysosomal storage disorders. Neurol Clin 2013;31:1051-71.

22. Wraith JE, Clarke LA, Beck M, et al. Enzyme replacement therapy for mucopolysaccharidosis I: a randomized, double-blinded, placebo-controlled, multinational study of recombinant human alpha-L-iduronidase (laronidase). J Pediatr 2004;144:581-8.

23. Bellettato CM, Scarpa M. Pathophysiology of neuropathic lysosomal storage disorders. J Inherit Metab Dis 2010;33:347-62.

24. Vellodi A. Lysosomal storage disorders. Br J Haematol 2005;128:413-31.

25. Aflaki E, Stubblefield BK, Maniwang E, et al. Macrophage models of Gaucher disease for evaluating disease pathogenesis and candidate drugs. Sci Transl Med 2014;6:240ra73.

26. Henry B, Ziobro R, Becker KA, Kolesnick R, Gulbins E. Acid sphingomyelinase. Handb Exp Pharmacol ;2013:77-88.

27. McGovern MM, Aron A, Brodie SE, Desnick RJ, Wasserstein MP. Natural history of Type A Niemann-Pick disease: possible endpoints for therapeutic trials. Neurology 2006;66:228-32.

28. McGovern MM, Lippa N, Bagiella E, et al. Morbidity and mortality in type B Niemann-Pick disease. Genet Med 2013;15:618-23.

29. Pick L. Niemann-Pick’s Disease and Other Forms of so-Called Xanthomatosis. Am J Med Sci 1933;185:615-16. Available from: https://onlinebooks.library.upenn.edu/webbin/serial?id=amjmedsci[Last accessed on 1 Apr 2022].

30. Desnick RJ, Ioannou YA, Eng CM, et al. The metabolic and molecular bases of inherited disease CD-ROM. Mol Pathol 1997;50:279. Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC379650/ [Last accessed on 1 Apr 2022].

31. Takahashi T, Desnick RJ, Takada G, Schuchman EH. Identification of a missense mutation (S436R) in the acid sphingomyelinase gene from a Japanese patient with type B Niemann-Pick disease. Hum Mutat 1992;1:70-1.

32. Diaz GA, Jones SA, Scarpa M, et al. One-year results of a clinical trial of olipudase alfa enzyme replacement therapy in pediatric patients with acid sphingomyelinase deficiency. Genet Med 2021;23:1543-50.

33. Wasserstein MP, Diaz GA, Lachmann RH, et al. Olipudase alfa for treatment of acid sphingomyelinase deficiency (ASMD): safety and efficacy in adults treated for 30 months. J Inherit Metab Dis 2018;41:829-38.

34. Samaranch L, Pérez-Cañamás A, Soto-Huelin B, et al. Adeno-associated viral vector serotype 9-based gene therapy for Niemann-Pick disease type A. Sci Transl Med 2019;11:eaat3738.

35. Ioannou YA, Zeidner KM, Gordon RE, Desnick RJ. Fabry disease: preclinical studies demonstrate the effectiveness of alpha-galactosidase A replacement in enzyme-deficient mice. Am J Hum Genet 2001;68:14-25.

36. Martins AM, D’Almeida V, Kyosen SO, et al. Guidelines to diagnosis and monitoring of Fabry disease and review of treatment experiences. J Pediatr 2009;155:S19-31.

37. Sun A. Lysosomal storage disease overview. Ann Transl Med 2018;6:476.

38. Staretz-Chacham O, Choi JH, Wakabayashi K, Lopez G, Sidransky E. Psychiatric and behavioral manifestations of lysosomal storage disorders. Am J Med Genet B Neuropsychiatr Genet 2010;153B:1253-65.

39. Hurowitz GI, Silver JM, Brin MF, Williams DT, Johnson WG. Neuropsychiatric aspects of adult-onset Tay-Sachs disease: two case reports with several new findings. J Neuropsychiatry Clin Neurosci 1993;5:30-6.

40. Gieselmann V, Matzner U, Hess B, et al. Metachromatic leukodystrophy: Molecular genetics and an animal model. J Inherit Metab Dis 1998;21:564-74.

41. Wraith JE, Baumgartner MR, Bembi B, et al. NP-C Guidelines Working Group. Recommendations on the diagnosis and management of Niemann-Pick disease type C. Mol Genet Metab 2009;98:152-65.

42. Sévin M, Lesca G, Baumann N, et al. The adult form of Niemann-Pick disease type C. Brain 2007;130:120-33.

43. Wang RY, Bodamer OA, Watson MS, Wilcox WR. ACMG Work Group on Diagnostic Confirmation of Lysosomal Storage Diseases. Lysosomal storage diseases: diagnostic confirmation and management of presymptomatic individuals. Genet Med 2011;13:457-84.

44. Mokhtariye A, Hagh-Nazari L, Varasteh AR, Keyfi F. Diagnostic methods for lysosomal storage disease. Rep Biochem Mol Biol 2019;7:119-28[Last accessed on 1 Apr 2022].

45. Hawkins AK, Ho A. Genetic counseling and the ethical issues around direct to consumer genetic testing. J Genet Couns 2012;21:367-73.

46. Filocamo M, Morrone A. Lysosomal storage disorders: molecular basis and laboratory testing. Hum Genomics 2011;5:156-69.

47. Sista RS, Wang T, Wu N, et al. Multiplex newborn screening for pompe, Fabry, hunter, Gaucher, and hurler diseases using a digital microfluidic platform. Clin Chim Acta 2013;424:12-8.

48. Hwu WL, Chien YH, Lee NC, et al. Newborn screening for fabry disease in taiwan reveals a high incidence of the later - onset GLA mutation c. 936+ 919G> A (IVS4 + 919G > A). Hum mutat 2009;30:1397-1405. Available from: https://pubmed.ncbi.nlm.nih.gov/19621417/[Last accessed on 1 Apr 2022].

49. Kasper DC, Herman J, De Jesus VR, Mechtler TP, Metz TF, Shushan B. The application of multiplexed, multi-dimensional ultra-high-performance liquid chromatography/tandem mass spectrometry to the high-throughput screening of lysosomal storage disorders in newborn dried bloodspots. Rapid Commun Mass Spectrom 2010;24:986-94.

50. Zhang XK, Elbin CS, Chuang WL, et al. Multiplex enzyme assay screening of dried blood spots for lysosomal storage disorders by using tandem mass spectrometry. Clin Chem 2008;54:1725-8.

51. Winchester B, Bali D, Bodamer OA, et al. Pompe Disease Diagnostic Working Group. Methods for a prompt and reliable laboratory diagnosis of pompe disease: report from an international consensus meeting. Mol Genet Metab 2008;93:275-81.

52. Valle DL, Antonarakis S, Ballabio A, Beaudet AL, Mitchell GA. The Online Metabolic and Molecular Bases of Inherited Disease. Available from: https://ommbid.mhmedical.com/book.aspx?bookID=2709 [Last accessed on 1 Apr 2022].

53. Giugliani R, Brusius-Facchin AC, Pasqualim G, et al. Current molecular genetics strategies for the diagnosis of lysosomal storage disorders. Expert Rev Mol Diagn 2016;16:113-23.

54. Metzker ML. Sequencing technologies - the next generation. Nat Rev Genet 2010;11:31-46.

55. Hadzsiev K, Komlosi K, Czako M, et al. Kleefstra syndrome in Hungarian patients: additional symptoms besides the classic phenotype. Mol Cytogenet 2016;9:22.

56. Hoffman JD, Greger V, Strovel ET, et al. Next-generation DNA sequencing of HEXA: a step in the right direction for carrier screening. Mol Genet Genomic Med 2013;1:260-8.

57. Yoshida S, Kido J, Matsumoto S, et al. Prenatal diagnosis of Gaucher disease using next-generation sequencing. Pediatr Int 2016;58:946-9.

58. Cognata V, Guarnaccia M, Polizzi A, Ruggieri M, Cavallaro S. Highlights on Genomics Applications for Lysosomal Storage Diseases. Cells 2020;9:1902.

59. O’rourke E, Laney D, Morgan C, Mooney K, Sullivan J. Genetic counseling for lysosomal storage diseases. Lysosomal Storage Disorders. Boston: Springer US; 2007. p. 179-95.

60. Kaback M, Lim-Steele J, Dabholkar D, et al. Tay-sachs disease - carrier screening, prenatal diagnosis, and the molecular era. an international perspective, 1970 to 1993. the international TSD data collection network. JAMA 1993;270:2307-15.

61. Gross SJ, Pletcher BA, Monaghan KG. Professional Practice and Guidelines Committee. Carrier screening in individuals of Ashkenazi Jewish descent. Genet Med 2008;10:54-6.

62. Baskovich B, Hiraki S, Upadhyay K, et al. Expanded genetic screening panel for the Ashkenazi Jewish population. Genet Med 2016;18:522-8.

63. Case LE, Kishnani PS. Physical therapy management of Pompe disease. Genet Med 2006;8:318-27.

64. Lund TC. Hematopoietic stem cell transplant for lysosomal storage diseases. Pediatr Endocrinol Rev 2013;11 Suppl 1:91-8.

65. Aldenhoven M, Wynn RF, Orchard PJ, et al. Long-term outcome of Hurler syndrome patients after hematopoietic cell transplantation: an international multicenter study. Blood 2015;125:2164-72.

66. Prasad VK, Kurtzberg J. Cord blood and bone marrow transplantation in inherited metabolic diseases: scientific basis, current status and future directions. Br J Haematol 2010;148:356-72.

67. Parenti G, Pignata C, Vajro P, Salerno M. New strategies for the treatment of lysosomal storage diseases (review). Int J Mol Med 2013;31:11-20.

68. Muenzer J, Wraith JE, Clarke LA. International Consensus Panel on Management and Treatment of Mucopolysaccharidosis I. Mucopolysaccharidosis I: management and treatment guidelines. Pediatrics 2009;123:19-29.

69. Lum SH, Stepien KM, Ghosh A, et al. Long term survival and cardiopulmonary outcome in children with Hurler syndrome after haematopoietic stem cell transplantation. J Inherit Metab Dis 2017;40:455-60.

70. Prasad VK, Kurtzberg J. Transplant outcomes in mucopolysaccharidoses. Semin Hematol 2010;47:59-69.

71. Mhanni AA, Auray-Blais C, Boutin M, et al. Therapeutic challenges in two adolescent male patients with Fabry disease and high antibody titres. Mol Genet Metab Rep 2020;24:100618.

72. Desai AK, Rosenberg AS, Kishnani PS. The potential impact of timing of IVIG administration on the efficacy of rituximab for immune tolerance induction for patients with Pompe disease. Clin Immunol 2020;219:108541.

73. Kishnani PS, Dickson PI, Muldowney L, et al. Immune response to enzyme replacement therapies in lysosomal storage diseases and the role of immune tolerance induction. Mol Genet Metab 2016;117:66-83.

74. Messinger YH, Mendelsohn NJ, Rhead W, et al. Successful immune tolerance induction to enzyme replacement therapy in CRIM-negative infantile Pompe disease. Genet Med 2012;14:135-42.

75. Lachmann RH. Enzyme replacement therapy for lysosomal storage diseases. Curr Opin Pediatr 2011;23:588-93.

76. Platt FM, Jeyakumar M, Andersson U, Heare T, Dwek RA, Butters TD. Substrate reduction therapy in mouse models of the glycosphingolipidoses. Philos Trans R Soc Lond B Biol Sci 2003;358:947-54.

77. Wenger DA, Coppola S, Liu SL. Insights into the diagnosis and treatment of lysosomal storage diseases. Arch Neurol 2003;60:322-8.

78. Cox TM, Aerts JMFG, Andria G, et al. The role of the iminosugar. N ;26:513-26.

79. Fecarotta S, Romano A, Della Casa R, et al. Long term follow-up to evaluate the efficacy of miglustat treatment in Italian patients with Niemann-Pick disease type C. Orphanet J Rare Dis 2015;10:22.

80. Peterschmitt MJ, Cox GF, Ibrahim J, et al. A pooled analysis of adverse events in 393 adults with Gaucher disease type 1 from four clinical trials of oral eliglustat: Evaluation of frequency, timing, and duration. Blood Cells Mol Dis 2018;68:185-91.

81. Larsen SD, Wilson MW, Abe A, et al. Property-based design of a glucosylceramide synthase inhibitor that reduces glucosylceramide in the brain. J Lipid Res 2012;53:282-91.

82. Marshall J, Ashe KM, Bangari D, et al. Substrate reduction augments the efficacy of enzyme therapy in a mouse model of Fabry disease. PLoS One 2010;5:e15033.

83. Gabig-Cimińska M, Jakóbkiewicz-Banecka J, Malinowska M, et al. Combined Therapies for Lysosomal Storage Diseases. Curr Mol Med 2015;15:746-71.

84. Hinderer C, Bell P, Gurda BL, et al. Intrathecal gene therapy corrects CNS pathology in a feline model of mucopolysaccharidosis I. Mol Ther 2014;22:2018-27.

85. Gilkes JA, Bloom MD, Heldermon CD. Preferred transduction with AAV8 and AAV9 via thalamic administration in the MPS IIIB model: A comparison of four rAAV serotypes. Mol Genet Metab Rep 2016;6:48-54.

86. Liu G, Chen YH, He X, et al. Adeno-associated virus type 5 reduces learning deficits and restores glutamate receptor subunit levels in MPS VII mice CNS. Mol Ther 2007;15:242-7.

87. Lin DS, Hsiao CD, Lee AY, et al. Mitigation of cerebellar neuropathy in globoid cell leukodystrophy mice by AAV-mediated gene therapy. Gene 2015;571:81-90.

88. Miyake N, Miyake K, Asakawa N, Yamamoto M, Shimada T. Long-term correction of biochemical and neurological abnormalities in MLD mice model by neonatal systemic injection of an AAV serotype 9 vector. Gene Ther 2014;21:427-33.

89. Cain JT, Likhite S, White KA, et al. Gene therapy corrects brain and behavioral pathologies in CLN6-batten disease. Mol Ther 2019;27:1836-47.

90. Weismann CM, Ferreira J, Keeler AM, et al. Systemic AAV9 gene transfer in adult GM1 gangliosidosis mice reduces lysosomal storage in CNS and extends lifespan. Hum Mol Genet 2015;24:4353-64.

91. Snyder BR, Gray SJ, Quach ET, et al. Comparison of adeno-associated viral vector serotypes for spinal cord and motor neuron gene delivery. Hum Gene Ther 2011;22:1129-35.

92. Ohno K, Samaranch L, Hadaczek P, et al. Kinetics and MR-based monitoring of AAV9 vector delivery into cerebrospinal fluid of nonhuman primates. Mol Ther Methods Clin Dev 2019;13:47-54.

93. Foust KD, Nurre E, Montgomery CL, et al. Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat Biotechnol 2009;27:59-65.

94. Samaranch L, Salegio EA, San Sebastian W, et al. Adeno-associated virus serotype 9 transduction in the central nervous system of nonhuman primates. Hum Gene Ther 2012;23:382-9.

95. Van Alstyne M, Tattoli I, Delestrée N, et al. Gain of toxic function by long-term AAV9-mediated SMN overexpression in the sensorimotor circuit. Nat Neurosci 2021;24:930-40.

96. Hinderer C, Katz N, Buza EL, et al. Severe toxicity in nonhuman primates and piglets following high-dose intravenous administration of an adeno-associated virus vector expressing human SMN. Hum Gene Ther 2018;29:285-98.

97. Samaranch L, Salegio EA, San Sebastian W, et al. Strong cortical and spinal cord transduction after AAV7 and AAV9 delivery into the cerebrospinal fluid of nonhuman primates. Hum Gene Ther 2013;24:526-32.

98. Hordeaux J, Buza EL, Dyer C, et al. Adeno-associated virus-induced dorsal root ganglion pathology. Human Gene Therapy 2020;31:808-18.

99. Ciesielska A, Mittermeyer G, Hadaczek P, et al. Anterograde axonal transport of AAV2-GDNF in rat basal ganglia. Mol Ther 2011;19:922-7.

100. Bobo RH, Laske DW, Akbasak A, et al. Convection-enhanced delivery of macromolecules in the brain. Proc Natl Acad Sci U S A 1994;91:2076-80.

101. Morrison PF, Chen MY, Chadwick RS, Lonser RR, Oldfield EH. Focal delivery during direct infusion to brain: role of flow rate, catheter diameter, and tissue mechanics. Am J Physiol 1999;277:R1218-29.

102. Gonzalez EA, Baldo G. Gene therapy for lysosomal storage disorders: recent advances and limitations. J. inborn errors metab 2017;5:232640981668978.

103. Baldo G, Giugliani R, Matte U. Gene delivery strategies for the treatment of mucopolysaccharidoses. Expert Opin Drug Deliv 2014;11:449-59.

104. Sevin C, Benraiss A, Van Dam D, et al. Intracerebral adeno-associated virus-mediated gene transfer in rapidly progressive forms of metachromatic leukodystrophy. Hum Mol Genet 2006;15:53-64.

105. Liu G, Martins I, Wemmie JA, Chiorini JA, Davidson BL. Functional correction of CNS phenotypes in a lysosomal storage disease model using adeno-associated virus type 4 vectors. J Neurosci 2005;25:9321-7.

106. Cressant A, Desmaris N, Verot L, et al. Improved behavior and neuropathology in the mouse model of Sanfilippo type IIIB disease after adeno-associated virus-mediated gene transfer in the striatum. J Neurosci 2004;24:10229-39.

107. Skorupa AF, Fisher KJ, Wilson JM, Parente MK, Wolfe JH. Sustained production of beta-glucuronidase from localized sites after AAV vector gene transfer results in widespread distribution of enzyme and reversal of lysosomal storage lesions in a large volume of brain in mucopolysaccharidosis VII mice. Exp Neurol 1999;160:17-27.

108. Brooks AI, Stein CS, Hughes SM, et al. Functional correction of established central nervous system deficits in an animal model of lysosomal storage disease with feline immunodeficiency virus-based vectors. Proc Natl Acad Sci U S A 2002;99:6216-21.

109. Ciron C, Desmaris N, Colle MA, et al. Gene therapy of the brain in the dog model of Hurler’s syndrome. Ann Neurol 2006;60:204-13.

110. Katz ML, Tecedor L, Chen Y, et al. AAV gene transfer delays disease onset in a TPP1-deficient canine model of the late infantile form of Batten disease. Sci Transl Med 2015;7:313ra180.

111. Pastores GM, Torres PA, Zeng BJ. Animal models for lysosomal storage disorders. Biochemistry (Mosc) 2013;78:721-5.

112. Deverman BE, Pravdo PL, Simpson BP, et al. Cre-dependent selection yields AAV variants for widespread gene transfer to the adult brain. Nat Biotechnol 2016;34:204-9.

113. Chan KY, Jang MJ, Yoo BB, et al. Engineered AAVs for efficient noninvasive gene delivery to the central and peripheral nervous systems. Nat Neurosci 2017;20:1172-9.

114. Davidsson M, Wang G, Aldrin-Kirk P, et al. A systematic capsid evolution approach performed in vivo for the design of AAV vectors with tailored properties and tropism. Proc Natl Acad Sci U S A ;2019:27053-62.

115. Liu Q, Segal DJ, Ghiara JB, Barbas CF 3rd. Design of polydactyl zinc-finger proteins for unique addressing within complex genomes. Proc Natl Acad Sci U S A 1997;94:5525-30.

116. Anguela XM, Sharma R, Doyon Y, et al. Robust ZFN-mediated genome editing in adult hemophilic mice. Blood 2013;122:3283-7.

117. Li H, Haurigot V, Doyon Y, et al. In vivo genome editing restores haemostasis in a mouse model of haemophilia. Nature 2011;475:217-21.

118. Sharma R, Anguela XM, Doyon Y, et al. In vivo genome editing of the albumin locus as a platform for protein replacement therapy. Blood 2015;126:1777-84.

119. Laoharawee K, DeKelver RC, Podetz-Pedersen KM, et al. Dose-dependent prevention of metabolic and neurologic disease in murine MPS II by ZFN-mediated in vivo genome editing. Mol Ther 2018;26:1127-36.

120. Ou L, DeKelver RC, Rohde M, et al. ZFN-mediated in vivo genome editing corrects murine hurler syndrome. Mol Ther 2019;27:178-87.

121. Jinek M, Chylinski K, Fonfara I, et al. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012;337:816-21.

122. Broeders M, Herrero-Hernandez P, Ernst MPT, van der Ploeg AT, Pijnappel WWMP. Sharpening the molecular scissors: advances in gene-editing technology. iScience 2020;23:100789.

123. Ou L, Przybilla MJ, Ahlat O, et al. A Highly efficacious ps gene editing system corrects metabolic and neurological complications of mucopolysaccharidosis type I. Mol Ther 2020;28:1442-54.

124. Ou L, Przybilla MJ, Tăbăran AF, et al. A novel gene editing system to treat both Tay-Sachs and Sandhoff diseases. Gene Ther 2020;27:226-36.

125. Levy JM, Yeh WH, Pendse N, et al. Cytosine and adenine base editing of the brain, liver, retina, heart and skeletal muscle of mice via adeno-associated viruses. Nat Biomed Eng 2020;4:97-110.

126. Böckenhoff A, Cramer S, Wölte P, et al. Comparison of five peptide vectors for improved brain delivery of the lysosomal enzyme arylsulfatase A. J Neurosci 2014;34:3122-9.

127. Zhou QH, Boado RJ, Lu JZ, Hui EK, Pardridge WM. Brain-penetrating IgG-iduronate 2-sulfatase fusion protein for the mouse. Drug Metab Dispos 2012;40:329-35.

128. Salvalaio M, Rigon L, Belletti D, et al. Targeted polymeric nanoparticles for brain delivery of high molecular weight molecules in lysosomal storage disorders. PLoS One 2016;11:e0156452.

129. Tancini B, Tosi G, Bortot B, et al. Use of polylactide-co-glycolide-nanoparticles for lysosomal delivery of a therapeutic enzyme in glycogenosis type II fibroblasts. J Nanosci Nanotechnol 2015;15:2657-66.

130. Nguyen QH, Witt RG, Wang B, et al. Tolerance induction and microglial engraftment after fetal therapy without conditioning in mice with Mucopolysaccharidosis type VII. Sci Transl Med 2020;12:eaay8980.

131. Nijagal A, Flake AW, MacKenzie TC. In utero hematopoietic cell transplantation for the treatment of congenital anomalies. Clin Perinatol 2012;39:301-10.

132. Derderian SC, Jeanty C, Walters MC, Vichinsky E, MacKenzie TC. In utero hematopoietic cell transplantation for hemoglobinopathies. Front Pharmacol 2014;5:278.

133. Ito H, Shiwaku H, Yoshida C, et al. In utero gene therapy rescues microcephaly caused by Pqbp1-hypofunction in neural stem progenitor cells. Mol Psychiatry 2015;20:459-71.

134. Massaro G, Mattar CNZ, Wong AMS, et al. Fetal gene therapy for neurodegenerative disease of infants. Nat Med 2018;24:1317-23.

135. Rashnonejad A, Amini Chermahini G, Gündüz C, et al. Fetal gene therapy using a single injection of recombinant AAV9 rescued SMA phenotype in mice. Mol Ther 2019;27:2123-33.

136. Shangaris P, Loukogeorgakis SP, Subramaniam S, et al. In utero gene therapy (IUGT) using GLOBE lentiviral vector phenotypically corrects the heterozygous humanised mouse model and its progress can be monitored using MRI techniques. Sci Rep 2019;9:11592.

137. Bose SK, White BM, Kashyap MV, et al. In utero adenine base editing corrects multi-organ pathology in a lethal lysosomal storage disease. Nat Commun 2021;12:4291.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Munjal V, Clarke MT, Vignolles-Jeong J, Valencia JA, Travis M, Samaranch L. Lysosomal storage disorders with neurological manifestations. Rare Dis Orphan Drugs J 2022;1:6. http://dx.doi.org/10.20517/rdodj.2021.05

AMA Style

Munjal V, Clarke MT, Vignolles-Jeong J, Valencia JA, Travis M, Samaranch L. Lysosomal storage disorders with neurological manifestations. Rare Disease and Orphan Drugs Journal. 2022; 1(2): 6. http://dx.doi.org/10.20517/rdodj.2021.05

Chicago/Turabian Style

Munjal, Vikas, Maria T. Clarke, Joshua Vignolles-Jeong, Jasmine A. Valencia, Meika Travis, Lluis Samaranch. 2022. "Lysosomal storage disorders with neurological manifestations" Rare Disease and Orphan Drugs Journal. 1, no.2: 6. http://dx.doi.org/10.20517/rdodj.2021.05

ACS Style

Munjal, V.; Clarke MT.; Vignolles-Jeong J.; Valencia JA.; Travis M.; Samaranch L. Lysosomal storage disorders with neurological manifestations. Rare. Dis. Orphan. Drugs. J. 2022, 1, 6. http://dx.doi.org/10.20517/rdodj.2021.05

About This Article

© The Author(s) 2022. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
2338
Downloads
1009
Citations
0
Comments
0
2

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 27 clicks
Like This Article 2 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Rare Disease and Orphan Drugs Journal
ISSN 2771-2893 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/