Next Article in Journal
Functional Role of microRNAs in Regulating Cardiomyocyte Death
Previous Article in Journal
T Cells, Interleukin-2 and Systemic Lupus Erythematosus—From Pathophysiology to Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Characterization of Portuguese Patients with Hereditary Cerebellar Ataxia

1
UnIGENe, IBMC-Institute for Molecular and Cell Biology, i3S-Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
2
Neurology Department, Centro Hospitalar Universitário do Porto, 4099-001 Porto, Portugal
3
CGPP-Centre for Predictive and Preventive Genetics, IBMC-Institute for Molecular and Cell Biology, i3S-Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
4
Department of Neurodegenerative Science, Van Andel Institute, Grand Rapids, MI 49503, USA
5
Department of Neurosciences and Mental Health, Neurology, Hospital de Santa Maria, Centro Hospitalar Universitário Lisboa Norte, 1649-028 Lisbon, Portugal
6
Instituto de Medicina Molecular João Lobo Antunes, Faculdade de Medicina, Universidade de Lisboa, 1649-028 Lisbon, Portugal
7
CNS-Campus Neurológico Sénior, 2560-280 Torres Vedras, Portugal
8
ICBAS-School of Medicine and Biomedical Sciences, Universidade do Porto, 4050-313 Porto, Portugal
9
Division of Psychiatry and Behavioral Medicine, Michigan State University College of Human Medicine, Grand Rapids, MI 49503, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2022, 11(6), 981; https://doi.org/10.3390/cells11060981
Submission received: 21 February 2022 / Revised: 10 March 2022 / Accepted: 11 March 2022 / Published: 12 March 2022
(This article belongs to the Section Cells of the Nervous System)

Abstract

:
Hereditary cerebellar ataxia (HCA) comprises a clinical and genetic heterogeneous group of neurodegenerative disorders characterized by incoordination of movement, speech, and unsteady gait. In this study, we performed whole-exome sequencing (WES) in 19 families with HCA and presumed autosomal recessive (AR) inheritance, to identify the causal genes. A phenotypic classification was performed, considering the main clinical syndromes: spastic ataxia, ataxia and neuropathy, ataxia and oculomotor apraxia (AOA), ataxia and dystonia, and ataxia with cognitive impairment. The most frequent causal genes were associated with spastic ataxia (SACS and KIF1C) and with ataxia and neuropathy or AOA (PNKP). We also identified three families with autosomal dominant (AD) forms arising from de novo variants in KIF1A, CACNA1A, or ATP1A3, reinforcing the importance of differential diagnosis (AR vs. AD forms) in families with only one affected member. Moreover, 10 novel causal-variants were identified, and the detrimental effect of two splice-site variants confirmed through functional assays. Finally, by reviewing the molecular mechanisms, we speculated that regulation of cytoskeleton function might be impaired in spastic ataxia, whereas DNA repair is clearly associated with AOA. In conclusion, our study provided a genetic diagnosis for HCA families and proposed common molecular pathways underlying cerebellar neurodegeneration.

Graphical Abstract

1. Introduction

Hereditary cerebellar ataxias (HCA) comprise a heterogeneous group of neurological disorders. The cardinal cerebellar syndrome of HCA may be present in isolation or accompanied by a wide constellation of neurological and non-neurological symptoms and signs. All modes of inheritance have been recognized, with the autosomal recessive (AR) forms being the most complex, both from a clinical and genetic point of view [1]. Prevalence of autosomal dominant (AD) HCA ranges from 0.0 to 5.6/100,000, and that of AR-HCA varies from 0.0 to 7.2/100,000 [2]. AD-HCA are generally caused by repeat expansions, with Machado-Joseph Disease/Spinocerebellar ataxia type 3 (MJD/SCA3) being, overall, the most frequent form and SCA2 the second on a global basis, but first in countries such as Cuba or India [2,3]. Friedreich ataxia is the most frequent recessive ataxia, followed by ataxia telangiectasia or ataxia with oculomotor apraxia (AOA), depending on the series [2,3]. The genetic diversity of HCA has greatly increased over the past years, due to the contribution of next generation sequencing (NGS) techniques. With a diagnostic yield of 20–25%, NGS has allowed for the identification of new ataxia-causing genes and redefining genotype–phenotype correlations; however, it has uncovered a large number of variants of unknown clinical significance [4,5].
Despite the diversity of HCA-genes and their ubiquitous expression throughout the brain, genetic defects in HCA predominantly affect the cerebellum, suggesting that cerebellar cells (particularly, Purkinje cells) are more vulnerable. Therefore, the identification of shared molecular pathways between mutated genes is crucial to understand the basis of cerebellar neurodegeneration [6,7]. Many genes function in common pathways, including protein quality control, transcription regulation, and calcium homeostasis in AD-HCA [8]; and DNA repair, mitochondria function, and lipid metabolism in AR-HCA [9]. Based on common pathways identified, it may be possible to develop effective therapeutic strategies, targeting more than one form of HCA [7].
This study focused on 19 Portuguese families with apparent AR-HCA, previously identified during a large population-based survey and studied by targeted testing for the most common genes [10], and now investigated using whole-exome sequencing (WES). The molecular mechanisms potentially responsible for the neurodegeneration associated with the identified HCA causal-genes were reviewed and discussed.

2. Materials and Methods

2.1. Patients and Clinical Study

A national, systematic, population-based survey was conducted in Portugal from 1994 until 2004, aiming at identifying Portuguese families with HCA and hereditary spastic paraplegia (HSP) [10]. Its detailed methodology was previously described [10]; briefly, clinical and family history, neurological evaluation and blood sampling were performed, and genomic DNA extracted. Written informed consent was obtained from participants.

2.2. Genetic Analysis

In the national survey, 250 families (344 patients) with presumed autosomal recessive ataxia were identified. Targeted genetic testing was conducted ab initium, according to the main clinical phenotype; 61 families (83 patients), however, remained without a genetic diagnosis [10]. Pathogenic variants in PNKP [11] were later identified in eight families (11 patients), and in MAG [12] in one family (3 patients), all with AOA. More recently, expansion of an intronic repeat in RFC1 [13] was excluded in all families with genomic DNA available who persisted without diagnosis.
From the 61 undiagnosed families, we only had genomic DNA available for 27 families, with a final diagnosis being reached in the 19 here described. We performed whole-exome sequencing (WES) in at least one patient of each family (depending on the availability of genomic DNA). Exome sequencing libraries were prepared using a SureSelect Exome Capture Kit v7 (Agilent), and sequencing was performed on NextSeq550 with 100 bp paired-end reads, according to the manufacturer’s instructions. After sequencing, reads were then aligned to the human reference genome hg19/GRCh37 using Burrows-Wheeler Aligner (bwa) v0.7.1 [14] and variants were called using GATK best practices v3.3-0 [15]. Duplicate reads were identified through samblaster v0.1.21 [16]. Variant annotation was performed with snpEff v4.2 [17] and dbNSFP v2.9 [18].
Annotated variants were filtered to include rare variants with a minor allele frequency (MAF) <1% in population databases such as Nucleotide Polymorphism Database (dbSNP; https://www.ncbi.nlm.nih.gov/snp (accessed on 11 February 2022)), Genome Aggregation Database (gnomAD v2.1; https://gnomad.broadinstitute.org (accessed on 11 February 2022)), and the 1000 Genome Project ((http://www.1000genomes.org (accessed on 11 February 2022)). Non-synonymous and splice-site variants, and variants for AR forms were filtered for primary analysis. This analysis was performed using Exomiser v7.2.1 [19], which also allowed variant prioritization based on the human phenotype ontology HP:0001251 (term name:ataxia). Then, we excluded intronic, UTR, intergenic, and synonymous variants, as well as homozygous variants reported in gnomAD. In addition, when de novo variants were suspected, those for AD forms were also filtered and analyzed. The functional predicted impact of missense variants was evaluated using SIFT [20], Mutation Taster [21], Polyphen2 [22], Mutation assessor [23], FATHMM [24], and UMD-Predictor [25]. Analysis of splice-site variants was performed using Splice Site Finder-like [26], MaxEntScan [27], Neural Network SPLICE (NNSPLICE) v0.9 [28], GeneSplicer [29], and Human Splicing Finder (HSF) v3.1 [30]. Variants were classified according to the ACMG/AMP Standards and Guidelines [31]. ClinVar (http://www.ncbi.nlm.nih.gov/clinvar (accessed on 11 February 2022)) and Human Gene Mutation Database (http://www.hgmd.cf.ac.uk/ac/index.php (accessed on 11 February 2022)) were also used for variant classification.
Relevant variants identified by WES were confirmed by Sanger sequencing, which also allowed verifying intrafamilial segregation in several families. For PCR amplifications, we used Ranger Mix (Bioline, London, UK), purified products with Exo/SAP (GRiSP, Porto, Portugal), and performed Sanger sequencing using Big Dye Terminator Cycle Sequencing v1.1 (Applied Biosystems, Foster City, CA, USA) in an ABI 3130xl Genetic Analyzer (Applied Biosystems, Foster City, CA, USA). Sequencing analysis was carried out using the SeqScape v2.6 software (Applied Biosystems, Foster City, CA, USA).

2.3. Minigene Assays

The SPG11 and KIF1C minigenes constructs were obtained by cloning the exonic and intronic sequences flanking the variants of interest (c.3039-5T > G in SPG11 and c.1166-2A > G in KIF1C) into the pCMVdi vector, kindly provided by Dr Alexandra Moreira [32]. Briefly, SPG11 exons 16, 17, and 18 with intronic regions or KIF1C exons 13, 14, and 15 with intronic regions were PCR amplified from patients’ genomic DNA and cloned into the pCMVdi vector, using the Gibson assembly method. Sequences were modified by site-directed mutagenesis to generate the wild-type constructs, using a QuikChange II Kit (Agilent, Santa Clara, CA, USA), according to the manufacturer’s protocol.
HEK293T cells were transfected with the minigene constructs using jetPRIME (Polyplus-transfection, Illkirch, France), according to the manufacturer’s protocol. RNA was extracted 48 h after transfection, using NZYol (Nzytech, Lisbon, Portugal), as per manufacturer’s recommendations, followed by purification of the RNA aqueous phase, using an RNeasy mini kit (Qiagen, Hilden, Germany). RNA quantification was performed on NanoDrop 2000 (ThermoFisher Scientific, Waltham, MA, USA). cDNA was synthesized by reverse transcription-PCR of 2 µg total RNA with oligo (dT), using a SuperScript III first-strand synthesis system (Invitrogen, Carlsbad, CA, USA), according to the manufacturer’s protocol. The resulting cDNA was amplified by PCR and loaded on agarose gel for extraction with a Zymoclean Gel DNA Recovery Kit (Zymo Research, Irvine, CA, USA). The resulting products were sequenced by Sanger sequencing using Big Dye Terminator Cycle Sequencing v1.1 (Applied Biosystems, Foster City, CA, USA) in an ABI 3130xl Genetic Analyzer (Applied Biosystems, Foster City, CA, USA).

3. Results

3.1. Genetic and Molecular Characterization

We used WES to identify variants and genes causing HCA in 19 Portuguese families with family history compatible with AR transmission (including 9 with known consanguinity), in a total of 30 individuals: 19 index cases, one affected and 10 non-affected relatives. At least, one relative of seven index cases was available. All relevant variants (rare, pathogenic, likely pathogenic, and/or predicted to be deleterious) were confirmed by Sanger sequencing, and segregation analysis was performed with all available family members. Phenotypic classification considered the main clinical symptom associated with ataxia: spasticity, neuropathy, oculomotor apraxia, dystonia, or cognitive impairment.
Overall, we identified 24 rare nucleotide variants in 13 genes (Table S1, supporting information): SACS, KIF1C, ANO10, SPG11, SYNE1, and CACNA1A were related to spastic ataxia (10/19 families, 52.6%); KIF1A, POLG, SETX, and PNKP to ataxia and neuropathy (4/19, 21.1%); PNKP to AOA (2/19, 10.5%); HEXB and ATP1A3 to ataxia and dystonia (2/19, 10.5%); and FA2H to ataxia with cognitive impairment (1/19, 5.3%). The most frequent genes were those associated with spastic ataxia, ataxia, and neuropathy or AOA: SACS (4 families, 21.1%), KIF1C (2 families, 10.5%), and PNKP (3 families, 15.8%); the remaining genes were identified in single families only.
Variants were classified as pathogenic when having been previously reported as disease-causing (14/24), or by causing an early stop codon (nonsense and frameshift; 8/24) in genes where loss-of-function is a known disease mechanism (Table S2, supporting information). In addition, two variants were predicted to affect splicing by bioinformatics analysis (Table S3, supporting information)—c.3039-5T > G in SPG11 and c.1166-2A > G in KIF1C—both in families with spastic ataxia; their detrimental effect on splicing was confirmed by minigene splicing-assays, concluding that both are expected to result in frameshifts (Figure 1; Figure S1, supporting information).
We also identified two novel missense variants (Figure 2 and Table S3, supporting information), both in heterozygosity, classified as likely pathogenic (2/24): one in CACNA1A (c.4996C > G; p.Arg1666Gly; family AR267), at an amino-acid residue where a different missense change has been determined to be pathogenic, located in a well-established functional (transmembrane) domain (Table S4, supporting information) and not found in population databases; and one in ATP1A3 (c.374T > A; p.Val125Glu; family AR278), a gene with low rate of benign missense variation, not reported in population databases, located in a mutation hotspot within a transmembrane domain (Table S4, supporting information) and predicted to be deleterious. Moreover, the WT residues of human CACNA1A (Arg1666) and ATP1A3 (Val125) are conserved across species, while the mutated residues are predicted to change their interactions, probably affecting protein structure (Figure 2). Besides the apparent AR inheritance, the variant in ATP1A3 was confirmed to occur de novo (absent in the unaffected parents, family AR78). Unfortunately, we did not have parental samples to test for the CACNA1A variant (family AR267), but this variant probably occurs de novo, as these are a recurrent cause of CACNA1A-related ataxia [33]. In addition, we also identified one de novo variant in KIF1A (c.761G > A, described as pathogenic) in a patient with spastic ataxia, and confirmed the unaffected parents and sibling did not carry it (family AR49).
Thus, we reported 10 novel disease-associated variants in nine families (Table S1, supporting information). All variants were absent from population databases or were present only in heterozygosity with very low MAF.

3.2. Clinical Characterization

Beyond the 20 patients tested, we had clinical information (but no DNA sample) from nine additional affected relatives of probands (detailed data Table 1).

3.2.1. Spastic Ataxia

Spastic ataxia was the largest phenotypic group, comprising 10 families (12 patients). With the exception of AR120 (SACS) and AR267 (CACNA1A), all patients had onset in the first decade of life. Its presenting symptoms comprised ataxia (AR53 (SACS), AR120 (SACS), AR103 (ANO10), AR109 (SYNE1), AR267(CACNA1A)), delayed motor milestones (AR77 (SACS), AR252 (SACS), AR111 (KIF1C)), cognitive regression (AR108 (SPG11)), and upper limb tremors (AR96(KIF1C)); we highlight, due to its rarity, the presence of seizures in AR77 (SACS) and the spastic ataxia phenotype in AR267 (CACNA1A). Regarding imaging, notably, family AR96 (KIF1C) showed basal ganglia hypointensities on T2/FLAIR sequences and AR111 (KIF1C) had brainstem atrophy.

3.2.2. Ataxia and Neuropathy

This group included four families (10 patients). Onset was in the first decade in AR49 (KIF1A) and AR92 (PNKP), second decade in AR4 (SETX), and third decade in AR126 (POLG). Ataxia was the presenting feature in all families, with neuropathy developing later. It is noteworthy that in AR4 (SETX) and AR92 (PNKP) none of the patients had oculomotor apraxia. All patients from AR4 presented diplopia and nystagmus, but none from AR92 displayed such features. The patient from AR49 (KIF1A) had an extensor plantar response in the absence of spasticity. AR126 had a POLG-related ataxia, with the typical optic atrophy, vertical gaze palsy, and epilepsy. One of the members from AR92 died at age 35, with 29 years of disease duration. Imaging was only available in two families (AR4, AR49), all displaying cerebellar atrophy.

3.2.3. Ataxia and Oculomotor Apraxia

Both families (4 patients) had onset in the first decade of life, and disease-causing variants in PNKP. Ataxia was the first symptom in all patients. OMA, neuropathy, and obesity were present in all individuals, and dystonia was highly frequent. Cerebellar atrophy was identified on magnetic resonance imaging (MRI); when tested, total proteins/albumin level was decreased, and cholesterol increased. In AR117, death occurred between 37 and 55 years of age, with a mean disease duration of 37 years.

3.2.4. Ataxia and Dystonia/Ataxia and Cognitive Impairment

AR2 (HEXB) and AR278 (ATP1A3), with one patient each, presented ataxia and dystonia. AR2 (HEXB) had adult-onset ataxia, very prominent oromandibular dystonia and muscle wasting, while AR278 (ATP1A3) had onset at age 11 years, with paroxysmal lower limbs dystonia induced by walking, and progressive ataxia four years later. Over the disease course, dystonia became permanent and generalized. Both AR2 and AR278 had cerebellar atrophy on MRI.
AR16 (FA2H) was the only family with ataxia and cognitive impairment as the major phenotype, in a patient with delayed motor and cognitive milestones, who later developed ataxia and seizures. Cortical and cerebellar atrophy was present on MRI, and increased latency was identified in visual evoked potentials.

4. Discussion

In this study, we performed WES on 30 individuals from 19 families with (apparently) recessive cerebellar ataxia, aiming at providing a conclusive genetic diagnosis for them. These families had been identified during our national population survey in Portugal [10], but remained for many years without a molecular diagnosis, after testing the most common genes. More recently, new causal-genes were identified in undiagnosed families, including PNKP in eight families with AOA4 [11], MAG in one family with AOA and neuropathy [12], and DAB1 in three AD families with SCA37 [34].
In this report, pathogenic and likely-pathogenic variants in several genes were identified and clearly associated with HCA (Table 1 and Table S1, supporting information). Division of the cohort into phenotypic subgroups had been performed during the original survey, and we retained this classification for a better genotype–phenotype characterization, and for clinician guidance in everyday practice. All the causal genes now identified had previously been described within their respective phenotypic subgroup [35,36,37,38]. The most frequent were SACS, KIF1C, and PNKP. To note, Friedreich ataxia, AOA (AOA1 and 2) and L-2-hydroxyglutaric aciduria had been previously screened and found to be the most prevalent types of AR-HCA [10]; KIF1C and PNKP were described only a few years later to be the causative genes for SPAX2 and AOA4 [11,39]. Present data reinforce PNKP-related ataxia as one of the most prevalent AR-HCA in Portugal [11]. Both PNKP and SETX were universally associated with neuropathy, but not with AOA; the reason why AR4 had not been previously tested for SETX. This has been described in the meantime [40]. Oculomotor apraxia, thus, should now be regarded as a clinical sign with high specificity, but less sensitivity for AOA2 and AOA4. Additionally, we identified four novel disease-variants in SACS and two in KIF1C, broadening the genetic spectrum of ARSACS and SPAX2.
We also provided functional data to validate the detrimental effect of two new splice-site variants (Figure 1) that would have been otherwise classified as variants of unknown clinical significance. Using minigene assays, we inferred that SPG11 c.3039-5T > G variant creates a new 3′ acceptor splice site, while KIF1C c.1166-2A > G abolished the canonical 3′ acceptor slice site. Both were predicted to modify the open reading-frame, resulting in a premature stop codon.

4.1. Heterozygous De Novo Variants

All families analyzed had been classified as AR-HCA, based either on the presence of affected relatives in the same generation with unaffected parents; absence of other affected relatives; or consanguinity. Nevertheless, de novo variants in heterozygosity were identified (or presumed) in three families with only one affected member (Figure 2), highlighting that various modes of inheritance should considered when analyzing sequencing data in such cases. We identified one previously reported de novo missense variant in KIF1A [41], and two novel variants in CACNA1A and ATP1A3 were classified as likely pathogenic. KIF1A has been associated with AR and de novo AD diseases [42]. Patients with de novo KIF1A-related variants have shown a complex phenotype, including developmental delay, ataxia, spastic paraplegia, and neuropathy, with onset in the first year of life [41,43,44]. Our patient from family AR49 had had cerebellar ataxia since age 5 years, no spasticity, but an extensor plantar response, and neuropathy. Age of onset and combination of neurological syndromes were similar to those observed in Friedreich ataxia, but this was not the case in other cases previously described [43,44]. Nevertheless, our patient raises the hypothesis that heterozygous variants in KIF1A should be considered in ‘Friedreich ataxia-like’ patients.
Opposed to KIF1A, CACNA1A and ATP1A3 are known to cause only autosomal dominant disorders [45,46,47,48]. CAG repeats in the last exon of the longest isoform of CACNA1A cause SCA6, and missense mutations affecting the calcium channel-encoding sequence cause progressive cerebellar ataxia, familial hemiplegic migraine (FHM1), and episodic ataxia type 2 (EA2) [49]. A congenital form of cerebellar ataxia with cognitive impairment has also been recently described [50]. A spastic ataxic phenotype, as observed in family AR267, has only been described once [35], probably being a less frequent presentation in CACNA1A-related disorders. Interestingly, CACNA1A variants were one of the most prevalent causes of AD-HCA, after SCA3 and dentatorubral-pallidoluysian atrophy (DRPLA), in our population-based survey [10]. Variants in ATP1A3 produce a wide spectrum of AD neurological and psychiatric disorders, ranging from infantile to adult onset. Cerebellar ataxia is classically present in CAPOS syndrome (cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss), but atypical phenotypes with ataxia (either relapsing, of acute onset or slowly progressive) have been increasingly recognized [51]. The proband of AR278 had paroxysmal dystonia induced by walking, resembling exercise-induced dyskinesia (PED), and, later, slowly progressive ataxia and generalized dystonia. A phenotype similar to PED has been reported in a family with four affected individuals, two of which also display cerebellar ataxia [52].

4.2. Molecular Mechanisms

HCA can be caused either by loss-of-function, gain-of function, or a dominant-negative effect, in a multitude of apparently unrelated genes. Generally, AR-HCA is associated with loss-of-function variants, whereas AD-HCA can be caused by a combination of gain and/or loss of function [7]. Our study highlights the diversity of cerebellar ataxia-related genes and phenotypes, possibly reflecting different disease mechanisms (Table S2, supporting information). A pivotal question that remains unanswered is why the Purkinje cells are particularly affected [6,7]. Most genes/proteins identified in our study do not show elevated cerebellar expression, compared to other brain regions (Table S5, supporting information), as per RNA and protein data from the Human Protein Atlas, and in agreement with current literature [7]. We reviewed the molecular mechanisms underlying the main clinical phenotypes (Table S2, supporting information) and discuss potential common disease-pathways in AR-HCA below.

4.2.1. Spastic Ataxia

Spasticity and pyramidal signs are hallmarks of several cerebellar ataxias [53,54]. As demonstrated by the diversity of genes causing spastic ataxia in our study, there is not a particular molecular pathway underlying this phenotype. Nevertheless, we identified one molecular mechanism that may be shared between a set of genes: (1) SACS, KIF1C, SPG11, and SYNE1 interact with the cytoskeleton and may function to ensure proper vesicle/organelle trafficking and/or neuronal structure.
SACS encodes the large protein sacsin, the function of which has not yet been clearly established. The presence of both ubiquitin-like (UBL) and DNaJ domain (implicated in chaperon-mediated folding process), and a potential role in the degradation of aberrant ataxin-1, suggest that sacsin may integrate the ubiquitin–proteasome system and Hsp70 chaperone function [55]. Moreover, studies in SACS knockout mice, revealed that sacsin regulates the neurofilament cytoskeleton and mitochondria dynamics [56,57]. Most SACS variants lead to the complete loss of sacsin, although some missense variants are associated with low levels of the protein [58,59]. Since most variants are upstream of the C-terminal DnaJ domain, it is speculated that ARSACS is associated with loss of chaperone function [60]. Concordantly, we identified known and novel variants in an unspecific region or within the sacsin-repeating region with homology with Hsp90 (Table S4, supporting information); all upstream of the DNaJ domain.
Kinesin family member 1C (KIF1C) is a ubiquitously expressed motor protein involved in both anterograde transport of vesicles and retrograde transport from Golgi to the ER [61]. Few reports have analyzed the impact of KIF1C variants, but it was suggested that variants in the motor domain may affect microtubule binding and impair ATP hydrolysis, while truncating variants lead to reduced protein levels and may impair cargo binding [62]. Therefore, truncating variants, such as the ones identified in our study, are probably associated with a loss-of-function mechanism.
SPG11 encodes spatacsin, a large protein involved with neuronal axonal growth, intracellular cargo trafficking, and lysosome function. SPG11 variants span the entire spatacsin protein [63]. Functional studies of truncating variants indicated that spatacsin loss-of-function is at the basis of neurodegeneration, as evidenced by reduced expression of spatacsin [64]; reduction in the anterograde vesicle trafficking, indicative of impaired axonal transport [65]; reduction in axonal complexity and neurite outgrowth [65]; or the presence of abnormal lysosomes [64]. Given the nature of the variants we identified, it is plausible that they also cause loss of spatacsin function.
SYNE1 is a large gene that encodes the synaptic nuclear envelope protein 1 (SYNE1), also known as nuclear envelope spectrin 1 (nesprin 1). This protein mediates the link from the nuclear envelope to the actin cytoskeleton (LINC complex), being important for nuclear migration [66,67]. Most SYNE1 variants associated with ARCA1 are nonsense or frameshift, and span the giant nesprin-1 isoform, but exclude the C-terminal Klarsicht-ANC-Syne-homology (KASH) [68]. Thus, ARCA1 may be caused by defective synaptic transmission through synaptic vesicles trafficking impairment and/or cerebellar dendrites defects; however, this remains to be confirmed [69]. Nevertheless, many missense variants were also identified in ARCA1, including one identified by us, within a spectrin repeat domain.
In addition, variants in ANO10, which encodes a putative calcium-activated chloride channel, named anoctamin-10 (ANO10), are associated with SCAR10/ARCA3 [70], with decreased coenzyme Q10 (CoQ10) levels. One study showed that truncating variants cause a reduction in ANO10 mRNA expression in patient’s skin fibroblasts [71], but the pathogenic mechanism of ANO10 variants remains unexplored.

4.2.2. Ataxia and Neuropathy

Peripheral neuropathy is a common feature associated to HCAs; but, as far as we know, this combination has not been linked with specific disease pathways. In this study, we found three genes (POLG, SETX, and PNKP) with various molecular functions, causing recessive ataxia and neuropathy.
POLG encodes the catalytic subunit of the mitochondrial DNA polymerase gamma protein (POLG), which functions in the replication and repair of mtDNA [72]. Functional studies showed that missense variants cause decreased activity, DNA binding and processivity of the polymerase, and reduction of functional mtDNA [73,74].The POLG variant identified in our study (p.Trp748Ser) is within the C-terminal domain of the polymerase, an important domain for its activity [75]. We can, therefore, speculate that it causes reduction or loss of POLG function.
SETX and PNKP are discussed in the following sub-section.

4.2.3. Ataxia and Oculomotor Apraxia

AOA has been associated with the DNA damage repair pathway. Ataxia telangiectasia, AOA1, AOA2, AOA4, XRCC1-AOA, and spinocerebellar ataxia with axonal neuropathy 1 are all caused by genes with a role in DNA repair [9]. We identified variants in SETX and PNKP, causing AOA2 and AOA4, respectively. Interestingly, in one family with pathogenic variants in SETX and one in PNKP, oculomotor apraxia was not present. Both had been phenotypically classified as ‘ataxia with neuropathy’, due to severe neuropathy. This possible occurrence has been previously described and is reinforced in this study.
SETX encodes a large nuclear protein termed senataxin, with an RNA/DNA helicase domain. Senataxin is involved in RNA metabolism, DNA maintenance, and damage response. Several types of biallelic variants in SETX were identified in AOA2 [76,77,78], probably causing a loss-of-function phenotype. In addition, dominant missense variants in SETX cause amyotrophic lateral sclerosis (ALS) with juvenile onset (ALS4) [79], suggesting that gain-of-function leads to ALS4. Altered gene expression and mRNA processing, and increased susceptibility to oxidative DNA damage, have all been associated with SETX variants [79,80,81,82]. There have been no functional studies of the SETX p.Gly2047Cys variant [83], but as it lies within the DNA/RNA helicase domain, it may cause a reduction or loss of senataxin function.
Variants in PNKP, encoding the nuclear polynucleotide kinase 3′ phosphatase protein (PNKP), result in a range of AR disorders. The enzymatic activity of PNKP is involved in the repair of both DNA double strand-break (DSB) and single-strand break (SSB); impaired function can cause neurodevelopmental dysfunction (microcephaly, seizures, and developmental delay (MCSZ)) [84] and neurodegeneration (AOA4 and Charcot-Marie-Tooth disease (CMTD)) [11,36]. Few functional studies have described the effects of PKNP variants. Nevertheless, most variants seem to cause reduced stability and levels of PNKP, and reduced DNA phosphatase activity in MCSZ or reduced kinase activity in neurodegenerative diseases. These ultimately lead to reduced DNA repair [85,86]. Moreover, several types of variants were identified in PNKP; most are in the kinase domain, particularly in neurodegenerative diseases [11,87], as shown in this study.

4.2.4. Ataxia and Cognitive Impairment or Dystonia

Additionally, other neurological signs may be present in AR-HCA, such as dystonia, and cognitive impairment. In our study, we identified a known variant in HEXB in one family with Sandhoff disease, and a new variant in FA2H causing ataxia with cognitive impairment.
Sandhoff disease is caused by variants in HEXB encoding the β subunit of the enzyme hexosaminidase (HEXB) [88]. HEXB is a lysosomal hydrolase that catalyzes the degradation of GM2 ganglioside. The spectrum of HEXB variants is wide, but most reduce GM2 hydrolysis, resulting in the accumulation of GM2 ganglioside [89,90]. Moreover, it was reported that several variants, including missense, lead to reduced mRNA expression and affect HEXB protein structure [91,92,93]. Particularly, the p.Arg505Gln variant, located in the catalytic domain, was reported to disturb the biochemical properties and cause loss of activity of HEXB and accumulation of GM2 ganglioside [90,94].
FA2H encodes the fatty acid 2-hydroxylase (FA2H) protein, which is involved in the synthesis of 2-hydroxy fatty acid galactolipids, the most abundant lipids in the myelin sheath [95]. Pathogenic variants in FA2H are speculated to cause neurodegeneration through a loss-of-function mechanism, since there is evidence of decreased hydroxylation of myelin lipids [96]. Moreover, a missense variant causing neurodegeneration with brain iron accumulation, led to reduced FA2H protein expression [97]. Since the FA2H variant we identified is predicted to cause a frameshift, it is also possible it leads to loss-of-function.

5. Conclusions

Our study provided the molecular diagnosis of 19 families with various types of HCA through WES, expanding the genetic and phenotypic spectrum of HCA. Most of the mutated genes caused a spastic ataxia phenotype, but PNKP associated with either ataxia and neuropathy or AOA was also amongst the most frequent causal-genes. We highlight three families with de novo variants in KIF1A, CACNA1A, and ATP1A3, despite having been initially classified has AR. These results evidence the importance of performing a differential diagnosis (AR vs. AD forms) in the absence of affected relatives of the proband. All the remaining families were associated with AR-HCA related genes. In addition, we provided a review on potential common mechanisms underlying neurodegeneration in AR-HCA, namely cytoskeleton function in spastic ataxia, and DNA repair in AOA. Translation of genetic findings into a better understanding of HCA mechanisms may help the development of effective therapies [6,7].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11060981/s1, Figure S1: Electropherogram with the position of the SPG11 (c.3039-5T>G); Table S1: Pathogenic and likely pathogenic variants identified by WES; Table S2: Molecular functions and disease mechanisms associated with recessive cerebellar ataxia genes and encoded proteins, grouped by main clinical phenotype [11,36,38,39,55,56,57,61,64,65,67,69,70,71,73,74,76,77,79,84,85,90,95,96,98,99,100,101,102,103,104,105,106,107,108,109]; Table S3: Pathogenic splice site variants and likely pathogenic missense variants, and pathogenicity predictions; Table S4: Protein domain location of variants identified; Table S5: RNA and protein expression data obtained from the Human Protein Atlas for genes and respective proteins identified in this study [110].

Author Contributions

Conceptualization, M.S., J.D., J.B. (José Brás) and R.G.; methodology, M.S., J.D., S.C., J.L.N., N.D., L.C.G., C.B., J.B. (José Barros), J.B. (José Brás) and R.G.; writing—original draft preparation, M.S. and J.D.; writing—review and editing, M.S., J.D., S.C., J.L.N., N.D., L.C.G., C.B., J.B. (José Barros), J.B. (José Brás), J.S. and R.G.; supervision, J.B. (José Brás), J.S. and R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by National Funds through FCT—Fundação para a Ciência e a Tecnologia, I.P., under the project UIDB/04293/2020. It was also supported in part by the FCT grant FCT-ANR/BEX-GMG/0008/2013; the Porto Neurosciences and Neurologic Disease Research Initiative at the i3S (Norte-01-0145-FEDER-000008), supported by Norte Portugal Regional Operational Programme (NORTE 2020) under the PORTUGAL 2020 Partnership Agreement, also through FEDER; and by GenomePT (POCI-01-0145-FEDER-022184). M.S. was the recipient of a fellowship (SFRH/BPD/116046/2016) and acknowledges funding from FCT through program DL 57/2016—Norma Transitória.

Institutional Review Board Statement

The reanalysis of the families here studied was approved by the Ethics Committee of Centro Hospitalar Universitário do Porto; approval code 2017-069(063-DEFI-063-CES); last approval date 2017.

Informed Consent Statement

All patients and relatives provided informed consent for research and genetic studies [10].

Data Availability Statement

Not applicable.

Acknowledgments

We sincerely thank the patients and relatives for their participation. We also thank Isabel Alonso for all her support in this study. Lastly, we are indebted to Paula Coutinho, who led the systematic population-based survey and collection of blood samples. Her work is an inspiration for us.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anheim, M.; Tranchant, C.; Koenig, M. The autosomal recessive cerebellar ataxias. N. Engl. J. Med. 2012, 366, 636–646. [Google Scholar] [CrossRef] [PubMed]
  2. Ruano, L.; Melo, C.; Silva, M.C.; Coutinho, P. The global epidemiology of hereditary ataxia and spastic paraplegia: A systematic review of prevalence studies. Neuroepidemiology 2014, 42, 174–183. [Google Scholar] [CrossRef] [PubMed]
  3. Sequeiros, J.; Martins, S.; Silveira, I. Epidemiology and population genetics of degenerative ataxias. Handb. Clin. Neurol. 2012, 103, 227–251. [Google Scholar] [CrossRef] [PubMed]
  4. Fogel, B.L.; Lee, H.; Deignan, J.L.; Strom, S.P.; Kantarci, S.; Wang, X.; Quintero-Rivera, F.; Vilain, E.; Grody, W.W.; Perlman, S.; et al. Exome sequencing in the clinical diagnosis of sporadic or familial cerebellar ataxia. JAMA Neurol. 2014, 71, 1237–1246. [Google Scholar] [CrossRef] [PubMed]
  5. van de Warrenburg, B.P.; Schouten, M.I.; de Bot, S.T.; Vermeer, S.; Meijer, R.; Pennings, M.; Gilissen, C.; Willemsen, M.A.; Scheffer, H.; Kamsteeg, E.J. Clinical exome sequencing for cerebellar ataxia and spastic paraplegia uncovers novel gene-disease associations and unanticipated rare disorders. Eur. J. Hum. Genet. 2016, 24, 1460–1466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Robinson, K.J.; Watchon, M.; Laird, A.S. Aberrant Cerebellar Circuitry in the Spinocerebellar Ataxias. Front. Neurosci. 2020, 14, 707. [Google Scholar] [CrossRef] [PubMed]
  7. Eidhof, I.; van de Warrenburg, B.P.; Schenck, A. Integrative network and brain expression analysis reveals mechanistic modules in ataxia. J. Med. Genet. 2019, 56, 283–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Duenas, A.M.; Goold, R.; Giunti, P. Molecular pathogenesis of spinocerebellar ataxias. Brain 2006, 129, 1357–1370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Fogel, B.L. Autosomal-recessive cerebellar ataxias. Handb. Clin. Neurol. 2018, 147, 187–209. [Google Scholar] [CrossRef] [PubMed]
  10. Coutinho, P.; Ruano, L.; Loureiro, J.L.; Cruz, V.T.; Barros, J.; Tuna, A.; Barbot, C.; Guimarães, J.; Alonso, I.; Silveira, I.; et al. Hereditary Ataxia and Spastic Paraplegia in Portugal. JAMA Neurol. 2013, 70, 746. [Google Scholar] [CrossRef] [Green Version]
  11. Bras, J.; Alonso, I.; Barbot, C.; Costa, M.M.; Darwent, L.; Orme, T.; Sequeiros, J.; Hardy, J.; Coutinho, P.; Guerreiro, R. Mutations in PNKP cause recessive ataxia with oculomotor apraxia type 4. Am. J. Hum. Genet 2015, 96, 474–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Santos, M.; Damasio, J.; Kun-Rodrigues, C.; Barbot, C.; Sequeiros, J.; Bras, J.; Alonso, I.; Guerreiro, R. Novel MAG Variant Causes Cerebellar Ataxia with Oculomotor Apraxia: Molecular Basis and Expanded Clinical Phenotype. J. Clin. Med. 2020, 9, 1212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cortese, A.; Simone, R.; Sullivan, R.; Vandrovcova, J.; Tariq, H.; Yau, W.Y.; Humphrey, J.; Jaunmuktane, Z.; Sivakumar, P.; Polke, J.; et al. Biallelic expansion of an intronic repeat in RFC1 is a common cause of late-onset ataxia. Nat. Genet. 2019, 51, 649–658. [Google Scholar] [CrossRef] [PubMed]
  14. Li, H.; Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 2009, 25, 1754–1760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Van der Auwera, G.A.; Carneiro, M.O.; Hartl, C.; Poplin, R.; Del Angel, G.; Levy-Moonshine, A.; Jordan, T.; Shakir, K.; Roazen, D.; Thibault, J.; et al. From FastQ data to high confidence variant calls: The Genome Analysis Toolkit best practices pipeline. Curr. Protoc. Bioinform. 2013, 43, 11.10.11–11.10.33. [Google Scholar] [CrossRef]
  16. Faust, G.G.; Hall, I.M. SAMBLASTER: Fast duplicate marking and structural variant read extraction. Bioinformatics 2014, 30, 2503–2505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Cingolani, P.; Platts, A.; Wang, L.L.; Coon, M.; Nguyen, T.; Wang, L.; Land, S.J.; Lu, X.; Ruden, D.M. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly 2012, 6, 80–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Liu, X.; Jian, X.; Boerwinkle, E. dbNSFP v2.0: A database of human non-synonymous SNVs and their functional predictions and annotations. Hum. Mutat. 2013, 34, E2393–E2402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Smedley, D.; Jacobsen, J.O.; Jager, M.; Kohler, S.; Holtgrewe, M.; Schubach, M.; Siragusa, E.; Zemojtel, T.; Buske, O.J.; Washington, N.L.; et al. Next-generation diagnostics and disease-gene discovery with the Exomiser. Nat. Protoc. 2015, 10, 2004–2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Sim, N.L.; Kumar, P.; Hu, J.; Henikoff, S.; Schneider, G.; Ng, P.C. SIFT web server: Predicting effects of amino acid substitutions on proteins. Nucleic Acids Res. 2012, 40, W452–W457. [Google Scholar] [CrossRef] [PubMed]
  21. Schwarz, J.M.; Cooper, D.N.; Schuelke, M.; Seelow, D. MutationTaster2: Mutation prediction for the deep-sequencing age. Nat. Methods 2014, 11, 361–362. [Google Scholar] [CrossRef] [PubMed]
  22. Adzhubei, I.A.; Schmidt, S.; Peshkin, L.; Ramensky, V.E.; Gerasimova, A.; Bork, P.; Kondrashov, A.S.; Sunyaev, S.R. A method and server for predicting damaging missense mutations. Nat. Methods 2010, 7, 248–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Reva, B.; Antipin, Y.; Sander, C. Predicting the functional impact of protein mutations: Application to cancer genomics. Nucleic Acids Res. 2011, 39, e118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Shihab, H.A.; Gough, J.; Cooper, D.N.; Stenson, P.D.; Barker, G.L.; Edwards, K.J.; Day, I.N.; Gaunt, T.R. Predicting the functional, molecular, and phenotypic consequences of amino acid substitutions using hidden Markov models. Hum. Mutat. 2013, 34, 57–65. [Google Scholar] [CrossRef] [PubMed]
  25. Salgado, D.; Desvignes, J.P.; Rai, G.; Blanchard, A.; Miltgen, M.; Pinard, A.; Levy, N.; Collod-Beroud, G.; Beroud, C. UMD-Predictor: A High-Throughput Sequencing Compliant System for Pathogenicity Prediction of any Human cDNA Substitution. Hum. Mutat. 2016, 37, 439–446. [Google Scholar] [CrossRef] [Green Version]
  26. Shapiro, M.B.; Senapathy, P. RNA splice junctions of different classes of eukaryotes: Sequence statistics and functional implications in gene expression. Nucleic Acids Res. 1987, 15, 7155–7174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Yeo, G.; Burge, C.B. Maximum entropy modeling of short sequence motifs with applications to RNA splicing signals. J. Comput. Biol. 2004, 11, 377–394. [Google Scholar] [CrossRef] [PubMed]
  28. Reese, M.G.; Eeckman, F.H.; Kulp, D.; Haussler, D. Improved splice site detection in Genie. J. Comput. Biol. 1997, 4, 311–323. [Google Scholar] [CrossRef] [PubMed]
  29. Pertea, M.; Lin, X.; Salzberg, S.L. GeneSplicer: A new computational method for splice site prediction. Nucleic Acids Res. 2001, 29, 1185–1190. [Google Scholar] [CrossRef] [Green Version]
  30. Desmet, F.O.; Hamroun, D.; Lalande, M.; Collod-Beroud, G.; Claustres, M.; Beroud, C. Human Splicing Finder: An online bioinformatics tool to predict splicing signals. Nucleic Acids Res. 2009, 37, e67. [Google Scholar] [CrossRef] [Green Version]
  31. Richards, S.; Aziz, N.; Bale, S.; Bick, D.; Das, S.; Gastier-Foster, J.; Grody, W.W.; Hegde, M.; Lyon, E.; Spector, E.; et al. Standards and guidelines for the interpretation of sequence variants: A joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet. Med. 2015, 17, 405–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. da Gloria, V.G.; Martins de Araujo, M.; Mafalda Santos, A.; Leal, R.; de Almeida, S.F.; Carmo, A.M.; Moreira, A. T cell activation regulates CD6 alternative splicing by transcription dynamics and SRSF1. J. Immunol. 2014, 193, 391–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Luo, X.; Rosenfeld, J.A.; Yamamoto, S.; Harel, T.; Zuo, Z.; Hall, M.; Wierenga, K.J.; Pastore, M.T.; Bartholomew, D.; Delgado, M.R.; et al. Clinically severe CACNA1A alleles affect synaptic function and neurodegeneration differentially. PLoS Genet. 2017, 13, e1006905. [Google Scholar] [CrossRef] [PubMed]
  34. Seixas, A.I.; Loureiro, J.R.; Costa, C.; Ordonez-Ugalde, A.; Marcelino, H.; Oliveira, C.L.; Loureiro, J.L.; Dhingra, A.; Brandao, E.; Cruz, V.T.; et al. A Pentanucleotide ATTTC Repeat Insertion in the Non-coding Region of DAB1, Mapping to SCA37, Causes Spinocerebellar Ataxia. Am. J. Hum. Genet. 2017, 101, 87–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Coutelier, M.; Hammer, M.B.; Stevanin, G.; Monin, M.L.; Davoine, C.S.; Mochel, F.; Labauge, P.; Ewenczyk, C.; Ding, J.; Gibbs, J.R.; et al. Efficacy of Exome-Targeted Capture Sequencing to Detect Mutations in Known Cerebellar Ataxia Genes. JAMA Neurol. 2018, 75, 591–599. [Google Scholar] [CrossRef]
  36. Pedroso, J.L.; Rocha, C.R.; Macedo-Souza, L.I.; De Mario, V.; Marques, W., Jr.; Barsottini, O.G.; Bulle Oliveira, A.S.; Menck, C.F.; Kok, F. Mutation in PNKP presenting initially as axonal Charcot-Marie-Tooth disease. Neurol. Genet. 2015, 1, e30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Van Goethem, G.; Martin, J.J.; Dermaut, B.; Löfgren, A.; Wibail, A.; Ververken, D.; Tack, P.; Dehaene, I.; Van Zandijcke, M.; Moonen, M.; et al. Recessive POLG mutations presenting with sensory and ataxic neuropathy in compound heterozygote patients with progressive external ophthalmoplegia. Neuromuscul. Disord. 2003, 13, 133–142. [Google Scholar] [CrossRef]
  38. Synofzik, M.; Smets, K.; Mallaret, M.; Di Bella, D.; Gallenmuller, C.; Baets, J.; Schulze, M.; Magri, S.; Sarto, E.; Mustafa, M.; et al. SYNE1 ataxia is a common recessive ataxia with major non-cerebellar features: A large multi-centre study. Brain 2016, 139, 1378–1393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Dor, T.; Cinnamon, Y.; Raymond, L.; Shaag, A.; Bouslam, N.; Bouhouche, A.; Gaussen, M.; Meyer, V.; Durr, A.; Brice, A.; et al. KIF1C mutations in two families with hereditary spastic paraparesis and cerebellar dysfunction. J. Med. Genet. 2014, 51, 137–142. [Google Scholar] [CrossRef] [PubMed]
  40. Anheim, M.; Monga, B.; Fleury, M.; Charles, P.; Barbot, C.; Salih, M.; Delaunoy, J.P.; Fritsch, M.; Arning, L.; Synofzik, M.; et al. Ataxia with oculomotor apraxia type 2: Clinical, biological and genotype/phenotype correlation study of a cohort of 90 patients. Brain 2009, 132, 2688–2698. [Google Scholar] [CrossRef] [PubMed]
  41. Ohba, C.; Haginoya, K.; Osaka, H.; Kubota, K.; Ishiyama, A.; Hiraide, T.; Komaki, H.; Sasaki, M.; Miyatake, S.; Nakashima, M.; et al. De novo KIF1A mutations cause intellectual deficit, cerebellar atrophy, lower limb spasticity and visual disturbance. J. Hum. Genet. 2015, 60, 739–742. [Google Scholar] [CrossRef] [PubMed]
  42. Morais, S.; Raymond, L.; Mairey, M.; Coutinho, P.; Brandao, E.; Ribeiro, P.; Loureiro, J.L.; Sequeiros, J.; Brice, A.; Alonso, I.; et al. Massive sequencing of 70 genes reveals a myriad of missing genes or mechanisms to be uncovered in hereditary spastic paraplegias. Eur. J. Hum. Genet. 2017, 25, 1217–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Lee, J.R.; Srour, M.; Kim, D.; Hamdan, F.F.; Lim, S.H.; Brunel-Guitton, C.; Decarie, J.C.; Rossignol, E.; Mitchell, G.A.; Schreiber, A.; et al. De novo mutations in the motor domain of KIF1A cause cognitive impairment, spastic paraparesis, axonal neuropathy, and cerebellar atrophy. Hum. Mutat. 2015, 36, 69–78. [Google Scholar] [CrossRef] [PubMed]
  44. Nicita, F.; Ginevrino, M.; Travaglini, L.; D’Arrigo, S.; Zorzi, G.; Borgatti, R.; Terrone, G.; Catteruccia, M.; Vasco, G.; Brankovic, V.; et al. Heterozygous KIF1A variants underlie a wide spectrum of neurodevelopmental and neurodegenerative disorders. J. Med. Genet. 2021, 58, 475–483. [Google Scholar] [CrossRef] [PubMed]
  45. Toru, S.; Murakoshi, T.; Ishikawa, K.; Saegusa, H.; Fujigasaki, H.; Uchihara, T.; Nagayama, S.; Osanai, M.; Mizusawa, H.; Tanabe, T. Spinocerebellar ataxia type 6 mutation alters P-type calcium channel function. J. Biol. Chem. 2000, 275, 10893–10898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Wappl, E.; Koschak, A.; Poteser, M.; Sinnegger, M.J.; Walter, D.; Eberhart, A.; Groschner, K.; Glossmann, H.; Kraus, R.L.; Grabner, M.; et al. Functional consequences of P/Q-type Ca2+ channel Cav2.1 missense mutations associated with episodic ataxia type 2 and progressive ataxia. J. Biol. Chem. 2002, 277, 6960–6966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Heinzen, E.L.; Swoboda, K.J.; Hitomi, Y.; Gurrieri, F.; Nicole, S.; de Vries, B.; Tiziano, F.D.; Fontaine, B.; Walley, N.M.; Heavin, S.; et al. De novo mutations in ATP1A3 cause alternating hemiplegia of childhood. Nat. Genet. 2012, 44, 1030–1034. [Google Scholar] [CrossRef] [Green Version]
  48. Demos, M.K.; van Karnebeek, C.D.; Ross, C.J.; Adam, S.; Shen, Y.; Zhan, S.H.; Shyr, C.; Horvath, G.; Suri, M.; Fryer, A.; et al. A novel recurrent mutation in ATP1A3 causes CAPOS syndrome. Orphanet J. Rare Dis. 2014, 9, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Klockgether, T.; Mariotti, C.; Paulson, H.L. Spinocerebellar ataxia. Nat. Rev. Dis. Primers 2019, 5, 24. [Google Scholar] [CrossRef] [PubMed]
  50. Gur-Hartman, T.; Berkowitz, O.; Yosovich, K.; Roubertie, A.; Zanni, G.; Macaya, A.; Heimer, G.; Duenas, B.P.; Sival, D.A.; Pode-Shakked, B.; et al. Clinical phenotypes of infantile onset CACNA1A-related disorder. Eur. J. Paediatr. Neurol. EJPN Off. J. Eur. Paediatr. Neurol. Soc. 2021, 30, 144–154. [Google Scholar] [CrossRef] [PubMed]
  51. Salles, P.A.; Mata, I.F.; Brunger, T.; Lal, D.; Fernandez, H.H. ATP1A3-Related Disorders: An Ever-Expanding Clinical Spectrum. Front. Neurol. 2021, 12, 637890. [Google Scholar] [CrossRef] [PubMed]
  52. Roubergue, A.; Roze, E.; Vuillaumier-Barrot, S.; Fontenille, M.J.; Meneret, A.; Vidailhet, M.; Fontaine, B.; Doummar, D.; Philibert, B.; Riant, F.; et al. The multiple faces of the ATP1A3-related dystonic movement disorder. Mov. Disord. Off. J. Mov. Disord. Soc. 2013, 28, 1457–1459. [Google Scholar] [CrossRef] [PubMed]
  53. Bereznyakova, O.; Dupre, N. Spastic ataxias. Handb. Clin. Neurol. 2018, 155, 191–203. [Google Scholar] [CrossRef] [PubMed]
  54. de Bot, S.T.; Willemsen, M.A.; Vermeer, S.; Kremer, H.P.; van de Warrenburg, B.P. Reviewing the genetic causes of spastic-ataxias. Neurology 2012, 79, 1507–1514. [Google Scholar] [CrossRef] [PubMed]
  55. Parfitt, D.A.; Michael, G.J.; Vermeulen, E.G.; Prodromou, N.V.; Webb, T.R.; Gallo, J.M.; Cheetham, M.E.; Nicoll, W.S.; Blatch, G.L.; Chapple, J.P. The ataxia protein sacsin is a functional co-chaperone that protects against polyglutamine-expanded ataxin-1. Hum. Mol. Genet. 2009, 18, 1556–1565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Girard, M.; Lariviere, R.; Parfitt, D.A.; Deane, E.C.; Gaudet, R.; Nossova, N.; Blondeau, F.; Prenosil, G.; Vermeulen, E.G.; Duchen, M.R.; et al. Mitochondrial dysfunction and Purkinje cell loss in autosomal recessive spastic ataxia of Charlevoix-Saguenay (ARSACS). Proc. Natl. Acad. Sci. USA 2012, 109, 1661–1666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Lariviere, R.; Gaudet, R.; Gentil, B.J.; Girard, M.; Conte, T.C.; Minotti, S.; Leclerc-Desaulniers, K.; Gehring, K.; McKinney, R.A.; Shoubridge, E.A.; et al. Sacs knockout mice present pathophysiological defects underlying autosomal recessive spastic ataxia of Charlevoix-Saguenay. Hum. Mol. Genet. 2015, 24, 727–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Lariviere, R.; Sgarioto, N.; Marquez, B.T.; Gaudet, R.; Choquet, K.; McKinney, R.A.; Watt, A.J.; Brais, B. Sacs R272C missense homozygous mice develop an ataxia phenotype. Mol. Brain 2019, 12, 19. [Google Scholar] [CrossRef]
  59. Longo, F.; De Ritis, D.; Miluzio, A.; Fraticelli, D.; Baets, J.; Scarlato, M.; Santorelli, F.M.; Biffo, S.; Maltecca, F. Assessment of Sacsin Turnover in Patients With ARSACS: Implications for Molecular Diagnosis and Pathogenesis. Neurology 2021, 97, e2315–e2327. [Google Scholar] [CrossRef] [PubMed]
  60. Bouhlal, Y.; Amouri, R.; El Euch-Fayeche, G.; Hentati, F. Autosomal recessive spastic ataxia of Charlevoix-Saguenay: An overview. Parkinsonism Relat. Disord. 2011, 17, 418–422. [Google Scholar] [CrossRef] [PubMed]
  61. Gabrych, D.R.; Lau, V.Z.; Niwa, S.; Silverman, M.A. Going Too Far Is the Same as Falling Short(dagger): Kinesin-3 Family Members in Hereditary Spastic Paraplegia. Front. Cell Neurosci. 2019, 13, 419. [Google Scholar] [CrossRef] [PubMed]
  62. Caballero Oteyza, A.; Battaloglu, E.; Ocek, L.; Lindig, T.; Reichbauer, J.; Rebelo, A.P.; Gonzalez, M.A.; Zorlu, Y.; Ozes, B.; Timmann, D.; et al. Motor protein mutations cause a new form of hereditary spastic paraplegia. Neurology 2014, 82, 2007–2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Manole, A.; Chelban, V.; Haridy, N.A.; Hamed, S.A.; Berardo, A.; Reilly, M.M.; Houlden, H. Severe axonal neuropathy is a late manifestation of SPG11. J. Neurol. 2016, 263, 2278–2286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Renvoise, B.; Chang, J.; Singh, R.; Yonekawa, S.; FitzGibbon, E.J.; Mankodi, A.; Vanderver, A.; Schindler, A.; Toro, C.; Gahl, W.A.; et al. Lysosomal abnormalities in hereditary spastic paraplegia types SPG15 and SPG11. Ann. Clin. Transl. Neurol. 2014, 1, 379–389. [Google Scholar] [CrossRef] [PubMed]
  65. Perez-Branguli, F.; Mishra, H.K.; Prots, I.; Havlicek, S.; Kohl, Z.; Saul, D.; Rummel, C.; Dorca-Arevalo, J.; Regensburger, M.; Graef, D.; et al. Dysfunction of spatacsin leads to axonal pathology in SPG11-linked hereditary spastic paraplegia. Hum. Mol. Genet. 2014, 23, 4859–4874. [Google Scholar] [CrossRef]
  66. Apel, E.D.; Lewis, R.M.; Grady, R.M.; Sanes, J.R. Syne-1, a dystrophin- and Klarsicht-related protein associated with synaptic nuclei at the neuromuscular junction. J. Biol. Chem. 2000, 275, 31986–31995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Gros-Louis, F.; Dupre, N.; Dion, P.; Fox, M.A.; Laurent, S.; Verreault, S.; Sanes, J.R.; Bouchard, J.P.; Rouleau, G.A. Mutations in SYNE1 lead to a newly discovered form of autosomal recessive cerebellar ataxia. Nat. Genet. 2007, 39, 80–85. [Google Scholar] [CrossRef] [PubMed]
  68. Yoshinaga, T.; Nakamura, K.; Ishikawa, M.; Yamaguchi, T.; Takano, K.; Wakui, K.; Kosho, T.; Yoshida, K.; Fukushima, Y.; Sekijima, Y. A novel frameshift mutation of SYNE1 in a Japanese family with autosomal recessive cerebellar ataxia type 8. Hum. Genome Var. 2017, 4, 17052. [Google Scholar] [CrossRef] [Green Version]
  69. Razafsky, D.; Hodzic, D. A variant of Nesprin1 giant devoid of KASH domain underlies the molecular etiology of autosomal recessive cerebellar ataxia type I. Neurobiol. Dis. 2015, 78, 57–67. [Google Scholar] [CrossRef] [Green Version]
  70. Vermeer, S.; Hoischen, A.; Meijer, R.P.; Gilissen, C.; Neveling, K.; Wieskamp, N.; de Brouwer, A.; Koenig, M.; Anheim, M.; Assoum, M.; et al. Targeted next-generation sequencing of a 12.5 Mb homozygous region reveals ANO10 mutations in patients with autosomal-recessive cerebellar ataxia. Am. J. Hum. Genet. 2010, 87, 813–819. [Google Scholar] [CrossRef] [Green Version]
  71. Balreira, A.; Boczonadi, V.; Barca, E.; Pyle, A.; Bansagi, B.; Appleton, M.; Graham, C.; Hargreaves, I.P.; Rasic, V.M.; Lochmuller, H.; et al. ANO10 mutations cause ataxia and coenzyme Q(1)(0) deficiency. J. Neurol. 2014, 261, 2192–2198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Lestienne, P. Evidence for a direct role of the DNA polymerase gamma in the replication of the human mitochondrial DNA in vitro. Biochem. Biophys. Res. Commun. 1987, 146, 1146–1153. [Google Scholar] [CrossRef]
  73. Luoma, P.T.; Luo, N.; Loscher, W.N.; Farr, C.L.; Horvath, R.; Wanschitz, J.; Kiechl, S.; Kaguni, L.S.; Suomalainen, A. Functional defects due to spacer-region mutations of human mitochondrial DNA polymerase in a family with an ataxia-myopathy syndrome. Hum. Mol. Genet. 2005, 14, 1907–1920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Schaller, A.; Hahn, D.; Jackson, C.B.; Kern, I.; Chardot, C.; Belli, D.C.; Gallati, S.; Nuoffer, J.M. Molecular and biochemical characterisation of a novel mutation in POLG associated with Alpers syndrome. BMC Neurol. 2011, 11, 4. [Google Scholar] [CrossRef]
  75. Van Goethem, G.; Luoma, P.; Rantamaki, M.; Al Memar, A.; Kaakkola, S.; Hackman, P.; Krahe, R.; Lofgren, A.; Martin, J.J.; De Jonghe, P.; et al. POLG mutations in neurodegenerative disorders with ataxia but no muscle involvement. Neurology 2004, 63, 1251–1257. [Google Scholar] [CrossRef] [PubMed]
  76. Moreira, M.C.; Klur, S.; Watanabe, M.; Nemeth, A.H.; Le Ber, I.; Moniz, J.C.; Tranchant, C.; Aubourg, P.; Tazir, M.; Schols, L.; et al. Senataxin, the ortholog of a yeast RNA helicase, is mutant in ataxia-ocular apraxia 2. Nat. Genet. 2004, 36, 225–227. [Google Scholar] [CrossRef]
  77. Airoldi, G.; Guidarelli, A.; Cantoni, O.; Panzeri, C.; Vantaggiato, C.; Bonato, S.; Grazia D’Angelo, M.; Falcone, S.; De Palma, C.; Tonelli, A.; et al. Characterization of two novel SETX mutations in AOA2 patients reveals aspects of the pathophysiological role of senataxin. Neurogenetics 2010, 11, 91–100. [Google Scholar] [CrossRef]
  78. Bernard, V.; Minnerop, M.; Burk, K.; Kreuz, F.; Gillessen-Kaesbach, G.; Zuhlke, C. Exon deletions and intragenic insertions are not rare in ataxia with oculomotor apraxia 2. BMC Med. Genet. 2009, 10, 87. [Google Scholar] [CrossRef] [Green Version]
  79. Chen, Y.Z.; Bennett, C.L.; Huynh, H.M.; Blair, I.P.; Puls, I.; Irobi, J.; Dierick, I.; Abel, A.; Kennerson, M.L.; Rabin, B.A.; et al. DNA/RNA helicase gene mutations in a form of juvenile amyotrophic lateral sclerosis (ALS4). Am. J. Hum. Genet. 2004, 74, 1128–1135. [Google Scholar] [CrossRef] [Green Version]
  80. Chen, Y.Z.; Hashemi, S.H.; Anderson, S.K.; Huang, Y.; Moreira, M.C.; Lynch, D.R.; Glass, I.A.; Chance, P.F.; Bennett, C.L. Senataxin, the yeast Sen1p orthologue: Characterization of a unique protein in which recessive mutations cause ataxia and dominant mutations cause motor neuron disease. Neurobiol. Dis. 2006, 23, 97–108. [Google Scholar] [CrossRef]
  81. Fogel, B.L.; Perlman, S. Novel mutations in the senataxin DNA/RNA helicase domain in ataxia with oculomotor apraxia 2. Neurology 2006, 67, 2083–2084. [Google Scholar] [CrossRef] [PubMed]
  82. Roda, R.H.; Rinaldi, C.; Singh, R.; Schindler, A.B.; Blackstone, C. Ataxia with oculomotor apraxia type 2 fibroblasts exhibit increased susceptibility to oxidative DNA damage. J. Clin. Neurosci. 2014, 21, 1627–1631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Sawyer, S.L.; Schwartzentruber, J.; Beaulieu, C.L.; Dyment, D.; Smith, A.; Warman Chardon, J.; Yoon, G.; Rouleau, G.A.; Suchowersky, O.; Siu, V.; et al. Exome sequencing as a diagnostic tool for pediatric-onset ataxia. Hum. Mutat. 2014, 35, 45–49. [Google Scholar] [CrossRef] [PubMed]
  84. Shen, J.; Gilmore, E.C.; Marshall, C.A.; Haddadin, M.; Reynolds, J.J.; Eyaid, W.; Bodell, A.; Barry, B.; Gleason, D.; Allen, K.; et al. Mutations in PNKP cause microcephaly, seizures and defects in DNA repair. Nat. Genet. 2010, 42, 245–249. [Google Scholar] [CrossRef] [PubMed]
  85. Reynolds, J.J.; Walker, A.K.; Gilmore, E.C.; Walsh, C.A.; Caldecott, K.W. Impact of PNKP mutations associated with microcephaly, seizures and developmental delay on enzyme activity and DNA strand break repair. Nucleic Acids Res. 2012, 40, 6608–6619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Kalasova, I.; Hailstone, R.; Bublitz, J.; Bogantes, J.; Hofmann, W.; Leal, A.; Hanzlikova, H.; Caldecott, K.W. Pathological mutations in PNKP trigger defects in DNA single-strand break repair but not DNA double-strand break repair. Nucleic Acids Res. 2020, 48, 6672–6684. [Google Scholar] [CrossRef] [PubMed]
  87. Gatti, M.; Magri, S.; Nanetti, L.; Sarto, E.; Di Bella, D.; Salsano, E.; Pantaleoni, C.; Mariotti, C.; Taroni, F. From congenital microcephaly to adult onset cerebellar ataxia: Distinct and overlapping phenotypes in patients with PNKP gene mutations. Am J. Med. Genet. A 2019, 179, 2277–2283. [Google Scholar] [CrossRef] [PubMed]
  88. Bolhuis, P.A.; Ponne, N.J.; Bikker, H.; Baas, F.; Vianney de Jong, J.M. Molecular basis of an adult form of Sandhoff disease: Substitution of glutamine for arginine at position 505 of the beta-chain of beta-hexosaminidase results in a labile enzyme. Biochim. Biophys. Acta 1993, 1182, 142–146. [Google Scholar] [CrossRef]
  89. Banerjee, P.; Siciliano, L.; Oliveri, D.; McCabe, N.R.; Boyers, M.J.; Horwitz, A.L.; Li, S.C.; Dawson, G. Molecular basis of an adult form of beta-hexosaminidase B deficiency with motor neuron disease. Biochem. Biophys. Res. Commun. 1991, 181, 108–115. [Google Scholar] [CrossRef]
  90. Redonnet-Vernhet, I.; Mahuran, D.J.; Salvayre, R.; Dubas, F.; Levade, T. Significance of two point mutations present in each HEXB allele of patients with adult GM2 gangliosidosis (Sandhoff disease) homozygosity for the Ile207-->Val substitution is not associated with a clinical or biochemical phenotype. Biochim. Biophys. Acta 1996, 1317, 127–133. [Google Scholar] [CrossRef] [Green Version]
  91. Zampieri, S.; Filocamo, M.; Buratti, E.; Stroppiano, M.; Vlahovicek, K.; Rosso, N.; Bignulin, E.; Regis, S.; Carnevale, F.; Bembi, B.; et al. Molecular and functional analysis of the HEXB gene in Italian patients affected with Sandhoff disease: Identification of six novel alleles. Neurogenetics 2009, 10, 49–58. [Google Scholar] [CrossRef] [PubMed]
  92. Santoro, M.; Modoni, A.; Sabatelli, M.; Madia, F.; Piemonte, F.; Tozzi, G.; Ricci, E.; Tonali, P.A.; Silvestri, G. Chronic GM2 gangliosidosis type Sandhoff associated with a novel missense HEXB gene mutation causing a double pathogenic effect. Mol. Genet. Metab. 2007, 91, 111–114. [Google Scholar] [CrossRef] [PubMed]
  93. Zampieri, S.; Cattarossi, S.; Oller Ramirez, A.M.; Rosano, C.; Lourenco, C.M.; Passon, N.; Moroni, I.; Uziel, G.; Pettinari, A.; Stanzial, F.; et al. Sequence and copy number analyses of HEXB gene in patients affected by Sandhoff disease: Functional characterization of 9 novel sequence variants. PLoS ONE 2012, 7, e41516. [Google Scholar] [CrossRef] [PubMed]
  94. Yasui, N.; Takaoka, Y.; Nishio, H.; Nurputra, D.K.; Sekiguchi, K.; Hamaguchi, H.; Kowa, H.; Maeda, E.; Sugano, A.; Miura, K.; et al. Molecular pathology of Sandhoff disease with p.Arg505Gln in HEXB: Application of simulation analysis. J. Hum. Genet. 2013, 58, 611–617. [Google Scholar] [CrossRef] [PubMed]
  95. Edvardson, S.; Hama, H.; Shaag, A.; Gomori, J.M.; Berger, I.; Soffer, D.; Korman, S.H.; Taustein, I.; Saada, A.; Elpeleg, O. Mutations in the fatty acid 2-hydroxylase gene are associated with leukodystrophy with spastic paraparesis and dystonia. Am. J. Hum. Genet. 2008, 83, 643–648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Dick, K.J.; Eckhardt, M.; Paisan-Ruiz, C.; Alshehhi, A.A.; Proukakis, C.; Sibtain, N.A.; Maier, H.; Sharifi, R.; Patton, M.A.; Bashir, W.; et al. Mutation of FA2H underlies a complicated form of hereditary spastic paraplegia (SPG35). Hum. Mutat. 2010, 31, E1251–E1260. [Google Scholar] [CrossRef] [PubMed]
  97. Kruer, M.C.; Paisan-Ruiz, C.; Boddaert, N.; Yoon, M.Y.; Hama, H.; Gregory, A.; Malandrini, A.; Woltjer, R.L.; Munnich, A.; Gobin, S.; et al. Defective FA2H leads to a novel form of neurodegeneration with brain iron accumulation (NBIA). Ann. Neurol. 2010, 68, 611–618. [Google Scholar] [CrossRef] [PubMed]
  98. Engert, J.C.; Berube, P.; Mercier, J.; Dore, C.; Lepage, P.; Ge, B.; Bouchard, J.P.; Mathieu, J.; Melancon, S.B.; Schalling, M.; et al. ARSACS, a spastic ataxia common in northeastern Quebec, is caused by mutations in a new gene encoding an 11.5-kb ORF. Nat. Genet. 2000, 24, 120–125. [Google Scholar] [CrossRef]
  99. Alfares, A.; Alfadhel, M.; Wani, T.; Alsahli, S.; Alluhaydan, I.; Al Mutairi, F.; Alothaim, A.; Albalwi, M.; Al Subaie, L.; Alturki, S.; et al. A multicenter clinical exome study in unselected cohorts from a consanguineous population of Saudi Arabia demonstrated a high diagnostic yield. Mol. Genet. Metab. 2017, 121, 91–95. [Google Scholar] [CrossRef]
  100. Stevanin, G.; Azzedine, H.; Denora, P.; Boukhris, A.; Tazir, M.; Lossos, A.; Rosa, A.L.; Lerer, I.; Hamri, A.; Alegria, P.; et al. Mutations in SPG11 are frequent in autosomal recessive spastic paraplegia with thin corpus callosum, cognitive decline and lower motor neuron degeneration. Brain 2008, 131, 772–784. [Google Scholar] [CrossRef] [Green Version]
  101. Orlacchio, A.; Babalini, C.; Borreca, A.; Patrono, C.; Massa, R.; Basaran, S.; Munhoz, R.P.; Rogaeva, E.A.; St George-Hyslop, P.H.; Bernardi, G.; et al. SPATACSIN mutations cause autosomal recessive juvenile amyotrophic lateral sclerosis. Brain 2010, 133, 591–598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Montecchiani, C.; Pedace, L.; Lo Giudice, T.; Casella, A.; Mearini, M.; Gaudiello, F.; Pedroso, J.L.; Terracciano, C.; Caltagirone, C.; Massa, R.; et al. ALS5/SPG11/KIAA1840 mutations cause autosomal recessive axonal Charcot-Marie-Tooth disease. Brain 2016, 139, 73–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Zhang, Q.; Bethmann, C.; Worth, N.F.; Davies, J.D.; Wasner, C.; Feuer, A.; Ragnauth, C.D.; Yi, Q.; Mellad, J.A.; Warren, D.T.; et al. Nesprin-1 and -2 are involved in the pathogenesis of Emery Dreifuss muscular dystrophy and are critical for nuclear envelope integrity. Hum. Mol. Genet. 2007, 16, 2816–2833. [Google Scholar] [CrossRef] [PubMed]
  104. Van Goethem, G.; Dermaut, B.; Lofgren, A.; Martin, J.J.; Van Broeckhoven, C. Mutation of POLG is associated with progressive external ophthalmoplegia characterized by mtDNA deletions. Nat. Genet 2001, 28, 211–212. [Google Scholar] [CrossRef]
  105. Naviaux, R.K.; Nguyen, K.V. POLG mutations associated with Alpers’ syndrome and mitochondrial DNA depletion. Ann. Neurol. 2004, 55, 706–712. [Google Scholar] [CrossRef]
  106. Winterthun, S.; Ferrari, G.; He, L.; Taylor, R.W.; Zeviani, M.; Turnbull, D.M.; Engelsen, B.A.; Moen, G.; Bindoff, L.A. Autosomal recessive mitochondrial ataxic syndrome due to mitochondrial polymerase gamma mutations. Neurology 2005, 64, 1204–1208. [Google Scholar] [CrossRef]
  107. Vissing, J.; Ravn, K.; Danielsen, E.R.; Duno, M.; Wibrand, F.; Wevers, R.A.; Schwartz, M. Multiple mtDNA deletions with features of MNGIE. Neurology 2002, 59, 926–929. [Google Scholar] [CrossRef]
  108. Fogel, B.L.; Cho, E.; Wahnich, A.; Gao, F.; Becherel, O.J.; Wang, X.; Fike, F.; Chen, L.; Criscuolo, C.; De Michele, G.; et al. Mutation of senataxin alters disease-specific transcriptional networks in patients with ataxia with oculomotor apraxia type 2. Hum. Mol. Genet. 2014, 23, 4758–4769. [Google Scholar] [CrossRef] [Green Version]
  109. Bolhuis, P.A.; Oonk, J.G.; Kamp, P.E.; Ris, A.J.; Michalski, J.C.; Overdijk, B.; Reuser, A.J. Ganglioside storage, hexosaminidase lability, and urinary oligosaccharides in adult Sandhoff’s disease. Neurology 1987, 37, 75–81. [Google Scholar] [CrossRef]
  110. Uhlen, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, A.; Kampf, C.; Sjostedt, E.; Asplund, A.; et al. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 1260419. [Google Scholar] [CrossRef]
Figure 1. Analysis of SPG11 (c.3039-5T > G) (A) and KIF1C (c.1166-2A > G) (B) splice-site variants, with minigene splicing assays. Agarose gel electrophoresis of the transcripts generated by the wild-type (WT) and mutant minigene constructs (on the left): B1 bands (<500 bp) represent transcripts affected by the variants; B2 bands (>500 bp) represent amplified transcripts with intron retention, not affected by the variant. Sanger sequencing of the transcripts corresponding to B1 bands (in the middle). Schematic representation of resulting splicing events (on the right): (A) SPG11 variant creates a new 3′ acceptor splice site, resulting in the retention of 4 nucleotides of intron 16; (B) KIF1C variant abolished the 3′ acceptor splice-site, leading to the usage of a splice site within exon 14 and resulting in the deletion of 8 nucleotides from exon 14. Electropherograms showing the location of the splice-site variants on genomic DNA are shown in Figure S1, supporting information.
Figure 1. Analysis of SPG11 (c.3039-5T > G) (A) and KIF1C (c.1166-2A > G) (B) splice-site variants, with minigene splicing assays. Agarose gel electrophoresis of the transcripts generated by the wild-type (WT) and mutant minigene constructs (on the left): B1 bands (<500 bp) represent transcripts affected by the variants; B2 bands (>500 bp) represent amplified transcripts with intron retention, not affected by the variant. Sanger sequencing of the transcripts corresponding to B1 bands (in the middle). Schematic representation of resulting splicing events (on the right): (A) SPG11 variant creates a new 3′ acceptor splice site, resulting in the retention of 4 nucleotides of intron 16; (B) KIF1C variant abolished the 3′ acceptor splice-site, leading to the usage of a splice site within exon 14 and resulting in the deletion of 8 nucleotides from exon 14. Electropherograms showing the location of the splice-site variants on genomic DNA are shown in Figure S1, supporting information.
Cells 11 00981 g001
Figure 2. Pedigrees of families harboring de novo variants (A), and conservation of the new missense variants and protein models of CACNA1A and ATP1A3 (B). (A) Black symbols represent individuals with cerebellar ataxia. (B) Sequence alignment of the residues surrounding the mutated residues in CACNA1A and ATP1A3 against other species was performed using the Clustal Omega program. Protein models of CACNA1A and ATP1A3 showing altered residues interactions upon amino acid changes (performed using DynaMut2). The Arg1666 and Val135 residues in CACNA1A and ATP1A3, respectively, are in the center, displaying the various interactions with nearby residues in different colors.
Figure 2. Pedigrees of families harboring de novo variants (A), and conservation of the new missense variants and protein models of CACNA1A and ATP1A3 (B). (A) Black symbols represent individuals with cerebellar ataxia. (B) Sequence alignment of the residues surrounding the mutated residues in CACNA1A and ATP1A3 against other species was performed using the Clustal Omega program. Protein models of CACNA1A and ATP1A3 showing altered residues interactions upon amino acid changes (performed using DynaMut2). The Arg1666 and Val135 residues in CACNA1A and ATP1A3, respectively, are in the center, displaying the various interactions with nearby residues in different colors.
Cells 11 00981 g002
Table 1. Clinical characterization of the families.
Table 1. Clinical characterization of the families.
FamilyInd. Code (Gender)TestedGeneDisease Causing VariantsConsanguinityAgeFirst SymptomAtaxiaOMADiplopiaPyramidal SignsPeripheral NeuropathyDystoniaCognitive ImpairmentOther Symptoms/SignsImagingOthers
OnsetObs.WheelchairDeath
Spastic ataxia
AR53IV.1 (F)YesSACSc.8148del; c.8148delYes65347NAAtaxia YesNoNoYesNoNoNo NAv
AR77II.3 (M)YesSACSc.1373C >T; c. 5440_5449delNo<131NANADelayed motor milestonesYesNoNoYesYesNoNoSeizures 1-16y NystagmusMRI: cerebellar atrophy
AR120II.1 (F)YesSACSc.3066del; c.2656_2666delNo156629NAAtaxiaYesNoNoYesYesNoNoNystagmusNAv
AR252II.1 (M)YesSACSc.814C > T; c.814C > TNot clear<15754NADelayed motor milestonesYesNoNoYesYesNoNoKyphoscoliosisNAv
AR96IV.1 (M)YesKIF1Cc.393_396del;
c.393_396del
Yes95241NAUL tremorYesNoNoYesYesNoMild deteriorationNystagmusMRI: pallidum, putamen, thalamus hypointensities (T2/FLAIR)
IV.2 (F)Yes85044NAUL tremorYes NoNoYesYesNoMild deteriorationNystagmusMRI: vermis atrophy
AR111III.3 (F)YesKIF1Cc.1166-2A > G; c.1166-2A > GYes<141NANADelayed motor milestonesYesNoNoYesYesNoNoVertical gaze restriction
Mild dysmorphism
Pes cavus
CT: brainstem atrophy
AR103III.3 (F)YesANO10c.132dup; c.132dupYes740NANAAtaxiaYesNoYesYesNoNoNoNystagmus
Migraine
MRI: cerebellar atrophy
AR108II.3 (F)YesSPG11c.1951C > T; c.3039-5T > GNo743NANACognitive regressionYesNoNoYesNoNoYes MRI: Thin corpus callosum
II.2 (M)No847NANACognitive regressionYesNoNoYesNoNoYes MRI: Chiari type1
AR109III.2 (M)YesSYNE1c.17270C > G; c.17270C > GNot clearNAv67NANAAtaxiaYesNoNoYesNoNoNo NAv
AR267III.4 (M)YesCACNA1Ac.4996C > GNo1921NANAAtaxiaYesNoNoYesNoNoYes MRI: cerebellar atrophy
Ataxia and neuropathy
AR4VI.27 (M)YesSETXc.1514G > A; c.1514G > AYes1437NANAAtaxiaYesNo *YesNoYesNo NystagmusCT: cerebellar atrophy
VI.7 (F)No1349NANAAtaxiaYesNo *YesNoYesNo NystagmusCT: cerebellar atrophy
VI.10 (F)No1143NANAAtaxiaYesNo *YesNoYesNo NystagmusCT: cerebellar atrophy
VI.19 (F)No1243NANAAtaxiaYesNo *YesNoYesNo Nystagmus
Obesity
CT: cerebellar atrophy
VI.20 (F)No1342NANAAtaxiaYesNo *YesNoYesNoMild retardationNystagmusCT: cerebellar atrophy
VI.31 (F)No1132NANAAtaxiaYesNo *YesNoYesNo NystagmusCT: cerebellar atrophy
AR49II.3 (F)YesKIF1Ac.761G > ANo526NANAAtaxiaYesNoNAvYesYesNoMild retardation CT: cortical cerebellar atrophy
AR92II.1 (M)YesPNKPc.1123G > T; c.1123G > T Not clear6331635AtaxiaYesNoNoNoYesNoNo NAv
II.4 (M)No72716NAAtaxiaYesNoNoNoYesNoNo NAvαFP: N
AR126V.7 (F)YesPOLGc.2243G > C;
c.2243G > C
Yes2541NANAAtaxiaYesNoNoNoYesNoYesOptic atrophy
Vertical gaze palsy
Nystagmus
Epilepsy
NAv
Ataxia and oculomotor apraxia
AR86II.2 (F)YesPNKPc.1123G > T;
c.1253_1269dup
No817NANAAtaxiaYesYesNoNoYesNoYesNystagmus
Obesity
MRI: cerebellar atrophy (++vermis)
AR117II.1 (M)YesPNKPc.1221_1223del; c.1123G > TYes8321637AtaxiaYesYesNAvNoYesYes (UL)NoObesityNAv
II.2 (F)No7351542AtaxiaYesYesNAvNoYesYes (UL)NoObesityMRI: cerebellar atrophyαFP: ↑
Proteins: ↓
Albumin: ↓
Cholesterol: ↑
II.3 (F)No7291555AtaxiaYesYesNAvNoYesYes (UL)NoObesityMRI: cerebellar atrophyProteins: ↓
Albumin: ↓
Cholesterol: ↑
Ataxia and dystonia
AR2IV.1 (M)YesHEXBc.1514G > A; c.1514G > AYes2451NANAAtaxiaYesNoNoYesNoYes (OM)NoNystagmus
Muscle atrophy
↓ vibration hallux
MRI: cerebellar, atrophy
AR278III.1 (M)YesATP1A3c.374T > ANo1144NANAParox dystoniaYesNoNoNoNoYes YesSupranuclear vertical gaze palsyMRI: cerebellar atrophy
Ataxia and cognitive impairment
AR16V.1 (F)YesFA2Hc.619_620del;
c.619_620del
Yes<114NANADelayed motor/cognitive milestonesYesNoNoYesNoNoYesSeizures,
Aggressive behavior
MRI: cerebellar atrophy, mild cortical atrophyVEP: ↑
BEP, SSEP: N
αFP—α feto-protein; BEP—Brainstem evoked potentials; CT—Computed tomography; F—Female; Ind—Individual; M—Male; MRI—Magnetic resonance imaging; NA—Not applicable; Nav—Not available; Obs—Observation; OM—Oromandibular; OMA—Oculomotor apraxia; Parox—paroxysmal; UL—Upper limbs; SSEP—Somatosensory evoked potentials; VEP—Visual evoked potentials; ↓—decreased; ↑—increased * No clear oculomotor apraxia, but all had a poverty of spontaneous saccadic eye movements.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Santos, M.; Damásio, J.; Carmona, S.; Neto, J.L.; Dehghani, N.; Guedes, L.C.; Barbot, C.; Barros, J.; Brás, J.; Sequeiros, J.; et al. Molecular Characterization of Portuguese Patients with Hereditary Cerebellar Ataxia. Cells 2022, 11, 981. https://doi.org/10.3390/cells11060981

AMA Style

Santos M, Damásio J, Carmona S, Neto JL, Dehghani N, Guedes LC, Barbot C, Barros J, Brás J, Sequeiros J, et al. Molecular Characterization of Portuguese Patients with Hereditary Cerebellar Ataxia. Cells. 2022; 11(6):981. https://doi.org/10.3390/cells11060981

Chicago/Turabian Style

Santos, Mariana, Joana Damásio, Susana Carmona, João Luís Neto, Nadia Dehghani, Leonor Correia Guedes, Clara Barbot, José Barros, José Brás, Jorge Sequeiros, and et al. 2022. "Molecular Characterization of Portuguese Patients with Hereditary Cerebellar Ataxia" Cells 11, no. 6: 981. https://doi.org/10.3390/cells11060981

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop