Next Article in Journal
Efficacy and Microbiota Modulation Induced by LimpiAL 2.5%, a New Medical Device for the Inverse Psoriasis Treatment
Next Article in Special Issue
Beyond Strains: Molecular Diversity in Alpha-Synuclein at the Center of Disease Heterogeneity
Previous Article in Journal
Transcriptomic Analysis from Normal Glucose Tolerance to T2D of Obese Individuals Using Bioinformatic Tools
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Evidence for Endolysosomal Dysfunction in Parkinson’s Disease: A Critical Overview

1
Dino Ferrari Center, Department of Pathophysiology and Transplantation, University of Milan, 20122 Milan, Italy
2
Fondazione IRCCS Ca’ Granda Ospedale Maggiore Policlinico, Neurology Unit, 20122 Milan, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6338; https://doi.org/10.3390/ijms24076338
Submission received: 10 January 2023 / Revised: 23 March 2023 / Accepted: 26 March 2023 / Published: 28 March 2023

Abstract

:
Parkinson’s disease (PD) is the second most common neurodegenerative disorder in the aging population, and no disease-modifying therapy has been approved to date. The pathogenesis of PD has been related to many dysfunctional cellular mechanisms, however, most of its monogenic forms are caused by pathogenic variants in genes involved in endolysosomal function (LRRK2, VPS35, VPS13C, and ATP13A2) and synaptic vesicle trafficking (SNCA, RAB39B, SYNJ1, and DNAJC6). Moreover, an extensive search for PD risk variants revealed strong risk variants in several lysosomal genes (e.g., GBA1, SMPD1, TMEM175, and SCARB2) highlighting the key role of lysosomal dysfunction in PD pathogenesis. Furthermore, large genetic studies revealed that PD status is associated with the overall “lysosomal genetic burden”, namely the cumulative effect of strong and weak risk variants affecting lysosomal genes. In this context, understanding the complex mechanisms of impaired vesicular trafficking and dysfunctional endolysosomes in dopaminergic neurons of PD patients is a fundamental step to identifying precise therapeutic targets and developing effective drugs to modify the neurodegenerative process in PD.

1. Introduction

Parkinson’s disease (PD) is the second most common neurodegenerative disorder in the aging population [1,2,3,4]. It is clinically defined by the presence of bradykinesia in combination with either rest tremor and/or rigidity, and a clear beneficial response to dopaminergic therapy [5]. Neuropathologically, it is characterized by the loss of dopaminergic neurons in the substantia nigra (SN) and the presence of α-synuclein positive inclusions (Lewy bodies, LB) in surviving neurons [6,7,8]. At present, there are no approved treatments capable of slowing neurodegeneration in PD. Therefore, it is of paramount importance to shed light on the molecular mechanism causing PD neurodegeneration, because this knowledge is the indispensable prerequisite to identifying therapeutic compounds that can address the dysfunctional cellular machinery specific to this neurodegenerative disorder [9,10]. In the past two decades, PD etiopathogenesis has been linked with several deranged cellular mechanisms, ranging from mitochondrial impairment (PRKN, PINK1, PARK7) and ubiquitination defects (FBXO7) to dysfunction of the endolysosomal pathway (LRRK2, VPS35, VPS13C, ATP13A2) and synaptic vesicle trafficking (SNCA, RAB39B, SYNJ1, DNAJC6). In addition, significant parts of the risk genes associated with PD encode for endolysosomal and synaptic vesicle proteins, confirming a particular susceptibility of PD-related brain structures to the impairment of these pathways (Figure 1) [11,12,13,14,15,16,17,18].
Lysosomes are membrane-bound cytoplasmic organelles equipped with acid hydrolases whose major function consists in preserving cellular homeostasis by breaking down cellular (organized in autophagosomes) and extracellular (imported with endosomes) macromolecules and organelles into their fundamental components [19,20]. Lysosomal storage diseases (LSDs) are inborn metabolism defects often caused by loss-of-function (LOF) mutations in genes encoding lysosomal hydrolases. The consequent aberrant buildup of waste material causes lysosomal impairment, which can ultimately result in cellular dysfunction. Neurodegeneration is observed in association with several LSDs, clearly indicating that lysosomal impairment (or the toxic effect of mutant hydrolases) plays a relevant role in neuronal dysfunction and death [18,21]. Neurons display unique biological characteristics that may explain their increased susceptibility to the dysfunction of lysosomes and endosomal trafficking, such as their arborization and long projections, their post-mitotic state, and their highly complex interactions with other cells [22,23,24,25].
The scope of this review is to provide a critical overview of the genetic evidence supporting the central role of the endolysosomal pathway, that PD researchers can adopt as a solid starting point for future studies. We pursue this goal by reviewing monogenic PD forms and “strong” genetic PD risk variants associated with dysfunction of the endolysosomal pathway, then explore the rapidly evolving field of polygenic “weak” genetic risk variants affecting lysosomal genes, and finally present a unifying perspective linking this genetic evidence with the pathogenesis of PD [11,12,13,14,15,16,17].

2. Monogenic Causes of PD Associated with Endolysosomal and Vesicular Dysfunction

Early epidemiological studies revealed that 10–15% of PD patients have positive family history of the disease while most cases are sporadic [26]. Through family-based linkage analysis and, more recently, next-generation sequencing approaches, thirteen genes have been definitively implicated in the etiology of PD. Mutations in the SNCA [27,28], LRRK2 [29,30,31], and VPS35 [32,33] genes cause autosomal dominant forms, whereas mutations in the PRKN [34], PARK7 [35], and PINK1 [36] genes cause autosomal recessive forms. In addition, biallelic mutations in the ATP13A2 [37], PLA2G6 [38], FBXO7 [39,40], DNAJC6 [41], SYNJ1 [42], and VPS13C [43] have been reported as rare causes of early-onset parkinsonism with atypical clinical features. Finally, RAB39B gene mutations have been associated with a form of X-linked levodopa-responsive parkinsonism in combination with various degrees of intellectual disability [44]. The possibility of relying on these established genetic PD forms represents a solid starting point to build hypotheses and generate genetic disease models for exploring the dysfunctional mechanisms underlying the PD pathology. In the following paragraphs, we focus on those Mendelian genes that support the role of impaired endolysosomal and vesicular function in the pathogenesis of PD (Figure 1).

2.1. LRRK2 (Leucine-Rich Repeat Kinase 2)

Mutations in the LRRK2 gene represent the most common monogenic cause of PD, being accountable for around 3% of all cases [45]. The causality link between LRRK2 mutations and autosomal dominant PD was unraveled in the early 2000s through linkage analysis and positional cloning [29,30,46,47] and was confirmed by many independent studies [31,48,49,50,51,52,53,54,55,56,57,58,59]. Although many candidate pathogenic variants were reported, pathogenicity was proven only for 8 of them, affecting the enzymatic domains of the Lrrk2 protein, including N1437H, R1441G/C/H/S, Y1699C, G2019S, and I2020T [60,61,62,63,64,65,66,67]. In addition, some other more common LRRK2 variants were associated with increased risk of PD (e.g., A419V, R1628P, M1646T, and G2385R) [65,66,67].
The G2019S variant, the most common LRRK2 pathogenic mutation in the western world, has a prevalence of 1% in all PD cases worldwide and is extremely frequent in Berber Arabs (up to 37% of PD patients) and Ashkenazi Jews (up to 23% of PD patients) [47,67,68,69,70,71]. The R1441C variant, the second most common LRRK2 mutation, is more frequent in Italy [72]. The R1441G is a founder mutation in the Basque population (up to 46% in familial PD Basque patients) [73]. The R1628P and the G2385R risk variants are diffused in the Chinese population [74,75,76].
The clinical features of LRRK2-related PD are difficult to distinguish from those of idiopathic PD. From a motor perspective, it is characterized by bradykinesia, rigidity, and rest tremor, with unilateral tremor being the most common onset symptom [59,62,77]. While dystonia seems to be more frequent, cognitive decline, psychiatric symptoms, dysautonomia, and hyposmia are apparently under-represented compared with idiopathic forms [51,59,60,61,65]. Response to levodopa is optimal, though levodopa-induced dyskinesias (LID) are frequent [47,61]. Patients with LRRK2-related PD are probably more susceptible to inflammatory bowel diseases, strokes, and certain cancers (e.g., leukemia, myeloproliferative diseases, colon cancer, breast cancer) [65,78,79,80,81].
Neuropathologically, the majority of LRRK2-PD brains show α-synuclein-positive Lewy pathology [82], but some LRRK2-PD cases lack LB and others show different neuropathological features, such as tau pathology. Remarkably, a review of 55 cases of LRRK2-PD revealed that about half had tau pathology [83].
Lrrk2 is a large protein, with multiple domains and diverse functions in different cells and tissues. The two catalytic domains, GTPase and kinase, have been extensively investigated, with the caveat that most of the studies rely on overexpression of the protein. Lrrk2 was shown to phosphorylate the vesicular Rab GTPases Rab8A and Rab10 [14,84,85,86,87,88].
The role of LRRK2 in PD has been thoroughly investigated in vitro (induced pluripotent stem cell, iPSC) and in vivo (C. elegans, Drosophila, and rodent models). Despite these models being able to identify several impaired mechanisms associated with Lrrk2 dysfunction, the precise link of mutated Lrrk2 to PD has not yet been fully elucidated [89,90,91,92]. Moreover, direct assessment of Lrrk2 endogenous subcellular localization is technically quite challenging, and the overexpression of the protein, fused with a tag in several studies, may be different from the physiological condition. Additionally, Lrrk2 subcellular localizations may change in different cell types and tissues, and under different conditions. Taking these premises into account, the localization of Lrrk2 in early and late endosomes has been repeatedly reported in several human cellular and animal models, supporting its involvement in endosomal trafficking and autophagy-lysosomal pathways [14,84,85,88,93]. In particular, Lrrk2 knockout murine models, which have no brain abnormalities, show peculiar abnormalities such as enlarged lamellar bodies (lysosome-related organelles) in lung cells and enlarged lysosomes with lipofuscin accumulation in kidneys, suggesting an important function of Lrrk2 in lysosomal homeostasis [94]. A direct interaction between Lrrk2 and α-synuclein has not been consistently demonstrated, and the murine models expressing mutant Lrrk2 do not recapitulate the most significant phenotypic features of the disease [95]. Pathogenic LRRK2 mutations are deemed to be gain-of-function (GOF) variants that increase the kinase activity and consequently increase Rab8A and Rab10 phosphorylation, resulting in dysregulation of vesicular transport and mitophagy [14,66,85,88,96]. Interestingly, Rab29 protein (encoded by RAB29, a gene proposed as a risk locus for PD) was shown to play a role in the recruitment of Lrrk2 to stressed lysosomes [97,98]. Finally, recent evidence has revealed that mutated Lrrk2 alters ceramide metabolism and acts as a modifier of glucocerebrosidase level and enzymatic activity, further reinforcing its link with the endolysosomal pathway [99,100,101]. With all the caveats of the proposed models, the studies performed to date indicate a possible role of endolysosomal dysfunction in LRRK2-PD, which requires further validation in more appropriate models, such as iPSCs-derived neurons or organoids from patients carrying LRRK2 mutations.

2.2. SNCA (Alpha-Synuclein)

SNCA was the first gene ever to be causally associated with PD. SNCA mutations were linked with PD in 1997, through linkage analysis in an Italian family (i.e., “Contursi kindred”). Shortly thereafter, α-synuclein was identified as a main component of LB [27,102,103,104]. To date, autosomal dominant PD was associated with SNCA duplications, triplications, and missense mutations including A30G/P, E46K, H50Q (debated role), G51D, and A53E/T/V [27,28,67,105,106,107,108,109,110,111,112,113,114,115].
Despite its great historical importance for understanding PD pathogenesis, SNCA-related parkinsonism is a very rare cause of disease, being accountable for up to 1–2% of familial and 0.2% of sporadic PD cases [67,112]. Clinically, it is characterized by earlier onset: around 43–48 years for A53T and duplications, around 31–36 years for triplications [67,112,116]. All classical motor features including bradykinesia, rigidity, tremor, postural instability, and gait disturbances are common. Dystonia has been reported to be more common in H50Q cases, myoclonus in A53E/T cases, pyramidal signs in G51D cases, and cerebellar symptoms in A30P cases. Additional atypical motor features include anterocollis, retrocollis, and alien limb syndrome. Among non-motor features, cognitive impairment is frequent for all mutations and psychosis is frequent for multiplications. Other overrepresented features are depression, anxiety, sleep disturbances, orthostatic hypotension, and urinary dysautonomia, while constipation and olfactory impairment are relatively rare [14,112,113]. Besides the typical SN degeneration, neuropathological studies showed several phenotypes including cortical Lewy pathology, diffuse Lewy body disease (LBD), α-synuclein glial cytoplasmic inclusions (GCIs), and CA2/3 hippocampus neuronal loss. Levodopa response is good, though motor fluctuation (MF) and LID are common [14,112,113].
The SNCA gene encodes α-synuclein, a protein with a poorly understood physiological function expressed in neurons and possibly oligodendrocytes, involved in the development and recycling of synaptic vesicles and plasticity of dopaminergic neurons [14,112,117,118,119,120,121]. SNCA mutations have been studied both with iPSC and in animal models, deepening the insight into its expression and aggregation but failing to reproduce neurodegeneration and motor symptoms in vivo [89,90].
Pathological α-synuclein tends to aggregate in variable size structures, from dimers to oligomers up to inclusion bodies, known as Lewy bodies in neurons, the pathological hallmark of PD [104,112,122,123,124]. Pathological α-synuclein aggregates are thought to cause neurodegeneration through endoplasmic reticulum (ER) and mitochondrial stress, protein misfolding and degradation, and synaptic and axonal dysfunction, impairing the retrograde transport of endosomes and autophagosomes to lysosomes located in the neuronal body [123,125,126,127,128,129,130,131]. Multiplications of SNCA increase the expression of α-synuclein and consequently its tendency to form pathological aggregates, while, curiously, all identified missense variants affect the N-terminal region of α-synuclein, possibly impacting its ability to bind at cellular membranes [112,132,133]. It is relevant to note that α-synuclein in physiological conditions is mainly degraded by the autophagosome–lysosome pathway. Therefore, a disruption of this pathway may cause α-synuclein aggregation and accumulation in neurons [13]. Supporting this hypothesis, PD-causing mutations such as A30P and A53T are probably associated with an impairment of α-synuclein degradation through autophagy [134,135] (Figure 1).

2.3. VPS35 (Vacuolar Protein Sorting 35 Ortholog)

In 2011, two independent studies using exome sequencing and linkage analysis revealed the link between the D620N mutation of VPS35 and autosomal dominant PD [32,33]. Since then, many variants have been associated with PD (R32S, G51S, M57I, I241M, P316S, R524W, I560T, M57I, H599R, M607V, L774M); however, their pathogenicity is still unconfirmed [136,137,138].
The clinical features of VPS35-related PD are similar to those of typical PD, with a relatively younger age of onset (46.9 ± 8.6 years). The reported motor features are usually classical (bradykinesia, rigidity, rest tremor) with asymmetric onset. Regarding non-motor features, dysautonomia, psychosis, and hallucinations are rare, and cognitive decline is observed in 15–30% of cases. The therapeutic response to dopaminergic therapy is good, though wearing-off and levodopa-induced dyskinesias appear in around 80% of cases [138,139,140,141]. Neuropathological findings remain to be determined [138,139,140,141].
The VPS35 gene encodes a subunit of the heteropentameric retromer complex, located at the endosomal membrane, where it facilitates the endosome-to-Golgi and the endosome-to-plasma membrane transport [138,142,143]. Rodent- and iPSC-based studies are providing greater insight into VPS35-related neurodegeneration [90,92,138,144,145]. The D620N variant was found to be associated with impaired autophagy, possibly due to abnormal sorting of the ATG9A autophagy receptor and decreased autophagosome formation [138,144]. Furthermore, it determines a defective sorting of cathepsin D (encoded by CTSD, see below), resulting in lysosomal dysfunction, which potentially implies an impaired degradation of aggregation-prone proteins such as α-synuclein [138,145]. A third mechanism may be identified in mitochondrial degradation, since Vps35 physiologically facilitates mitophagy via mitochondrial-derived vesicles and the mutated Vps35 might enhance the activity of Dlp1, a protein involved in mitochondrial fission, thus resulting in mitochondrial fragmentation [138,146,147]. Vps35 D620N mutation was shown to enhance Lrrk2-mediated phosphorylation of Rab10 as well as autophosphorylation, suggesting that Vps35 may be an upstream regulator of Lrrk2. Overexpression of wildtype Vps35 was demonstrated in flies and murine models to rescue retromer-mediated defects, such as lysosomal enlargement, caused by Lrrk2 G2019S overexpression or Rab29 knockdown. These observations support a possible shared mechanism of Vps35, Lrrk2, and Rab29 in PD pathogenesis [138,148]. Importantly, the demonstration of a common shared altered mechanism of these proteins involving vesicular and endolysosomal trafficking will require further validation in human-based neuronal models hopefully capable of recapitulating PD pathology. In this direction, recent studies of human iPSC-derived neurons carrying the VPS35 D620N mutation showed decreased autophagic flux [149].

2.4. VPS13C (Vacuolar Protein Sorting 13 Homolog C)

Biallelic VPS13C mutations cause autosomal recessive early-onset PD (EOPD) [43,150]. The link between VPS13C mutations and PD was identified through a genome-wide association study (GWAS) in 2014 and confirmed in 2016 through homozygosity mapping and exome sequencing study of consanguineous families, when the homozygous c.8445 + 2T > G variant was identified in a Turkish patient with PD [43,151].
To our knowledge, only 18 VPS13C-related PD cases have been clinically and genetically described to date, with earlier onset (37.5 ± 10.5 years of age) and faster clinical progression compared to idiopathic PD [152]. From a motor perspective, VPS13C-related PD is generally characterized by classical features such as bradykinesia, rigidity, rest tremor, freezing, and postural instability, as well as by distinct features including dystonia and, more rarely, pyramidal signs. Concerning non-motor features, dysautonomia, hyposmia, early cognitive decline, and visual hallucinations are presented. Dopaminergic therapy is usually effective, though MF and LID are common [43,150,152,153,154,155,156,157]. Neuropathological studies have shown diffuse LBD [43,150,152,153,154,155,156,157].
The gene VPS13C encodes an intermembrane lipid transfer protein localized at contact sites between the ER and late endosomes/lysosomes and on the outer mitochondrial membrane [43,158]. Vps13C regulates lysosomal homeostasis and controls mitophagy, modulating the Pink1/Parkin pathway in cellular models [43,159]. The neurodegeneration associated with the loss of VPS13C function thus seems primarily attributable to an alteration of lysosomal homeostasis and an upregulation of Pink1/Parkin-dependent mitophagy [43,153,154,159]. However, whether this impaired mechanism observed in vitro reflects the pathological process in VPS13C-PD brains is far from being clear yet.

2.5. ATP13A2 (ATPase Cation Transporting 13A2)

Biallelic ATP13A2 mutations have been associated with a multitude of phenotypes including Kufor–Rakeb disease (KRD), neuronal ceroid lipofuscinosis, hereditary spastic paraplegia, and an amyotrophic lateral sclerosis-like form [37,160,161]. KRD is a rare AR, levodopa-responsive, rigid-akinetic parkinsonism with atypical features including pyramidal signs, supranuclear gaze palsy, and cognitive decline. It was first described in 1994 in a Jordanian family, and was associated with ATP13A2 through linkage analysis in 2006 [37,162,163].
Around 50 cases of KRD have been reported to date. The age of onset is usually in the second decade of life. Additional clinical features may include tremor, facial-faucial-finger mini myoclonus, oculogyric dystonic spasms, slow saccadic eye movement, peripheral neuropathy, and visual hallucinations [163,164,165,166,167,168,169,170,171,172,173]. Brain MRI may display generalized brain atrophy with bilateral putaminal and caudate iron accumulation [165,174,175,176]. After an initial response, MF and LID rapidly occur [163,164,177,178]. While the KRD phenotype is generally associated with truncating ATP13A2 mutations, missense mutations in ATP13A2 tend to be associated with a clinical form similar to EOPD [177,179,180,181,182]. A single neuropathologic case of KRD was reported, showing widespread neuronal and glial lipofuscin accumulation with no LB-type inclusions and absence of α-synuclein-positive pathology [183].
ATP13A2 encodes a lysosomal P-type ATPase which has been associated with the homeostasis of metal cations (e.g., Fe3+, Mn2+, Zn2+), mitochondrial clearance, and, possibly, α-synuclein detoxification [171,184,185]. In cell lines, the endogenous level of Atp13a2 repeatedly appeared below immunodetection, despite using different currently available Atp13a2 antibodies. Therefore, most of the studies showing co-localization with lysosomal and late endosomal/intraluminal vesicle markers are based on overexpression of the protein. Atp13a2-deficient mice show sensorimotor deficits, and accumulation of insoluble α-synuclein in the brain, which is exacerbated by overexpression of the human wildtype α-synuclein [186]. Apoptosis dysregulation, mitochondrial dysfunction, and ER stress have emerged from animal and cellular models. Moreover, Atp13a2 LOF determines lysosomal dysfunction with defective polyamine export and autophagosome dysfunction, as effectively explored both in vivo and in vitro [176,187,188,189,190,191].

2.6. RAB39B (RAB39B, Member RAS Oncogene Family)

LOF mutations in RAB39B were identified as a cause of X-linked recessive (XLR) early-onset parkinsonism (EOP) in 2014 through linkage analysis in three Australian brothers presenting intellectual disability (ID) with EOP, and confirmed via linkage analysis and direct RAB39B sequencing in a Wisconsin family with 13 males presenting a similar phenotype [44].
The clinical features of RAB39B-related disease range from classical PD to syndromic forms with various combinations of EOP (akinetic-rigid or with tremor), dystonia, ID (delayed speech, learning difficulties), autistic spectrum disorders, seizures, frontal lobe reflexes, macrocephaly, strabismus, and short stature. Brain magnetic resonance imaging (MRI) may display SN and globus pallidus iron deposition, while brain computerized tomography (CT) may show symmetrical globus pallidus calcification. Response to levodopa is often present, though saddled with MF and LID [44,192,193,194,195,196,197,198,199,200,201,202,203]. Neuropathological findings include SN and cortical neuronal death and LBD, SN tau-positive neurofibrillary tangles, and basal ganglia axonal spheroids [44,195,196,198,202,203,204]. It is to be noted that RAB39B-related parkinsonism is caused by large-scale deletions, splicing abnormalities, and frameshift, nonsense, and missense mutations resulting in LOF. Duplications and triplications of RAB39B may result in a pathologic GOF leading to complex syndromes with ID and behavioral abnormalities [200,203,204,205,206,207,208].
RAB39B encodes a neuronal GTPase involved in vesicular trafficking and recycling between synaptic terminals, endosomes, and the Golgi apparatus. Rab39B is thought to control many cellular functions including α-synuclein homeostasis and GluA2/GluA3 AMPAR subunit trafficking from the ER to the Golgi apparatus, as suggested by several murine, iPSC, and isogenic human embryonic stem cell models [44,193,194,195,200,202,209,210].

2.7. SYNJ1 (Synaptojanin 1)

Biallelic SYNJ1 mutations were linked with autosomal recessive EOPD in 2013 through homozygosity mapping and exome sequencing, in two parallel studies describing an Iranian and an Italian family [42,211].
To date, 34 patients originating from 20 families with biallelic SYNJ1 missense, nonsense, frameshift, or splicing mutations have been reported to constitute a heterogeneous multitude of phenotypes ranging from EOPD or atypical parkinsonism (13 families) up to severe epileptic encephalopathies without parkinsonism (seven families) [42,211,212,213,214,215,216,217,218,219,220,221,222,223,224].
SYNJ1-related parkinsonism is generally characterized by disease onset in the third decade of life. Classical features include bradykinesia, rest tremor, and postural instability. A variety of atypical features have been reported including vertical gaze palsy, eyelid apraxia, diplopia, dysarthria/anarthria, hypophonia, dysphagia, drooling, oromandibular tremor, dystonia, pyramidal signs, cognitive decline, and seizures. In some cases, neuroimaging was reportedly normal, while in other cases it displayed variable abnormalities including cortical atrophy/hypometabolism, caudate hypometabolism, and bilateral nigrostriatal dopaminergic denervation. Levodopa response was variable and a multitude of adverse effects including dyskinesia, dystonia, and postural hypotension have been reported. Among them, five families with various ethnicities carried the R258Q missense variant, which is therefore thought to be a possible mutational hotspot [42,211,212,213,214,215,216,217,218,219,220,221,222,223,224]. A single autoptic examination displayed neuronal loss in SN without LBD, neurofibrillary degeneration, and tau protein staining in cell bodies and axonal hillocks [212].
SYNJ1 encodes an inositol phosphatase that allows the shedding of clathrin coats and other endocytic factors from their membranes in the Golgi apparatus, endosomes, and plasma membrane, playing a fundamental role in synaptic vesicle endocytosis and autophagy [42,225,226,227,228,229]. The pathogenesis of SYNJ1-related parkinsonism has been studied in zebrafish, Drosophila, and rodent models [92,228,230,231,232]. Loss of SYNJ1 function causes synaptic autophagy and transmission defects manifesting with delayed synaptic vesicle endocytic recycling and accumulation of clathrin-coated vesicles [228,229,231,232].

2.8. DNAJC6 (DnaJ Heat Shock Protein Family Member C6)

In 2012, biallelic DNAJC6 mutations were linked with juvenile-onset autosomal recessive parkinsonism through homozygosity mapping and exome sequencing in a Palestinian family [41].
To date, 21 cases of DNAJC6-related parkinsonism have been reported, carrying biallelic missense, nonsense, frameshift, or splicing mutations [41,233,234,235,236,237,238,239]. DNAJC6-related parkinsonism is typically characterized by disease onset in the second decade of life and rapid clinical progression until loss of walking within around 10 years from onset. Initial symptoms tend to be rest tremor and bradykinesia, followed by rigidity and postural instability. Although classic PD is a possible form of DNAJC6-related parkinsonism, there usually are atypical manifestations such as dystonia, pyramidal signs, postural tremor, dysarthria, anarthria, epilepsy, ID, and psychosis. Levodopa response is extremely variable and often burdened by dyskinesias. No consistent neuropathological features have yet been described [67,237,240].
DNAJC6 encodes a neuron-specific isoform of auxilin-1, a co-chaperone involved in clathrin-coat detachment after endocytosis, to ease vesicle recycling, the role of which was explored using human embryonic stem cells with CRISPR-Cas9–mediated gene editing [14,67,241,242,243]. DNAJC6 LOF disrupts synaptic vesicle endocytosis and induces α-synuclein overexpression, thus possibly leading to dopaminergic neurodegeneration [241,242].
In conclusion, monogenic evidence seems to indicate two different groups of genes within the endolysosomal pathway: those more involved in lysosomal-autophagic functions and those encoding for proteins related to the machinery of synaptic vesicles, which are considered a specialized form of recycling endosome present in neurons. The clinical phenotype and neuropathologic findings of the first group seem to recapitulate better the idiopathic form of PD (e.g., LRRK2, VPS35, VPS13C). Conversely, the second group is characterized by atypical phenotypes and, as far as is known, neuropathology (e.g., SYNJ1 and DNAJC6). Therefore, one could speculate that the best models to identify disease mechanisms and therapeutic compounds for the common idiopathic form of PD should be based on the first group of genes, to which GBA1 should also be added (see below). In this context, SNCA and RAB39B probably deserve a separate discussion: although they seem functionally more involved in synaptic function, they display widespread LBD, more in line with idiopathic forms of PD. Therefore, these genes may represent a bridging mechanism between the two groups described above (Figure 1 and Figure 2, Table 1).

3. Genetic Risk Factors for PD Associated with Endolysosomal Dysfunction

Despite these great advances in the field of PD Mendelian genetics, a monogenic cause can be identified in only a minor fraction of PD cases. In general, non-monogenic PD is thought to be a multifactorial disorder influenced by genetic and environmental factors. The emergence of new technological approaches and the increasing sizes of international PD cohorts are leading to the identification of common variants with small effects contributing to PD [244,245]. Pathway analysis of the genes in which these variants are located has shown the involvement of several pathways including mitochondrial biology, inflammation/immune response, and lysosomal-autophagic functions [246,247]. Here, we focus on the risk genes involved in the endolysosomal function (Figure 1).

3.1. GBA1 (Glucosylceramidase Beta 1 or Glucocerebrosidase)

Pathogenic variants in the GBA1 gene represent the most common genetic risk factor for PD: it is estimated that about 8.5% of PD patients worldwide carry a GBA1 mutation. GBA1 carriers display a five- to seven-fold increased risk of developing PD, with a lifelong penetrance of 10–30% [45,151,248,249,250,251,252,253,254,255]. This is especially relevant for specific groups with a high frequency of GBA1 mutations, such as Jews (Ashkenazi Jews PD: 10–31%, non-Ashkenazi Jews PD: 2.9–12%), whereas the frequency in the general population is around 1% [248,256].
The GBA1 gene encodes the glucosylceramidase beta 1 (GCase), a lysosomal hydrolase that breaks glucosylceramide into glucose and acylsphingosine [257,258]. Biallelic GBA1 mutations cause Gaucher’s disease (GD), a lysosomal storage disorder biologically characterized by a significant reduction in GCase activity, which leads to toxic accumulation of glucosylceramide and glucosylsphingosine [259]. Monoallelic GBA1 mutation carriers do not clinically develop GD nor present a pathological accumulation of glucosylceramide and glucosylsphingosine [250,260]. The link between monoallelic GBA1 mutations and PD emerged between 1985 and 1988 when parkinsonism was described as a possible neurological manifestation in GD patients and their relatives carrying heterozygous GBA1 mutations [260,261,262,263,264]. This link was confirmed by subsequent large-scale genetic studies, including GWAS [11].
The clinical features of GBA1-related PD are similar to those of idiopathic PD [250,256,265,266,267,268,269]. However, the average age of onset tends to be slightly earlier (1 to 6 years), clinical progression is generally faster, and survival is shorter [250,270,271,272,273,274,275,276]. Non-motor symptoms including hyposmia, constipation, urinary dysautonomia, orthostatic hypotension, sleep disturbances, depression, anxiety, and especially cognitive impairment (with a decline in working memory, visuospatial, and executive function), are more frequent and impactful [250,256,268,269,271,273,274,275,276,277,278,279,280,281,282,283]. Neuropathological features of GBA1-related PD precisely resemble those of idiopathic PD, with dopaminergic neuron loss and LBD in SN. Cortical LBD was also reported [256,264,269,271,284,285,286,287]. Levodopa response is effective, although MF and LID occur earlier and more frequently [269,276,288] (Figure 2).
Point mutations and complex alleles in GBA1 (resulting from rearrangement with the pseudogene GBAP1) are both reported as strong genetic risk factors for PD. Missense mutations are most represented, including the main causative mutations for GD (e.g., N370S and L444P) as well as PD risk variants that are not associated with GD (e.g., E326K and T369M) [256,289,290,291,292,293,294]. Severity classification for GBA1 mutations considers their impact on the GD phenotype and correlates with the residual GCase activity. The degree of severity of GBA1 mutations, which can be defined as “severe” or “mild”, has a differential effect on penetrance, age at onset, and clinical progression of PD [282,295]. However, it is still not completely clear whether this severity classification derived from GD patients can be automatically translated to the PD field [256,267,271,282,295,296,297].
The disease mechanisms of GBA1-related PD are yet to be understood and may not perfectly match with those of GD, as there are PD risk variants that do not cause GD in homozygosis (e.g., E326K and T369M), and since no definitive correlation has been demonstrated to date between GCase activity and GBA1-related PD [256,290,292,293,298,299]. Nevertheless, the GCase protein was identified in 32–90% of Lewy bodies in patients with GBA1-related PD or Dementia with Lewy bodies (DLB) and less than 10% of Lewy bodies in patients with idiopathic PD, suggesting a possible interaction between GCase and α-synuclein [12,17,256,269,300]. Several pathogenetic hypotheses for GBA1-related PD have been proposed to date, following large studies on PD brains, Drosophila, murine models, cell models including midbrain-like organoids, and iPSC-derived dopamine neurons [256,301] The suggestions include different mechanisms that may depend on both gain- and loss-of-function mutations, including ER stress, ER-Golgi transport impairment, α-synuclein clearance impairment, lysosomal and autophagic dysfunction, lipid homeostasis disruption, mitochondrial dysfunction, and neuroinflammation [256,269,301,302,303,304,305,306,307,308,309,310,311,312]. An intriguing hypothesis is that inactive GCase located on the surface of the lysosomal membrane may change the composition of glycolipids of the membrane affecting the lysosomal internalization of α-synuclein for degradation [97,313]. Another important study showed that lysosomal glucosylceramide accumulation due to GCase deficiency directly interacts with α-synuclein and leads to its accumulation, and vice versa, so that the pathological buildup of α-synuclein may inhibit the transport of GCase to the lysosome, inducing a toxic vicious cycle for neurons [303].
Pathogenic GBA1 variants alone are not sufficient to develop PD, because in groups of patients carrying the same mutations, some develop PD while others do not. Among GBA1 mutation carriers, the development of PD probably depends on other genetic and environmental factors. In this context, two risk-modifier variants (rs356219 in SNCA and rs1293298 in CTSB) were initially identified and relatively little impact was attributed to them, while a larger cumulative impact was reported for common variants affecting genes involved in lysosomal function [12]. Furthermore, other SNCA variants and TMEM175 M393T (see below) may influence the onset age of GBA1-PD [12]. Extending the research into risk-modifying factors to rare predicted deleterious variants in lysosomal genes, the largest contribution to the development of PD in GBA1 mutation carriers was attributed to a second deleterious variant in GBA1 or a deleterious variant in genes associated with mucopolysaccharidoses, evidencing the importance of the overall lysosomal burden in the development of PD [12,17,314].

3.2. SMPD1 (Sphingomyelin Phosphodiesterase 1 or Acid-Sphingomyelinase)

The gene SMPD1 encodes acid sphingomyelinase (ASMase), a lysosomal hydrolase that breaks sphingomyelin into ceramide and phosphocholine [280,281,282,283]. Biallelic SMPD1 mutations cause Niemann–Pick disease (NPD), an LSD characterized by sphingomyelin accumulation, with a heterogeneous range of phenotypes including a severe infantile multisystemic disorder with neurodegeneration (type A) and a later-onset form with absent or minimal neurologic involvement (type B) [315,316,317].
SMPD1 variants L302P and P330fs, highly prevalent among Ashkenazi Jews, were repeatedly associated with PD in this population through case–control studies, as about 1.5% of PD patients carried these mutations, compared with 0.4% of controls [318,319,320,321,322,323]. Previous studies showed that reduced ASMase levels lead to α-synuclein accumulation [323]. It is therefore possible to assume that these mutations increase the risk of developing PD by reducing lysosomal localization of ASMase and thus causing accumulation of pathological α-synuclein [323].
Moreover, several other studies on European and Chinese populations identified additional SMPD1 variants suspected to be associated with PD [254,324,325,326]. Nevertheless, it is important to note that only some of the SMPD1 mutations causing NPD type A/B were demonstrated to be associated with PD [323].

3.3. TMEM175 (Transmembrane Protein 175)

GWAS approaches in PD cases repeatedly identified an association peak on chromosome 4 (TMEM175/GAK/DGKQ locus) representing the fourth strongest risk locus. Two coding variants in this locus localized in the gene TMEM175 (M393T and Q65P) were found to be associated with PD. Transmembrane protein 175 is an integral membrane protein involved in potassium ion transmembrane transport in endolysosomes. Interestingly, the M393T variant was shown to be associated with reduced GCase activity. Therefore, TMEM175 variants are probably responsible for the GWAS association in the TMEM175/GAK/DGKQ locus, as also supported by a Mendelian randomization study [12,13,155,327,328,329,330,331,332,333].

3.4. SCARB2 (Scavenger Receptor Class B Member 2)

SCARB2 variants have been repeatedly identified as risk factors for PD [12,334,335]. SCARB2 encodes the scavenger receptor class B member 2, which, among other endolysosomal activities, transports the GCase protein from the Golgi apparatus to the lysosome. However, unlike the TMEM175 M393T variant, the identified SCARB2 risk variants do not seem to affect GCase activity [335,336,337,338].

3.5. Polygenic “Lysosomal Burden”

An important genetic study using a sequence kernel association test (SKAT-O) investigated deleterious variant burden among lysosomal storage disorder genes using whole exome sequencing data from vast cohorts of PD cases and control subjects. The authors identified a significant burden of rare, potentially damaging lysosomal gene variants in PD patients compared with controls. The association persisted even when the GBA1 gene was excluded from the analysis, suggesting a significant “lysosomal burden” in idiopathic forms of PD. Consistent results were obtained in two independent replication cohorts. The lysosomal genes found to drive the association were GBA1, SMPD1, CTSD, SLC17A5, and ASAH1. Remarkably, in the discovery cohort, the majority of PD cases had at least one deleterious variant in an LSD gene, and 21% carried multiple damaging alleles [254]. Additional clues linking lysosomal dysfunction and “idiopathic” PD include the identification of a deficiency in the activity of the GCase enzyme in blood, CSF, and brain structures not only in GBA1-PD but also in idiopathic forms [339,340].
A recent study investigated the role of deleterious variants affecting LSD genes in modulating the penetrance of GBA1 risk variants in PD. The analysis in the discovery cohort revealed a significantly increased burden of deleterious variants in GBA1-PD patients compared to healthy GBA1 mutation carriers. Moreover, the authors demonstrated that the two strongest modifiers of GBA1 penetrance were a second variation in GBA1 (5.6% vs. 1.4%) and variants in genes causing mucopolysaccharidoses (6.9% vs. 1%) [17]. Furthermore, the largest GWAS on PD to date, based on about 40,000 PD cases including genetic and idiopathic forms, with 1.4M controls, found that an SNP in the gene GALC (rs979812) is associated with PD [11]. GALC encodes galactosylceramidase, a lysosomal hydrolase involved in ceramide catabolism, similar to ASAH1, GBA1, and SMPD1 [13,21,341,342]. Interestingly, the GALC rs979812 variant seems to be associated with increased enzymatic activity of galactosylceramidase [343].
Enrichment analysis based on the same GWAS study revealed that pathways related to lysosomal function are among the most significantly enriched [11]. Other endolysosomal genes identified through a GWAS approach were BAG3, CTSB, GPNMB, GRN, GUSB, HIP1R, LAMP3, NEU1, RAB29, SCARB2, SH3GL2, and TMEM175 [13]. Another GWAS demonstrated significant effects of BAG3, GBA, LAMP3, SCARB2, SNCA, and TMEM175 loci on age at onset of PD [12]. However, it is worth noting that large GWAS, for which perfect matching of cases and controls is virtually impossible, can result in spurious associations because of population stratification. This is also a major concern when studying recently admixed populations, particularly regarding variants with very small effect size [344]. Moreover, the scope of GWAS is limited to identifying genetic associations between PD status and genetic variants, tagging genomic regions encompassing several candidate genes. The “true” causal genes in each identified locus and the mechanisms by which they confer increased risk of PD often remain unclear. This important issue can be addressed by complementing GWAS data with quantitative trait loci (QTL) datasets correlating risk genotypes with gene expression, methylation, and proteomic data (i.e., Mendelian randomization studies—MRS); this knowledge of the molecular mechanisms by which genetic variants localized in PD risk loci increase the disease risk is a key step to translating genetic evidence into possible therapeutic targets. Remarkably, the link between endolysosomal impairment and synaptic dysfunction within PD derives also from these approaches, which have recently confirmed the causal role of several “lysosomal” and “synaptic” genetic hits (i.e., ARSA, CTSB, GALC, IDUA, RAB29, RAB7L1, SH3GL2, SMPD1, STX1B, TMEM175, VAMP42, and ZSWIM7) [333,343,345,346,347,348,349].

4. Unifying Perspective and Conclusions

The previous paragraphs support the hypothesis that genetic variants of genes involved in endolysosomal function are major determinants of PD pathogenesis. Nonetheless, the existence of PD genes and risk variants in genes not directly involved in this pathway suggests that isolated dysfunction of the endolysosomal pathway is unable to explain the full picture, and additional genetic or environmental hits are important factors at play, especially in late-onset non-monogenic forms (Figure 1).
It is an established fact that neuropathological studies of PD brains display abnormal aggregates of misfolded proteins, among which the major actor seems to be aggregated α-synuclein. Other pathological proteins have also been described, for example, aggregated tau in Lrrk2-PD patients [83] (Figure 1).
A dysfunctional endolysosomal system may determine the reduction of α-synuclein turnover, increasing its cytoplasmic concentration and ultimately promoting the formation of oligomeric and fibrillary species, which are considered damaging for dopaminergic neurons [350,351]. An additional link to endolysosomal dysfunction is the presence of membranaceous structures in Lewy bodies, the pathological hallmark of the disease [352]. Moreover, dysfunctional endomembrane trafficking may induce the exploitation of alternative α-synuclein clearance routes, involving the plasma membrane and exocytosis, and ultimately promoting the spread of disease (“prion-like spread hypothesis” of PD) [353,354]. In our view, a vicious circle started by abnormal protein accumulation and subsequent activation of dysfunctional endolysosomal and autophagic processes may be the culprit of this neurodegenerative disease (Figure 1). However, it should be noted that it remains open to debate whether α-synuclein aggregates should be portrayed as the cause or a result of the neurodegenerative process in PD. Taking this forward, a deficiency of functional monomeric α-synuclein has been proposed as a pathogenetic mechanism underlying PD (“proteinopenia” vs. “proteinopathy” hypothesis) [355]. In any case, the strong genetic evidence supporting the role of vesicular and endolysosomal impairment in PD should be carefully considered when formulating any new hypothesis for PD pathogenesis.
The causes of the selective vulnerability of dopaminergic neurons in PD are still unclear. Some peculiar characteristics of these neurons, such as the high metabolic-energetic demands, pace-making activity for the regulated release of dopamine, the pro-oxidant metabolism of this neurotransmitter, and the presence of long and complex arborizations, may suggest possible mechanisms [356]. In this scenario, normal endolysosomal and synaptic functions seem particularly important in the continuous process of recycling synaptic vesicles and the maintenance of long and complex neuronal projections, which are known to recede in PD brains in a dying-back process [357,358,359]. Conversely, high energy needs and oxidative stress management may indicate the mitochondrion as the prime suspect. Unsurprisingly, genetic evidence supports a very significant role also of the mitochondrial pathway (e.g., PRKN and PINK1 mutations) [356].
This review highlights the great importance of research in the field of PD genetics. Starting from the genetic findings accumulated over the past few decades and solidly based on that knowledge, a great number of in vitro and in vivo functional studies of disease models have been conducted, leading to an increased understanding of PD disease mechanisms. However, it is likely that some of the abnormalities observed in these models, although reproducible, are not pathogenetically linked to PD in humans. Some of the animal models do not present the neuropathological or phenotypical features of PD, calling into question the significance of the findings. It is probable that more advanced models, such as patient-derived midbrain organoids carrying specific gene mutations, represent more suitable tools to recapitulate the disease process that occurs in PD patients.
In conclusion, multiple lines of genetic evidence support a major role of vesicular and endolysosomal dysfunction in the pathogenesis of PD. As we have shown, several genes involved in these pathways have been demonstrated to be causative in monogenic forms of PD or have been shown to be associated with increased risk of PD. The crucial role of the endolysosomal system in PD has also generated therapeutic prospects for PD treatment. Indeed, several studies have suggested that the maintenance or upregulation of lysosomal activity may protect against neurodegeneration in PD [360]. In this view, promising therapeutic endolysosomal and autophagic targets include enzymatic activity enhancement (e.g., ambroxol), substrate reduction agents (e.g., venglustat), lysosomal activation (e.g., farnesyltransferase and mTOR inhibitors), and autophagic induction (e.g., trealose and nilotinib) [360]. Currently, many active therapeutic trials in humans have targetted this pathway in patients with genetic and idiopathic forms of PD [269,361,362,363,364,365]. Hopefully, an increased understanding of the specific mechanisms by which dysfunction of these pathways causes PD may allow for the development of targeted drugs that can modify the invariably progressive course of PD.

Author Contributions

Writing—original draft preparation: V.Y. and E.M.; writing—review and editing: V.Y., A.D.F. and E.M. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by Italian Ministry of Health—current research IRCCS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADautosomal dominant
AMPARα-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor
ASMaseacid sphingomyelinase
ARautosomal recessive
CTcomputerized tomography
GCIglial cytoplasmic inclusion
GOFgain of function
EOPearly-onset parkinsonism
EOPDearly-onset Parkinson’s disease
ERendoplasmic reticulum
GCaseglucosylceramidase
GDGaucher’s disease
GOFgain-of-function
GWASgenome-wide association study
IDintellectual disability
iPSCinduced pluripotent stem cell
KRDKufor–Rakeb disease
LBDLewy body disease
LIDlevodopa-induced dyskinesia
LOFloss-of-function
LSDlysosomal storage disease
MFmotor fluctuation
MRImagnetic resonance imaging
NPDNiemann–Pick disease
PDParkinson’s disease
SNsubstantia nigra
XLRX-linked recessive

References

  1. Pringsheim, T.; Jette, N.; Frolkis, A.; Steeves, T.D.L. The Prevalence of Parkinson’s Disease: A Systematic Review and Meta-Analysis. Mov. Disord. 2014, 29, 1583–1590. [Google Scholar] [CrossRef]
  2. Kalia, L.V.; Lang, A.E. Parkinson’s Disease. Lancet 2015, 386, 896–912. [Google Scholar] [CrossRef] [PubMed]
  3. Dorsey, E.R.; Elbaz, A.; Nichols, E.; Abbasi, N.; Abd-Allah, F.; Abdelalim, A.; Adsuar, J.C.; Ansha, M.G.; Brayne, C.; Choi, J.-Y.J.; et al. Global, Regional, and National Burden of Parkinson’s Disease, 1990–2016: A Systematic Analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2018, 17, 939–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tysnes, O.B.; Storstein, A. Epidemiology of Parkinson’s Disease. J. Neural. Transm. 2017, 124, 901–905. [Google Scholar] [CrossRef]
  5. Postuma, R.B.; Berg, D.; Stern, M.; Poewe, W.; Olanow, C.W.; Oertel, W.; Obeso, J.; Marek, K.; Litvan, I.; Lang, A.E.; et al. MDS Clinical Diagnostic Criteria for Parkinson’s Disease. Mov. Disord. 2015, 30, 1591–1601. [Google Scholar] [CrossRef] [PubMed]
  6. Daniel, S.E.; Lees, A.J. Parkinson’s Disease Society Brain Bank, London: Overview and Research. J. Neural. Transm. Suppl. 1993, 39, 165–172. [Google Scholar] [PubMed]
  7. Gelb, D.J.; Oliver, E.; Gilman, S. Diagnostic Criteria for Parkinson Disease. Arch. Neurol. 1999, 56, 33–39. [Google Scholar] [CrossRef]
  8. Dickson, D.W.; Braak, H.; Duda, J.E.; Duyckaerts, C.; Gasser, T.; Halliday, G.M.; Hardy, J.; Leverenz, J.B.; del Tredici, K.; Wszolek, Z.K.; et al. Neuropathological Assessment of Parkinson’s Disease: Refining the Diagnostic Criteria. Lancet Neurol. 2009, 8, 1150–1157. [Google Scholar] [CrossRef]
  9. Balestrino, R.; Schapira, A.H.V. Parkinson Disease. Eur. J. Neurol. 2020, 27, 27–42. [Google Scholar] [CrossRef]
  10. Bandres-Ciga, S.; Diez-Fairen, M.; Kim, J.J.; Singleton, A.B. Genetics of Parkinson’s Disease: An Introspection of Its Journey towards Precision Medicine. Neurobiol. Dis. 2020, 137, 104782. [Google Scholar] [CrossRef]
  11. Nalls, M.A.; Blauwendraat, C.; Vallerga, C.L.; Heilbron, K.; Bandres-Ciga, S.; Chang, D.; Tan, M.; Kia, D.A.; Noyce, A.J.; Xue, A.; et al. Identification of Novel Risk Loci, Causal Insights, and Heritable Risk for Parkinson’s Disease: A Meta-Analysis of Genome-Wide Association Studies. Lancet Neurol. 2019, 18, 1091–1102. [Google Scholar] [CrossRef] [PubMed]
  12. Blauwendraat, C.; Reed, X.; Krohn, L.; Heilbron, K.; Bandres-Ciga, S.; Tan, M.; Gibbs, J.R.; Hernandez, D.G.; Kumaran, R.; Langston, R.; et al. Genetic Modifiers of Risk and Age at Onset in GBA Associated Parkinson’s Disease and Lewy Body Dementia. Brain 2020, 143, 234–248. [Google Scholar] [CrossRef]
  13. Senkevich, K.; Gan-Or, Z. Autophagy Lysosomal Pathway Dysfunction in Parkinson’s Disease; Evidence from Human Genetics. Park. Relat. Disord. 2020, 73, 60–71. [Google Scholar] [CrossRef] [PubMed]
  14. Smolders, S.; van Broeckhoven, C. Genetic Perspective on the Synergistic Connection between Vesicular Transport, Lysosomal and Mitochondrial Pathways Associated with Parkinson’s Disease Pathogenesis. Acta Neuropathol. Commun. 2020, 8, 63. [Google Scholar] [CrossRef] [PubMed]
  15. Zou, L.; Tian, Y.; Zhang, Z. Dysfunction of Synaptic Vesicle Endocytosis in Parkinson’s Disease. Front. Integr. Neurosci. 2021, 15, 619160. [Google Scholar] [CrossRef] [PubMed]
  16. Grover, S.; Sreelatha, A.A.K.; Pihlstrom, L.; Domenighetti, C.; Schulte, C.; Sugier, P.E.; Radivojkov-Blagojevic, M.; Lichtner, P.; Mohamed, O.; Portugal, B.; et al. Genome-Wide Association and Meta-Analysis of Age at Onset in Parkinson Disease: Evidence from the COURAGE-PD Consortium. Neurology 2022, 99, E698–E710. [Google Scholar] [CrossRef]
  17. Straniero, L.; Rimoldi, V.; Monfrini, E.; Bonvegna, S.; Melistaccio, G.; Lake, J.; Soldà, G.; Aureli, M.; Shankaracharya; Keagle, P.; et al. Role of Lysosomal Gene Variants in Modulating GBA-Associated Parkinson’s Disease Risk. Mov. Disord. 2022, 37, 1202–1210. [Google Scholar] [CrossRef]
  18. Ysselstein, D.; Shulman, J.M.; Krainc, D. Emerging Links between Pediatric Lysosomal Storage Diseases and Adult Parkinsonism. Mov. Disord. 2019, 34, 614–624. [Google Scholar] [CrossRef]
  19. Ballabio, A.; Bonifacino, J.S. Lysosomes as Dynamic Regulators of Cell and Organismal Homeostasis. Nat. Rev. Mol. Cell. Biol. 2020, 21, 101–118. [Google Scholar] [CrossRef]
  20. Patra, S.; Patil, S.; Klionsky, D.J.; Bhutia, S.K. Lysosome Signaling in Cell Survival and Programmed Cell Death for Cellular Homeostasis. J. Cell. Physiol. 2022, 238, 287–305. [Google Scholar] [CrossRef]
  21. Platt, F.M.; d’Azzo, A.; Davidson, B.L.; Neufeld, E.F.; Tifft, C.J. Lysosomal Storage Diseases. Nat. Rev. Dis. Prim. 2018, 4, 27. [Google Scholar] [CrossRef]
  22. Pissadaki, E.K.; Bolam, J.P. The Energy Cost of Action Potential Propagation in Dopamine Neurons: Clues to Susceptibility in Parkinson’s Disease. Front. Comput. Neurosci. 2013, 7, 13. [Google Scholar] [CrossRef] [Green Version]
  23. Pacelli, C.; Giguère, N.; Bourque, M.J.; Lévesque, M.; Slack, R.S.; Trudeau, L.É. Elevated Mitochondrial Bioenergetics and Axonal Arborization Size Are Key Contributors to the Vulnerability of Dopamine Neurons. Curr. Biol. 2015, 25, 2349–2360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Giguère, N.; Delignat-Lavaud, B.; Herborg, F.; Voisin, A.; Li, Y.; Jacquemet, V.; Anand-Srivastava, M.; Gether, U.; Giros, B.; Trudeau, L.É. Increased Vulnerability of Nigral Dopamine Neurons after Expansion of Their Axonal Arborization Size through D2 Dopamine Receptor Conditional Knockout. PLoS Genet. 2019, 15, e1008352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Yi, S.; Wang, L.; Wang, H.; Ho, M.S.; Zhang, S. Pathogenesis of α-Synuclein in Parkinson’s Disease: From a Neuron-Glia Crosstalk Perspective. Int. J. Mol. Sci. 2022, 23, 14753. [Google Scholar] [CrossRef]
  26. de Lau, L.M.L.; Breteler, M.M.B. Epidemiology of Parkinson’s Disease. Lancet Neurol. 2006, 5, 525–535. [Google Scholar] [CrossRef] [PubMed]
  27. Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R.; et al. Mutation in the α-Synuclein Gene Identified in Families with Parkinson’s Disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [Green Version]
  28. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. α-Synuclein Locus Triplication Causes Parkinson’s Disease. Science 2003, 302, 841. [Google Scholar] [CrossRef] [Green Version]
  29. Paisán-Ruíz, C.; Jain, S.; Evans, E.W.; Gilks, W.P.; Simón, J.; van der Brug, M.; de Munain, A.L.; Aparicio, S.; Gil, A.M.; Khan, N.; et al. Cloning of the Gene Containing Mutations That Cause PARK8-Linked Parkinson’s Disease. Neuron 2004, 44, 595–600. [Google Scholar] [CrossRef] [Green Version]
  30. Zimprich, A.; Biskup, S.; Leitner, P.; Lichtner, P.; Farrer, M.; Lincoln, S.; Kachergus, J.; Hulihan, M.; Uitti, R.J.; Calne, D.B.; et al. Mutations in LRRK2 Cause Autosomal-Dominant Parkinsonism with Pleomorphic Pathology. Neuron 2004, 44, 601–607. [Google Scholar] [CrossRef] [Green Version]
  31. Berg, D.; Schweitzer, K.; Leitner, P.; Zimprich, A.; Lichtner, P.; Belcredi, P.; Brussel, T.; Schulte, C.; Maass, S.; Nagele, T. Type and Frequency of Mutations in the LRRK2 Gene in Familial and Sporadic Parkinson’s Disease. Brain 2005, 128, 3000–3011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zimprich, A.; Benet-Pagès, A.; Struhal, W.; Graf, E.; Eck, S.H.; Offman, M.N.; Haubenberger, D.; Spielberger, S.; Schulte, E.C.; Lichtner, P.; et al. A Mutation in VPS35, Encoding a Subunit of the Retromer Complex, Causes Late-Onset Parkinson Disease. Am. J. Hum. Genet. 2011, 89, 168–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Vilariño-Güell, C.; Wider, C.; Ross, O.A.; Dachsel, J.C.; Kachergus, J.M.; Lincoln, S.J.; Soto-Ortolaza, A.I.; Cobb, S.A.; Wilhoite, G.J.; Bacon, J.A.; et al. VPS35 Mutations in Parkinson Disease. Am. J. Hum. Genet. 2011, 89, 162–167. [Google Scholar] [CrossRef] [Green Version]
  34. Kitada, T.; Asakawa, S.; Hattori, N.; Matsumine, H.; Yamamura, Y.; Minoshima, S.; Yokochi, M.; Mizuno, Y.; Shimizu, N. Mutations in the Parkin Gene Cause Autosomal Recessive Juvenile Parkinsonism. Nature 1998, 392, 605–608. [Google Scholar] [CrossRef] [PubMed]
  35. Bonifati, V.; Rizzu, P.; van Baren, M.J.; Schaap, O.; Breedveld, G.J.; Krieger, E.; Dekker, M.C.J.; Squitieri, F.; Ibanez, P.; Joosse, M.; et al. Mutations in the DJ-1 Gene Associated with Autosomal Recessive Early-Onset Parkinsonism. Science 2003, 299, 256–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Valente, E.M.; Salvi, S.; Ialongo, T.; Marongiu, R.; Elia, A.E.; Caputo, V.; Romito, L.; Albanese, A.; Dallapiccola, B.; Bentivoglio, A.R. PINK1 Mutations Are Associated with Sporadic Early-Onset Parkinsonism. Ann. Neurol. 2004, 56, 336–341. [Google Scholar] [CrossRef]
  37. Ramirez, A.; Heimbach, A.; Gründemann, J.; Stiller, B.; Hampshire, D.; Cid, L.P.; Goebel, I.; Mubaidin, A.F.; Wriekat, A.L.; Roeper, J.; et al. Hereditary Parkinsonism with Dementia Is Caused by Mutations in ATP13A2, Encoding a Lysosomal Type 5 P-Type ATPase. Nat. Genet. 2006, 38, 1184–1191. [Google Scholar] [CrossRef]
  38. Paisan-Ruiz, C.; Bhatia, K.P.; Li, A.; Hernandez, D.; Davis, M.; Wood, N.W.; Hardy, J.; Houlden, H.; Singleton, A.; Schneider, S.A. Characterization of PLA2G6 as a Locus for Dystonia-Parkinsonism. Ann. Neurol. 2009, 65, 19–23. [Google Scholar] [CrossRef]
  39. di Fonzo, A.; Dekker, M.C.J.; Montagna, P.; Baruzzi, A.; Yonova, E.H.; Guedes, L.C.; Szczerbinska, A.; Zhao, T.; Dubbel-Hulsman, L.O.M.; Wouters, C.H.; et al. FBXO7 Mutations Cause Autosomal Recessive, Early-Onset Parkinsonian-Pyramidal Syndrome. Neurology 2009, 72, 240–245. [Google Scholar] [CrossRef]
  40. Shojaee, S.; Sina, F.; Banihosseini, S.S.; Kazemi, M.H.; Kalhor, R.; Shahidi, G.A.; Fakhrai-Rad, H.; Ronaghi, M.; Elahi, E. Genome-Wide Linkage Analysis of a Parkinsonian-Pyramidal Syndrome Pedigree by 500 K SNP Arrays. Am. J. Hum. Genet. 2008, 82, 1375–1384. [Google Scholar] [CrossRef] [Green Version]
  41. Edvardson, S.; Cinnamon, Y.; Ta-Shma, A.; Shaag, A.; Yim, Y.I.; Zenvirt, S.; Jalas, C.; Lesage, S.; Brice, A.; Taraboulos, A.; et al. A Deleterious Mutation in DNAJC6 Encoding the Neuronal-Specific Clathrin-Uncoating Co-Chaperone Auxilin, Is Associated with Juvenile Parkinsonism. PLoS ONE 2012, 7, 4–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Quadri, M.; Fang, M.; Picillo, M.; Olgiati, S.; Breedveld, G.J.; Graafland, J.; Wu, B.; Xu, F.; Erro, R.; Amboni, M.; et al. Mutation in the SYNJ1 Gene Associated with Autosomal Recessive, Early-Onset Parkinsonism. Hum. Mutat. 2013, 34, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  43. Lesage, S.; Drouet, V.; Majounie, E.; Deramecourt, V.; Jacoupy, M.; Nicolas, A.; Cormier-Dequaire, F.; Hassoun, S.M.; Pujol, C.; Ciura, S.; et al. Loss of VPS13C Function in Autosomal-Recessive Parkinsonism Causes Mitochondrial Dysfunction and Increases PINK1/Parkin-Dependent Mitophagy. Am. J. Hum. Genet. 2016, 98, 500–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wilson, G.R.; Sim, J.C.H.; McLean, C.; Giannandrea, M.; Galea, C.A.; Riseley, J.R.; Stephenson, S.E.M.; Fitzpatrick, E.; Haas, S.A.; Pope, K.; et al. Mutations in RAB39B Cause X-Linked Intellectual Disability and Early-Onset Parkinson Disease with α-Synuclein Pathology. Am. J. Hum. Genet. 2014, 95, 729–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Skrahina, V.; Gaber, H.; Vollstedt, E.J.; Förster, T.M.; Usnich, T.; Curado, F.; Brüggemann, N.; Paul, J.; Bogdanovic, X.; Zülbahar, S.; et al. The Rostock International Parkinson’s Disease (ROPAD) Study: Protocol and Initial Findings. Mov. Disord. 2021, 36, 1005–1010. [Google Scholar] [CrossRef]
  46. Funayama, M.; Hasegawa, K.; Kowa, H.; Saito, M.; Tsuji, S.; Obata, F. A New Locus for Parkinson’s Disease (PARK8) Maps to Chromosome 12p11.2-Q13.1. Ann. Neurol. 2002, 51, 296–301. [Google Scholar] [CrossRef]
  47. Paisán-Ruiz, C.; Lewis, P.A.; Singleton, A.B. LRRK2: Cause, Risk, and Mechanism. J. Park. Dis. 2013, 3, 85–103. [Google Scholar] [CrossRef] [Green Version]
  48. Farrer, M.; Stone, J.; Mata, I.F.; Lincoln, S.; Kachergus, J.; Hulihan, M.; Strain, K.J.; Maraganore, D.M. LRRK2 Mutations in Parkinson Disease. Neurology 2005, 65, 738–740. [Google Scholar] [CrossRef]
  49. Mata, I.F.; Kachergus, J.M.; Taylor, J.P.; Lincoln, S.; Aasly, J.; Lynch, T.; Hulihan, M.M.; Cobb, S.A.; Wu, R.M.; Lu, C.S.; et al. Lrrk2 Pathogenic Substitutions in Parkinson’s Disease. Neurogenetics 2005, 6, 171–177. [Google Scholar] [CrossRef]
  50. Lesage, S.; Ibanez, P.; Lohmann, E.; Pollak, P.; Tison, F.; Tazir, M.; Leutenegger, A.L.; Guimaraes, J.; Bonnet, A.M.; Agid, Y.; et al. G2019S LRRK2 Mutation in French and North African Families with Parkinson’s Disease. Ann. Neurol. 2005, 58, 784–787. [Google Scholar] [CrossRef]
  51. di Fonzo, A.; Rohé, C.F.; Ferreira, J.; Chien, H.F.; Vacca, L.; Stocchi, F.; Guedes, L.; Fabrizio, E.; Manfredi, M.; Vanacore, N.; et al. A Frequent LRRK2 Gene Mutation Associated with Autosomal Dominant Parkinson’s Disease. Lancet 2005, 365, 412–415. [Google Scholar] [CrossRef] [PubMed]
  52. Gilks, W.P.; Abou-Sleiman, P.M.; Gandhi, S.; Jain, S.; Singleton, A.; Lees, A.J.; Shaw, K.; Bhatia, K.P.; Bonifati, V.; Quinn, N.P.; et al. A Common LRRK2 Mutation in Idiopathic Parkinson’s Disease. Lancet 2005, 365, 415–416. [Google Scholar] [CrossRef] [PubMed]
  53. Papapetropoulos, S.; Singer, C.; Ross, O.A.; Toft, M.; Johnson, J.L.; Farrer, M.J.; Mash, D.C. Clinical Heterogeneity of the LRRK2 G2019S Mutation. Arch. Neurol. 2006, 63, 1242–1246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Orr-Urtreger, A.; Shifrin, C.; Rozovski, U.; Rosner, S.; Bercovich, D.; Gurevich, T.; Yagev-More, H.; Bar-Shira, A.; Giladi, N. The LRRK2 G2019S Mutation in Ashkenazi Jews with Parkinson Disease: Is There a Gender Effect? Neurology 2007, 69, 1595–1602. [Google Scholar] [CrossRef] [PubMed]
  55. Xiromerisiou, G.; Hadjigeorgiou, G.M.; Gourbali, V.; Johnson, J.; Papakonstantinou, I.; Papadimitriou, A.; Singleton, A.B. Screening for SNCA and LRRK2 Mutations in Greek Sporadic and Autosomal Dominant Parkinson’s Disease: Identification of Two Novel LRRK2 Variants. Eur. J. Neurol. 2007, 14, 7–11. [Google Scholar] [CrossRef]
  56. Paisán-Ruíz, C.; Nath, P.; Washecka, N.; Gibbs, J.R.; Singleton, A.B. Comprehensive Analysis of LRRK2 in Publicly Available Parkinson’s Disease Cases and Neurologically Normal Controls. Hum. Mutat. 2008, 29, 485–490. [Google Scholar] [CrossRef]
  57. Change, N.; Mercier, G.; Lucotte, G. Genetic Screening of the G2019S Mutation of the LRRK2 Gene in Southwest European, North African, and Sephardic Jewish Subjects. Genet. Test 2008, 12, 333–339. [Google Scholar] [CrossRef]
  58. Bonifati, V.; Wu-Chou, Y.H.; Schweiger, D.; di Fonzo, A.; Lu, C.S.; Oostra, B. Lrrk2 Mutation Analysis in Parkinson Disease Families With Evidence of Linkage To Park8. Neurology 2008, 70, 2348–2349. [Google Scholar] [CrossRef]
  59. Healy, D.G.; Falchi, M.; O’Sullivan, S.S.; Bonifati, V.; Durr, A.; Bressman, S.; Brice, A.; Aasly, J.; Zabetian, C.P.; Goldwurm, S.; et al. Phenotype, Genotype, and Worldwide Genetic Penetrance of LRRK2-Associated Parkinson’s Disease: A Case-Control Study. Lancet Neurol. 2008, 7, 583–590. [Google Scholar] [CrossRef] [Green Version]
  60. Aasly, J.O.; Toft, M.; Fernandez-Mata, I.; Kachergus, J.; Hulihan, M.; White, L.R.; Farrer, M. Clinical Features of LRRK2-Associated Parkinson’s Disease in Central Norway. Ann. Neurol. 2005, 57, 762–765. [Google Scholar] [CrossRef]
  61. di Fonzo, A.; Tassorelli, C.; de Mari, M.; Chien, H.F.; Ferreira, J.; Rohé, C.F.; Riboldazzi, G.; Antonini, A.; Albani, G.; Mauro, A.; et al. Comprehensive Analysis of the LRRK2 Gene in Sixty Families with Parkinson’s Disease. Eur. J. Hum. Genet. 2006, 14, 322–331. [Google Scholar] [CrossRef]
  62. Paisán-Ruiz, C. LRRK2 Gene Variation and Its Contribution to Parkinson Disease. Hum. Mutat. 2009, 30, 1153–1160. [Google Scholar] [CrossRef]
  63. Bardien, S.; Lesage, S.; Brice, A.; Carr, J. Genetic Characteristics of Leucine-Rich Repeat Kinase 2 (LRRK2) Associated Parkinson’s Disease. Park. Relat. Disord. 2011, 17, 501–508. [Google Scholar] [CrossRef]
  64. Lorenzo-Betancor, O.; Samaranch, L.; Ezquerra, M.; Tolosa, E.; Lorenzo, E.; Irigoyen, J.; Gaig, C.; Pastor, M.A.; Soto-Ortolaza, A.I.; Ross, O.A.; et al. LRRK2 Haplotype-Sharing Analysis in Parkinson’s Disease Reveals a Novel p.S1761R Mutation. Mov. Disord. 2012, 27, 146–150. [Google Scholar] [CrossRef]
  65. Chittoor-Vinod, V.G.; Jeremy Nichols, R.; Schüle, B. Genetic and Environmental Factors Influence the Pleomorphy of Lrrk2 Parkinsonism. Int. J. Mol. Sci. 2021, 22, 1045. [Google Scholar] [CrossRef]
  66. Bryant, N.; Malpeli, N.; Ziaee, J.; Blauwendraat, C.; Liu, Z.; West, A.B. Identification of LRRK2 Missense Variants in the Accelerating Medicines Partnership Parkinson’s Disease Cohort. Hum. Mol. Genet. 2021, 30, 454–466. [Google Scholar] [CrossRef]
  67. Jia, F.; Fellner, A.; Kumar, K.R. Monogenic Parkinson’s Disease: Genotype, Phenotype, Pathophysiology, and Genetic Testing. Genes 2022, 13, 471. [Google Scholar] [CrossRef]
  68. Ozelius, L.J.; Senthil, G.; Saunders-Pullman, R.; Ohmann, E.; Deligtisch, A.; Tagliati, M.; Hunt, A.L.; Klein, C.; Henick, B.; Hailpern, S.M.; et al. LRRK2 G2019S as a Cause of Parkinson’s Disease in Ashkenazi Jews. New Engl. J. Med. 2006, 354, 424–425. [Google Scholar] [CrossRef]
  69. Lesage, S.; Dürr, A.; Brice, A. LRRK2 is a major gene in North African parkinsonism. Med. Sci. 2006, 22, 470–471. [Google Scholar] [CrossRef] [Green Version]
  70. Lesage, S.; Dürr, A.; Tazir, M.; Lohmann, E.; Leutenegger, A.-L.; Janin, S.; Pollak, P.; Brice, A. LRRK2 G2019S as a Cause of Parkinson’s Disease in North African Arabs. New Engl. J. Med. 2006, 354, 422–423. [Google Scholar] [CrossRef]
  71. Iwaki, H.; Blauwendraat, C.; Makarious, M.B.; Bandrés-Ciga, S.; Leonard, H.L.; Gibbs, J.R.; Hernandez, D.G.; Scholz, S.W.; Faghri, F.; Nalls, M.A.; et al. Penetrance of Parkinson’s Disease in LRRK2 p.G2019S Carriers Is Modified by a Polygenic Risk Score. Mov. Disord. 2020, 35, 774–780. [Google Scholar] [CrossRef]
  72. Criscuolo, C.; de Rosa, A.; Guacci, A.; Simons, E.J.; Breedveld, G.J.; Peluso, S.; Volpe, G.; Filla, A.; Oostra, B.A.; Bonifati, V.; et al. The LRRK2 R1441C Mutation Is More Frequent than G2019S in Parkinson’s Disease Patients from Southern Italy. Mov. Disord. 2011, 26, 1732–1736. [Google Scholar] [CrossRef] [PubMed]
  73. Simón-Sánchez, J.; Martí-Massó, J.F.; Sánchez-Mut, J.V.; Paisán-Ruiz, C.; Martínez-Gil, A.; Ruiz-Martínez, J.; Sáenz, A.; Singleton, A.B.; Lopéz de Munain, A.; Pérez-Tur, J. Parkinson’s Disease Due to the R1441G Mutation in Dardarin: A Founder Effect in the Basques. Mov. Disord. 2006, 21, 1954–1959. [Google Scholar] [CrossRef] [PubMed]
  74. Xie, C.L.; Pan, J.L.; Wang, W.W.; Zhang, Y.; Zhang, S.F.; Gan, J.; Liu, Z.G. The Association between the LRRK2 G2385R Variant and the Risk of Parkinson’s Disease: A Meta-Analysis Based on 23 Case-Control Studies. Neurol. Sci. 2014, 35, 1495–1504. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, Y.; Sun, Q.; Yi, M.; Zhou, X.; Guo, J.; Xu, Q.; Tang, B.; Yan, X. Genetic Analysis of LRRK2 R1628P in Parkinson’s Disease in Asian Populations. Park. Dis. 2017, 2017, 1–6. [Google Scholar] [CrossRef] [Green Version]
  76. Lim, S.Y.; Tan, A.H.; Ahmad-Annuar, A.; Klein, C.; Tan, L.C.S.; Rosales, R.L.; Bhidayasiri, R.; Wu, Y.R.; Shang, H.F.; Evans, A.H.; et al. Parkinson’s Disease in the Western Pacific Region. Lancet Neurol. 2019, 18, 865–879. [Google Scholar] [CrossRef]
  77. Marras, C.; Schüle, B.; Munhoz, R.P.; Rogaeva, E.; Langston, J.W.; Kasten, M.; Meaney, C.; Klein, C.; Wadia, P.M.; Lim, S.-Y.; et al. Phenotype in Parkinsonian and Nonparkinsonian LRRK2 G2019S Mutation Carriers. Neurology 2011, 77, 325–333. [Google Scholar] [CrossRef]
  78. Agalliu, I.; San Luciano, M.; MirelmanMD, A.; Giladi, N.; Waro, B.; Aasly, J.; Inzelberg, R.; Hassin-Baer, S.; Friedman, E.; Ruiz-Martinez, J.; et al. Higher Frequency of Certain Cancers in LRRK2 G2019S Mutation Carriers with Parkinson Disease a Pooled Analysis. JAMA Neurol. 2015, 72, 58–65. [Google Scholar] [CrossRef] [Green Version]
  79. Warø, B.J.; Aasly, J.O. Exploring Cancer in LRRK2 Mutation Carriers and Idiopathic Parkinson’s Disease. Brain Behav. 2018, 8, e00858. [Google Scholar] [CrossRef] [Green Version]
  80. Agalliu, I.; Ortega, R.A.; Luciano, M.S.; Mirelman, A.; Pont-Sunyer, C.; Brockmann, K.; Vilas, D.; Tolosa, E.; Berg, D.; Warø, B.; et al. Cancer Outcomes among Parkinson’s Disease Patients with Leucine Rich Repeat Kinase 2 Mutations, Idiopathic Parkinson’s Disease Patients, and Nonaffected Controls. Mov. Disord. 2019, 34, 1392–1398. [Google Scholar] [CrossRef]
  81. Macías-García, D.; Periñán, M.T.; Muñoz-Delgado, L.; Jesús, S.; Jimenez-Jaraba, M.V.; Buiza-Rueda, D.; Bonilla-Toribio, M.; Adarmes-Gómez, A.; Carrillo, F.; Gómez-Garre, P.; et al. Increased Stroke Risk in Patients with Parkinson’s Disease with LRRK2 Mutations. Mov. Disord. 2022, 37, 225–227. [Google Scholar] [CrossRef]
  82. Kalia, L.V.; Lang, A.E.; Hazrati, L.N.; Fujioka, S.; Wszolek, Z.K.; Dickson, D.W.; Ross, O.A.; van Deerlin, V.M.; Trojanowski, J.Q.; Hurtig, H.I.; et al. Clinical Correlations with Lewy Body Pathology in LRRK2-Related Parkinson Disease. JAMA Neurol. 2015, 72, 100–105. [Google Scholar] [CrossRef] [Green Version]
  83. Schneider, S.A.; Alcalay, R.N. Neuropathology of Genetic Synucleinopathies with Parkinsonism: Review of the Literature. Mov. Disord. 2017, 32, 1504–1523. [Google Scholar] [CrossRef] [PubMed]
  84. Roosen, D.A.; Cookson, M.R. LRRK2 at the Interface of Autophagosomes, Endosomes and Lysosomes. Mol. Neurodegener. 2016, 11, 73. [Google Scholar] [CrossRef] [Green Version]
  85. Steger, M.; Tonelli, F.; Ito, G.; Davies, P.; Trost, M.; Vetter, M.; Wachter, S.; Lorentzen, E.; Duddy, G.; Wilson, S.; et al. Phosphoproteomics Reveals That Parkinson’ s Disease Kinase LRRK2 Regulates a Subset of Rab GTPases. Elife 2016, 2, e12813. [Google Scholar] [CrossRef] [Green Version]
  86. Ito, G.; Katsemonova, K.; Tonelli, F.; Lis, P.; Baptista, M.A.S.; Shpiro, N.; Duddy, G.; Wilson, S.; Ho, P.W.-L.; Ho, S.-L.; et al. Phos-Tag Analysis of Rab10 Phosphorylation by LRRK2: A Powerful Assay for Assessing Kinase Function and Inhibitors. Biochem. J. 2016, 473, 2671–2685. [Google Scholar] [CrossRef] [Green Version]
  87. Thirstrup, K.; Dächsel, J.C.; Oppermann, F.S.; Williamson, D.S.; Smith, G.P.; Fog, K.; Christensen, K.V. Selective LRRK2 Kinase Inhibition Reduces Phosphorylation of Endogenous Rab10 and Rab12 in Human Peripheral Mononuclear Blood Cells. Sci. Rep. 2017, 7, 10300. [Google Scholar] [CrossRef] [Green Version]
  88. Steger, M.; Diez, F.; Dhekne, H.S.; Lis, P.; Nirujogi, R.S.; Karayel, O.; Tonelli, F.; Martinez, T.N.; Lorentzen, E.; Pfeffer, S.R.; et al. Systematic Proteomic Analysis of LRRK2-Mediated Rab GTPase Phosphorylation Establishes a Connection to Ciliogenesis. Elife 2017, 6, e31012. [Google Scholar] [CrossRef] [Green Version]
  89. Kim, J.; Daadi, E.W.; Oh, T.; Daadi, E.S.; Daadi, M.M. Human Induced Pluripotent Stem Cell Phenotyping and Preclinical Modeling of Familial Parkinson’s Disease. Genes 2022, 13, 1937. [Google Scholar] [CrossRef]
  90. Bose, A.; Petsko, G.A.; Studer, L. Induced Pluripotent Stem Cells: A Tool for Modeling Parkinson’s Disease. Trends Neurosci. 2022, 45, 608–620. [Google Scholar] [CrossRef]
  91. Xiao, B.; Tan, E.K. Targeting LRRK2 in Parkinson’s Disease. Cell Rep. Med. 2022, 3, 100778. [Google Scholar] [CrossRef]
  92. Sanchiz-Calvo, M.; Bentea, E.; Baekelandt, V. Rodent Models Based on Endolysosomal Genes Involved in Parkinson’s Disease. Curr. Opin. Neurobiol. 2022, 72, 55–62. [Google Scholar] [CrossRef]
  93. Piccoli, G.; Volta, M. LRRK2 along the Golgi and Lysosome Connection: A Jamming Situation. Biochem. Soc. Trans. 2021, 49, 2063–2072. [Google Scholar] [CrossRef] [PubMed]
  94. Erb, M.L.; Moore, D.J. LRRK2 and the Endolysosomal System in Parkinson’s Disease. J. Park. Dis. 2020, 10, 1271–1291. [Google Scholar] [CrossRef]
  95. Volta, M.; Melrose, H. LRRK2 Mouse Models: Dissecting the Behavior, Striatal Neurochemistry and Neurophysiology of PD Pathogenesis. Biochem. Soc. Trans. 2017, 45, 113–122. [Google Scholar] [CrossRef] [PubMed]
  96. Alessi, D.R.; Sammler, E. LRRK2 Kinase in Parkinson’s Disease. Science 2018, 360, 36–37. [Google Scholar] [CrossRef] [Green Version]
  97. Gan-Or, Z.; Dion, P.A.; Rouleau, G.A. Genetic Perspective on the Role of the Autophagy-Lysosome Pathway in Parkinson Disease. Autophagy 2015, 11, 1443–1457. [Google Scholar] [CrossRef] [Green Version]
  98. Purlyte, E.; Dhekne, H.S.; Sarhan, A.R.; Gomez, R.; Lis, P.; Wightman, M.; Martinez, T.N.; Tonelli, F.; Pfeffer, S.R.; Alessi, D.R. Rab29 Activation of the Parkinson’s Disease-Associated LRRK2 Kinase. EMBO J. 2018, 37, 1–18. [Google Scholar] [CrossRef]
  99. Ferrazza, R.; Cogo, S.; Melrose, H.; Bubacco, L.; Greggio, E.; Guella, G.; Civiero, L.; Plotegher, N. LRRK2 Deficiency Impacts Ceramide Metabolism in Brain. Biochem. Biophys. Res. Commun. 2016, 478, 1141–1146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Ysselstein, D.; Nguyen, M.; Young, T.J.; Severino, A.; Schwake, M.; Merchant, K.; Krainc, D. LRRK2 Kinase Activity Regulates Lysosomal Glucocerebrosidase in Neurons Derived from Parkinson’s Disease Patients. Nat. Commun. 2019, 10, 5570. [Google Scholar] [CrossRef] [Green Version]
  101. Kedariti, M.; Frattini, E.; Baden, P.; Cogo, S.; Civiero, L.; Ziviani, E.; Zilio, G.; Bertoli, F.; Aureli, M.; Kaganovich, A.; et al. LRRK2 Kinase Activity Regulates GCase Level and Enzymatic Activity Differently Depending on Cell Type in Parkinson’s Disease. NPJ Park. Dis. 2022, 8, 92. [Google Scholar] [CrossRef]
  102. Golbe, L.I.; di Iorio, G.; Sanges, G.; Lazzarini, A.M.; la Sala, S.; Bonavita, V.; Duvoisin, R.C. Clinical Genetic Analysis of Parkinson’s Disease in the Contursi Kindred. Ann. Neurol. 1996, 40, 767–775. [Google Scholar] [CrossRef]
  103. Polymeropoulos, M.H.; Higgins, J.J.; Golbe, L.I.; Johnson, W.G.; Ide, S.E.; di Iorio, G.; Sanges, G.; Stenroos, E.S.; Pho, L.T.; Schaffer, A.A.; et al. Mapping of a Gene for Parkinson’s Disease to Chromosome 4q21-Q23. Science 1996, 274, 1197–1199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.-Y.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. α-Synuclein in Lewy Bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef] [PubMed]
  105. Krüger, R.; Kuhn, W.; Müller, T.; Woitalla, D.; Graeber, M.; Kösel, S.; Przuntek, H.; Epplen, J.T.; Schöls, L.; Riess, O. Ala30Pro Mutation in the Gene Encoding Alpha-Synuclein in Parkinson’s Disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef]
  106. Zarranz, J.J.; Alegre, J.; Gómez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atarés, B.; et al. The New Mutation, E46K, of α-Synuclein Causes Parkinson and Lewy Body Dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef]
  107. Appel-Cresswell, S.; Vilarino-Guell, C.; Encarnacion, M.; Sherman, H.; Yu, I.; Shah, B.; Weir, D.; Thompson, C.; Szu-Tu, C.; Trinh, J.; et al. Alpha-Synuclein p.H50Q, a Novel Pathogenic Mutation for Parkinson’s Disease. Mov. Disord. 2013, 28, 811–813. [Google Scholar] [CrossRef]
  108. Proukakis, C.; Dudzik, C.G.; Brier, T.; MacKay, D.S.; Cooper, J.M.; Millhauser, G.L.; Houlden, H.; Schapira, A.H. A Novel α-Synuclein Missense Mutation in Parkinson Disease. Neurology 2013, 80, 1062–1064. [Google Scholar] [CrossRef] [Green Version]
  109. Kiely, A.P.; Asi, Y.T.; Kara, E.; Limousin, P.; Ling, H.; Lewis, P.; Proukakis, C.; Quinn, N.; Lees, A.J.; Hardy, J.; et al. A-Synucleinopathy Associated with G51D SNCA Mutation: A Link between Parkinson’s Disease and Multiple System Atrophy? Acta Neuropathol. 2013, 125, 753–769. [Google Scholar] [CrossRef] [Green Version]
  110. Lesage, S.; Anheim, M.; Letournel, F.; Bousset, L.; Honoré, A.; Rozas, N.; Pieri, L.; Madiona, K.; Dürr, A.; Melki, R.; et al. G51D α-Synuclein Mutation Causes a Novel Parkinsonian-Pyramidal Syndrome. Ann. Neurol. 2013, 73, 459–471. [Google Scholar] [CrossRef]
  111. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Pöyhönen, M.; Paetau, A. A Novel α-Synuclein Mutation A53E Associated with Atypical Multiple System Atrophy and Parkinson’s Disease-Type Pathology. Neurobiol. Aging 2014, 35, 2180.e1–2180.e5. [Google Scholar] [CrossRef] [PubMed]
  112. Rosborough, K.; Patel, N.; Kalia, L.V. α-Synuclein and Parkinsonism: Updates and Future Perspectives. Curr. Neurol. Neurosci. Rep. 2017, 17, 31. [Google Scholar] [CrossRef]
  113. Trinh, J.; Zeldenrust, F.M.J.; Huang, J.; Kasten, M.; Schaake, S.; Petkovic, S.; Madoev, H.; Grünewald, A.; Almuammar, S.; König, I.R.; et al. Genotype-Phenotype Relations for the Parkinson’s Disease Genes SNCA, LRRK2, VPS35: MDSGene Systematic Review. Mov. Disord. 2018, 33, 1857–1870. [Google Scholar] [CrossRef] [PubMed]
  114. Blauwendraat, C.; Kia, D.A.; Pihlstrøm, L.; Gan-Or, Z.; Lesage, S.; Gibbs, J.R.; Ding, J.; Alcalay, R.N.; Hassin-Baer, S.; Pittman, A.M.; et al. Insufficient Evidence for Pathogenicity of SNCA His50Gln (H50Q) in Parkinson’s Disease. Neurobiol. Aging 2018, 64, 159.e5–159.e8. [Google Scholar] [CrossRef] [PubMed]
  115. Liu, H.; Koros, C.; Strohäker, T.; Schulte, C.; Bozi, M.; Varvaresos, S.; Ibáñez de Opakua, A.; Simitsi, A.M.; Bougea, A.; Voumvourakis, K.; et al. A Novel SNCA A30G Mutation Causes Familial Parkinsonʼs Disease. Mov. Disord. 2021, 36, 1624–1633. [Google Scholar] [CrossRef] [PubMed]
  116. Book, A.; Guella, I.; Candido, T.; Brice, A.; Hattori, N.; Jeon, B.; Farrer, M.J. A Meta-Analysis of α-Synuclein Multiplication in Familial Parkinsonism. Front. Neurol. 2018, 9, 1021. [Google Scholar] [CrossRef] [PubMed]
  117. Lotharius, J.; Brundin, P. Impaired Dopamine Storage Resulting from α-Synuclein Mutations May Contribute to the Pathogenesis of Parkinson’s Disease. Hum. Mol. Genet. 2002, 11, 2395–2407. [Google Scholar] [CrossRef] [Green Version]
  118. Liu, S.; Ninan, I.; Antonova, I.; Battaglia, F.; Trinchese, F.; Narasanna, A.; Kolodilov, N.; Dauer, W.; Hawkins, R.D.; Arancio, O. A-Synuclein Produces a Long-Lasting Increase in Neurotransmitter Release. EMBO J. 2004, 23, 4506–4516. [Google Scholar] [CrossRef] [Green Version]
  119. Nemani, V.M.; Lu, W.; Berge, V.; Nakamura, K.; Onoa, B.; Lee, M.K.; Chaudhry, F.A.; Nicoll, R.A.; Edwards, R.H. Increased Expression of α-Synuclein Reduces Neurotransmitter Release by Inhibiting Synaptic Vesicle Reclustering after Endocytosis. Neuron 2010, 65, 66–79. [Google Scholar] [CrossRef] [Green Version]
  120. Asi, Y.T.; Simpson, J.E.; Heath, P.R.; Wharton, S.B.; Lees, A.J.; Revesz, T.; Houlden, H.; Holton, J.L. Alpha-synuclein MRNA Expression in Oligodendrocytes in MSA. Glia 2014, 62, 964–970. [Google Scholar] [CrossRef] [Green Version]
  121. Wang, C.; Zhao, C.; Li, D.; Tian, Z.; Lai, Y.; Diao, J.; Liu, C. Versatile Structures of α-Synuclein. Front. Mol. Neurosci. 2016, 9, 48. [Google Scholar] [CrossRef] [PubMed]
  122. Arima, K.; Uéda, K.; Sunohara, N.; Arakawa, K.; Hirai, S.; Nakamura, M.; Tonozuka-Uehara, H.; Kawai, M. NACP/α-Synuclein Immunoreactivity in Fibrillary Components of Neuronal and Oligodendroglial Cytoplasmic Inclusions in the Pontine Nuclei in Multiple System Atrophy. Acta Neuropathol. 1998, 96, 439–444. [Google Scholar] [CrossRef] [PubMed]
  123. Hinault, M.P.; Cuendet, A.F.H.; Mattoo, R.U.H.; Mensi, M.; Dietler, G.; Lashuel, H.A.; Goloubinoff, P. Stable α-Synuclein Oligomers Strongly Inhibit Chaperone Activity of the Hsp70 System by Weak Interactions with J-Domain Co-Chaperones. J. Biol. Chem. 2010, 285, 38173–38182. [Google Scholar] [CrossRef] [Green Version]
  124. Kalia, L.V.; Kalia, S.K.; McLean, P.J.; Lozano, A.M.; Lang, A.E. A-Synuclein Oligomers and Clinical Implications for Parkinson Disease. Ann. Neurol. 2013, 73, 155–169. [Google Scholar] [CrossRef]
  125. Colla, E.; Coune, P.; Liu, Y.; Pletnikova, O.; Troncoso, J.C.; Iwatsubo, T.; Schneider, B.L.; Lee, M.K. Endoplasmic Reticulum Stress Is Important for the Manifestations of α-Synucleinopathy in Vivo. J. Neurosci. 2012, 32, 3306–3320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Calo, L.; Wegrzynowicz, M.; Santivañez-Perez, J.; Grazia Spillantini, M. Synaptic Failure and α-Synuclein. Mov. Disord. 2016, 31, 169–177. [Google Scholar] [CrossRef]
  127. di Maio, R.; Barrett, P.J.; Hoffman, E.K.; Barrett, C.W.; Zharikov, A.; Borah, A.; Hu, X.; McCoy, J.; Chu, C.T.; Burton, E.A.; et al. α-Synuclein Binds to TOM20 and Inhibits Mitochondrial Protein Import in Parkinson’s Disease. Sci. Transl. Med. 2016, 8, 342ra78. [Google Scholar] [CrossRef] [Green Version]
  128. Xilouri, M.; Brekk, O.R.; Stefanis, L. Autophagy and Alpha-Synuclein: Relevance to Parkinson’s Disease and Related Synucleopathies. Mov. Disord. 2016, 31, 178–192. [Google Scholar] [CrossRef]
  129. Maday, S.; Holzbaur, E.L.F. Compartment-Specific Regulation of Autophagy in Primary Neurons. J. Neurosci. 2016, 36, 5933–5945. [Google Scholar] [CrossRef] [Green Version]
  130. Sacino, A.N.; Brooks, M.M.; Chakrabarty, P.; Saha, K.; Khoshbouei, H.; Golde, T.E.; Giasson, B.I. Proteolysis of α-Synuclein Fibrils in the Lysosomal Pathway Limits Induction of Inclusion Pathology. J. Neurochem. 2017, 140, 662–678. [Google Scholar] [CrossRef] [Green Version]
  131. Volpicelli-Daley, L.A. Effects of α-Synuclein on Axonal Transport. Neurobiol. Dis. 2017, 105, 321–327. [Google Scholar] [CrossRef]
  132. Miller, D.W.; Hague, S.M.; Clarimon, J.; Baptista, M.; Gwinn-Hardy, K.; Cookson, M.R.; Singleton, A.B. Alpha-Synuclein in Blood and Brain from Familial Parkinson Disease with SNCA Locus Triplication. Neurology 2004, 62, 1835–1838. [Google Scholar] [CrossRef]
  133. Devine, M.J.; Gwinn, K.; Singleton, A.; Hardy, J. Parkinson’s Disease and α-Synuclein Expression. Mov. Disord. 2011, 26, 2160–2168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Cuervo, A.M.; Stefanis, L.; Fredenburg, R.; Lansbury, P.T.; Sulzer, D. Impaired Degradation of Mutant Alpha-Synuclein by Chaperone-Mediated Autophagy. Science 2004, 305, 1292–1295. [Google Scholar] [CrossRef]
  135. Stykel, M.G.; Humphries, K.M.; Kamski-Hennekam, E.; Buchner-Duby, B.; Porte-Trachsel, N.; Ryan, T.; Coackley, C.L.; Bamm, V.V.; Harauz, G.; Ryan, S.D. α-Synuclein Mutation Impairs Processing of Endomembrane Compartments and Promotes Exocytosis and Seeding of α-Synuclein Pathology. Cell Rep. 2021, 35, 109099. [Google Scholar] [CrossRef] [PubMed]
  136. Gambardella, S.; Biagioni, F.; Ferese, R.; Busceti, C.L.; Frati, A.; Novelli, G.; Ruggieri, S.; Fornai, F. Vacuolar Protein Sorting Genes in Parkinson’s Disease: A Re-Appraisal of Mutations Detection Rate and Neurobiology of Disease. Front. Neurosci. 2016, 10, 532. [Google Scholar] [CrossRef]
  137. Rahman, A.A.; Morrison, B.E. Contributions of VPS35 Mutations to Parkinson’s Disease. Neuroscience 2019, 401, 1–10. [Google Scholar] [CrossRef] [PubMed]
  138. Williams, E.T.; Chen, X.; Otero, P.A.; Moore, D.J. Understanding the Contributions of VPS35 and the Retromer in Neurodegenerative Disease. Neurobiol. Dis. 2022, 170, 105768. [Google Scholar] [CrossRef] [PubMed]
  139. Wider, C.; Skipper, L.; Solida, A.; Brown, L.; Farrer, M.; Dickson, D.; Wszolek, Z.K.; Vingerhoets, F.J.G. Autosomal Dominant Dopa-Responsive Parkinsonism in a Multigenerational Swiss Family. Park. Relat. Disord. 2008, 14, 465–470. [Google Scholar] [CrossRef]
  140. Struhal, W.; Presslauer, S.; Spielberger, S.; Zimprich, A.; Auff, E.; Bruecke, T.; Poewe, W.; Ransmayr, G.; Wolf, E.; Seppi, K.; et al. VPS35 Parkinson’s Disease Phenotype Resembles the Sporadic Disease. J. Neural. Transm. 2014, 121, 755–759. [Google Scholar] [CrossRef] [PubMed]
  141. Ishiguro, M.; Li, Y.; Yoshino, H.; Daida, K.; Ishiguro, Y.; Oyama, G.; Saiki, S.; Funayama, M.; Hattori, N.; Nishioka, K. Clinical Manifestations of Parkinson’s Disease Harboring VPS35 Retromer Complex Component p.D620N with Long-Term Follow-Up. Park. Relat. Disord. 2021, 84, 139–143. [Google Scholar] [CrossRef] [PubMed]
  142. Williams, E.T.; Chen, X.; Moore, D.J. VPS35, the Retromer Complex and Parkinson’s Disease. J. Park. Dis. 2017, 7, 219–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Seaman, M.N.J. The Retromer Complex: From Genesis to Revelations. Trends Biochem. Sci. 2021, 46, 608–620. [Google Scholar] [CrossRef] [PubMed]
  144. Zavodszky, E.; Seaman, M.N.J.; Moreau, K.; Jimenez-Sanchez, M.; Breusegem, S.Y.; Harbour, M.E.; Rubinsztein, D.C. Mutation in VPS35 Associated with Parkinson’s Disease Impairs WASH Complex Association and Inhibits Autophagy. Nat. Commun. 2014, 5, 3828. [Google Scholar] [CrossRef] [Green Version]
  145. Follett, J.; Norwood, S.J.; Hamilton, N.A.; Mohan, M.; Kovtun, O.; Tay, S.; Zhe, Y.; Wood, S.A.; Mellick, G.D.; Silburn, P.A.; et al. The Vps35 D620N Mutation Linked to Parkinson’s Disease Disrupts the Cargo Sorting Function of Retromer. Traffic 2014, 15, 230–244. [Google Scholar] [CrossRef]
  146. Braschi, E.; Goyon, V.; Zunino, R.; Mohanty, A.; Xu, L.; McBride, H.M. Vps35 Mediates Vesicle Transport between the Mitochondria and Peroxisomes. Curr. Biol. 2010, 20, 1310–1315. [Google Scholar] [CrossRef] [Green Version]
  147. Wang, W.; Wang, X.; Fujioka, H.; Hoppel, C.; Whone, A.L.; Caldwell, M.A.; Cullen, P.J.; Liu, J.; Zhu, X. Parkinson’s Disease-Associated Mutant VPS35 Causes Mitochondrial Dysfunction by Recycling DLP1 Complexes. Nat. Med. 2016, 22, 54–63. [Google Scholar] [CrossRef] [Green Version]
  148. MacLeod, D.A.; Rhinn, H.; Kuwahara, T.; Zolin, A.; DiPaolo, G.; McCabe, B.D.; Marder, K.S.; Honig, L.S.; Clark, L.N.; Small, S.A.; et al. RAB7L1 Interacts with LRRK2 to Modify Intraneuronal Protein Sorting and Parkinson’s Disease Risk. Neuron 2013, 79, 202–203. [Google Scholar] [CrossRef] [Green Version]
  149. Hanss, Z.; Larsen, S.B.; Antony, P.; Mencke, P.; Massart, F.; Jarazo, J.; Schwamborn, J.C.; Barbuti, P.A.; Mellick, G.D.; Krüger, R. Mitochondrial and Clearance Impairment in p.D620N VPS35 Patient-Derived Neurons. Mov. Disord. 2021, 36, 704–715. [Google Scholar] [CrossRef]
  150. Kobayashi, R.; Naruse, H.; Koyama, S.; Kawakatsu, S.; Hayashi, H.; Ishiura, H.; Mitsui, J.; Ohta, Y.; Toda, T.; Tsuji, S.; et al. Familial Dementia with Lewy Bodies with VPS13C Mutations. Park. Relat. Disord. 2020, 81, 31–33. [Google Scholar] [CrossRef]
  151. Nalls, M.A.; Pankratz, N.; Lill, C.M.; Do, C.B.; Hernandez, D.G.; Saad, M.; Destefano, A.L.; Kara, E.; Bras, J.; Sharma, M.; et al. Large-Scale Meta-Analysis of Genome-Wide Association Data Identifies Six New Risk Loci for Parkinson’s Disease. Nat. Genet. 2014, 46, 989–993. [Google Scholar] [CrossRef] [Green Version]
  152. Monfrini, E.; Spagnolo, F.; Canesi, M.; Seresini, A.; Passarella, B.; Percetti, M.; Seia, M.; Goldwurm, S.; Cereda, V.; Comi, G.P.; et al. VPS13C-Associated Parkinson’s Disease: Two Novel Cases and Review of the Literature. Park. Relat. Disord. 2022, 94, 37–39. [Google Scholar] [CrossRef] [PubMed]
  153. Schormair, B.; Kemlink, D.; Mollenhauer, B.; Fiala, O.; Machetanz, G.; Roth, J.; Berutti, R.; Strom, T.M.; Haslinger, B.; Trenkwalder, C.; et al. Diagnostic Exome Sequencing in Early-Onset Parkinson’s Disease Confirms VPS13C as a Rare Cause of Autosomal-Recessive Parkinson’s Disease. Clin. Genet. 2018, 93, 603–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Darvish, H.; Bravo, P.; Tafakhori, A.; Azcona, L.J.; Ranji-Burachaloo, S.; Johari, A.H.; Paisán-Ruiz, C. Identification of a Large Homozygous VPS13C Deletion in a Patient with Early-Onset Parkinsonism. Mov. Disord. 2018, 33, 1968–1970. [Google Scholar] [CrossRef] [PubMed]
  155. Hopfner, F.; Mueller, S.H.; Szymczak, S.; Junge, O.; Tittmann, L.; May, S.; Lohmann, K.; Grallert, H.; Lieb, W.; Strauch, K.; et al. Rare Variants in Specific Lysosomal Genes Are Associated With Parkinson’s Disease. Mov. Disord. 2020, 35, 1245–1248. [Google Scholar] [CrossRef] [PubMed]
  156. Gu, X.; Li, C.; Chen, Y.; Ou, R.; Cao, B.; Wei, Q.; Hou, Y.; Zhang, L.; Song, W.; Zhao, B.; et al. Mutation Screening and Burden Analysis of VPS13C in Chinese Patients with Early-Onset Parkinson’s Disease. Neurobiol. Aging 2020, 94, 311.e1–311.e4. [Google Scholar] [CrossRef]
  157. Smolders, S.; Philtjens, S.; Crosiers, D.; Sieben, A.; Hens, E.; Heeman, B.; van Mossevelde, S.; Pals, P.; Asselbergh, B.; dos Santos Dias, R.; et al. Contribution of Rare Homozygous and Compound Heterozygous VPS13C Missense Mutations to Dementia with Lewy Bodies and Parkinson’s Disease. Acta Neuropathol. Commun. 2021, 9, 25. [Google Scholar] [CrossRef]
  158. Kumar, N.; Leonzino, M.; Hancock-Cerutti, W.; Horenkamp, F.A.; Li, P.Q.; Lees, J.A.; Wheeler, H.; Reinisch, K.M.; de Camilli, P. VPS13A and VPS13C Are Lipid Transport Proteins Differentially Localized at ER Contact Sites. J. Cell Biol. 2018, 217, 3625–3639. [Google Scholar] [CrossRef] [Green Version]
  159. Hancock-Cerutti, W.; Wu, Z.; Xu, P.; Yadavalli, N.; Leonzino, M.; Tharkeshwar, A.K.; Ferguson, S.M.; Shadel, G.S.; de Camilli, P. ER-Lysosome Lipid Transfer Protein VPS13C/PARK23 Prevents Aberrant MtDNA-Dependent STING Signaling. J. Cell Biol. 2022, 221, e202106046. [Google Scholar] [CrossRef]
  160. Bras, J.; Verloes, A.; Schneider, S.A.; Mole, S.E.; Guerreiro, R.J. Mutation of the Parkinsonism Gene ATP13A2 Causes Neuronal Ceroid-Lipofuscinosis. Hum. Mol. Genet. 2012, 21, 2646–2650. [Google Scholar] [CrossRef] [Green Version]
  161. Estrada-Cuzcano, A.; Martin, S.; Chamova, T.; Synofzik, M.; Timmann, D.; Holemans, T.; Andreeva, A.; Reichbauer, J.; de Rycke, R.; Chang, D.I.; et al. Loss-of-Function Mutations in the ATP13A2/PARK9 Gene Cause Complicated Hereditary Spastic Paraplegia (SPG78). Brain 2017, 140, 287–305. [Google Scholar] [CrossRef] [Green Version]
  162. Najim al-Din, A.S.; Wriekat, A.; Mubaidin, A.; Dasouki, M.; Hiari, M. Pallido-Pyramidal Degeneration, Supranuclear Upgaze Paresis and Dementia: Kufor-Rakeb Syndrome. Acta Neurol. Scand. 1994, 89, 347–352. [Google Scholar] [CrossRef] [PubMed]
  163. Williams, D.R.; Hadeed, A.; Najim al-Din, A.S.; Wreikat, A.L.; Lees, A.J. Kufor Rakeb Disease: Autosomal Recessive, Levodopa-Responsive Parkinsonism with Pyramidal Degeneration, Supranuclear Gaze Palsy, and Dementia. Mov. Disord. 2005, 20, 1264–1271. [Google Scholar] [CrossRef] [PubMed]
  164. Hampshire, D.J.; Roberts, E.; Crow, Y.; Bond, J.; Mubaidin, A.; Wriekat, A.L.; Al-Din, A.; Woods, C.G. Kufor-Rakeb Syndrome, Pallido-Pyramidal Degeneration with Supranuclear Upgaze Paresis and Dementia, Maps to 1p36. J. Med. Genet. 2001, 38, 680–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Schneider, S.A.; Paisan-Ruiz, C.; Quinn, N.P.; Lees, A.J.; Houlden, H.; Hardy, J.; Bhatia, K.P. ATP13A2 Mutations (PARK9) Cause Neurodegeneration with Brain Iron Accumulation. Mov. Disord. 2010, 25, 979–984. [Google Scholar] [CrossRef] [PubMed]
  166. Crosiers, D.; Ceulemans, B.; Meeus, B.; Nuytemans, K.; Pals, P.; van Broeckhoven, C.; Cras, P.; Theuns, J. Juvenile Dystonia-Parkinsonism and Dementia Caused by a Novel ATP13A2 Frameshift Mutation. Park. Relat. Disord. 2011, 17, 135–138. [Google Scholar] [CrossRef] [PubMed]
  167. Park, J.S.; Mehta, P.; Cooper, A.A.; Veivers, D.; Heimbach, A.; Stiller, B.; Kubisch, C.; Fung, V.S.; Krainc, D.; Mackay-Sim, A.; et al. Pathogenic Effects of Novel Mutations in the P-Type ATPase ATP13A2 (PARK9) Causing Kufor-Rakeb Syndrome, a Form of Early-Onset Parkinsonism. Hum. Mutat. 2011, 32, 956–964. [Google Scholar] [CrossRef] [PubMed]
  168. Chien, H.F.; Bonifati, V.; Barbosa, E.R. ATP13A2-Related Neurodegeneration (PARK9) without Evidence of Brain Iron Accumulation. Mov. Disord. 2011, 26, 1364–1365. [Google Scholar] [CrossRef]
  169. Ugolino, J.; Fang, S.; Kubisch, C.; Monteiro, M.J. Mutant Atp13a2 Proteins Involved in Parkinsonism Are Degraded by ER-Associated Degradation and Sensitize Cells to ER-Stress Induced Cell Death. Hum. Mol. Genet. 2011, 20, 3565–3577. [Google Scholar] [CrossRef] [Green Version]
  170. Eiberg, H.; Hansen, L.; Korbo, L.; Nielsen, I.; Svenstrup, K.; Bech, S.; Pinborg, L.; Friberg, L.; Hjermind, L.; Olsen, O.; et al. Novel Mutation in ATP13A2 Widens the Spectrum of Kufor-Rakeb Syndrome (PARK9). Clin. Genet. 2012, 82, 256–263. [Google Scholar] [CrossRef]
  171. van Veen, S.; Sørensen, D.M.; Holemans, T.; Holen, H.W.; Palmgren, M.G.; Vangheluwe, P. Cellular Function and Pathological Role of ATP13A2 and Related P-Type Transport ATPases in Parkinson’s Disease and Other Neurological Disorders. Front. Mol. Neurosci. 2014, 7, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. McNeil-Gauthier, A.-L.; Brais, B.; Rouleau, G.; Anoja, N.; Ducharme, S. Successful Treatment of Psychosis in a Patient with Kufor-Rakeb Syndrome with Low Dose Aripiprazole: A Case Report. Neurocase 2019, 25, 133–137. [Google Scholar] [CrossRef]
  173. Jellinger, K.A. Pallidal Degenerations and Related Disorders: An Update. J. Neural. Transm. 2022, 129, 521–543. [Google Scholar] [CrossRef] [PubMed]
  174. Behrens, M.I.; Brüggemann, N.; Chana, P.; Venegas, P.; Kägi, M.; Parrao, T.; Orellana, P.; Garrido, C.; Rojas, C.V.; Hauke, J.; et al. Clinical Spectrum of Kufor-Rakeb Syndrome in the Chilean Kindred with ATP13A2 Mutations. Mov. Disord. 2010, 25, 1929–1937. [Google Scholar] [CrossRef]
  175. Brüggemann, N.; Hagenah, J.; Reetz, K.; Schmidt, A.; Kasten, M.; Buchmann, I.; Eckerle, S.; Bähre, M.; Münchau, A.; Djarmati, A.; et al. Recessively Inherited Parkinsonism Effect OfATP13A2 Mutations on the Clinical and Neuroimaging Phenotype. Arch. Neurol. 2010, 67, 1357–1363. [Google Scholar] [CrossRef] [Green Version]
  176. Park, J.S.; Blair, N.F.; Sue, C.M. The Role of ATP13A2 in Parkinson’s Disease: Clinical Phenotypes and Molecular Mechanisms. Mov. Disord. 2015, 30, 770–779. [Google Scholar] [CrossRef]
  177. di Fonzo, A.; Chien, H.F.; Socal, M.; Giraudo, S.; Tassorelli, C.; Iliceto, G.; Fabbrini, G.; Marconi, R.; Fincati, E.; Abbruzzese, G.; et al. ATP13A2 Missense Mutations in Juvenile Parkinsonism and Young Onset Parkinson Disease. Neurology 2007, 68, 1557–1562. [Google Scholar] [CrossRef] [PubMed]
  178. Malakouti-Nejad, M.; Shahidi, G.A.; Rohani, M.; Shojaee, S.M.; Hashemi, M.; Klotzle, B.; Fan, J.B.; Elahi, E. Identification of p.Gln858* in ATP13A2 in Two EOPD Patients and Presentation of Their Clinical Features. Neurosci. Lett. 2014, 577, 106–111. [Google Scholar] [CrossRef]
  179. Ning, Y.P.; Kanai, K.; Tomiyama, H.; Li, Y.; Funayama, M.; Yoshino, H.; Sato, S.; Asahina, M.; Kuwabara, S.; Takeda, A.; et al. PARK9-Linked Parkinsonism in Eastern Asia, Mutation Detection in ATP13A2 and Clinical Phenotype. Neurology 2008, 70, 1491–1493. [Google Scholar] [CrossRef]
  180. Lin, C.H.; Tan, E.K.; Chen, M.L.; Tan, L.C.; Lim, H.Q.; Chen, G.S.; Wu, R.M. Novel ATP13A2 Variant Associated with Parkinson Disease in Taiwan and Singapore. Neurology 2008, 71, 1727–1732. [Google Scholar] [CrossRef]
  181. Djarmati, A.; Hagenah, J.; Reetz, K.; Winkler, S.; Behrens, M.I.; Pawlack, H.; Lohmann, K.; Ramirez, A.; Tadić, V.; Brüggemann, N.; et al. ATP13A2 Variants in Early-Onset Parkinson’s Disease Patients and Controls. Mov. Disord. 2009, 24, 2104–2111. [Google Scholar] [CrossRef]
  182. Santoro, L.; Breedveld, G.J.; Manganelli, F.; Iodice, R.; Pisciotta, C.; Nolano, M.; Punzo, F.; Quarantelli, M.; Pappatà, S.; di Fonzo, A.; et al. Novel ATP13A2 (PARK9) Homozygous Mutation in a Family with Marked Phenotype Variability. Neurogenetics 2011, 12, 33–39. [Google Scholar] [CrossRef] [Green Version]
  183. Chien, H.F.; Rodriguez, R.D.; Bonifati, V.; Nitrini, R.; Pasqualucci, C.A.; Gelpi, E.; Barbosa, E.R. Neuropathologic Findings in a Patient With Juvenile-Onset Levodopa-Responsive Parkinsonism Due to ATP13A2 Mutation. Neurology 2021, 97, 763–766. [Google Scholar] [CrossRef]
  184. Martin, S.; Holemans, T.; Vangheluwe, P. Unlocking Atp13a2/Park9 Activity. Cell Cycle 2015, 14, 3341–3342. [Google Scholar] [CrossRef] [Green Version]
  185. Sørensen, D.M.; Holemans, T.; van Veen, S.; Martin, S.; Arslan, T.; Haagendahl, I.W.; Holen, H.W.; Hamouda, N.N.; Eggermont, J.; Palmgren, M.; et al. Parkinson Disease Related ATP13A2 Evolved Early in Animal Evolution. PLoS ONE 2018, 13, e0193228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Dirr, E.R.; Ekhator, O.R.; Blackwood, R.; Holden, J.G.; Masliah, E.; Schultheis, P.J.; Fleming, S.M. Exacerbation of Sensorimotor Dysfunction in Mice Deficient in Atp13a2 and Overexpressing Human Wildtype Alpha-Synuclein. Behav. Brain Res. 2018, 343, 41–49. [Google Scholar] [CrossRef] [PubMed]
  187. Radi, E.; Formichi, P.; di Maio, G.; Battisti, C.; Federico, A. Altered Apoptosis Regulation in Kufor-Rakeb Syndrome Patients with Mutations in the ATP13A2 Gene. J. Cell Mol. Med. 2012, 16, 1916–1923. [Google Scholar] [CrossRef] [PubMed]
  188. Park, J.S.; Koentjoro, B.; Davis, R.L.; Sue, C.M. Loss of ATP13A2 Impairs Glycolytic Function in Kufor-Rakeb Syndrome Patient-Derived Cell Models. Park. Relat. Disord. 2016, 27, 67–73. [Google Scholar] [CrossRef]
  189. Wang, R.; Tan, J.; Chen, T.; Han, H.; Tian, R.; Tan, Y.; Wu, Y.; Cui, J.; Chen, F.; Li, J.; et al. ATP13A2 Facilitates HDAC6 Recruitment to Lysosome to Promote Autophagosome–Lysosome Fusion. J. Cell Biol. 2019, 218, 267–284. [Google Scholar] [CrossRef]
  190. van Veen, S.; Martin, S.; van den Haute, C.; Benoy, V.; Lyons, J.; Vanhoutte, R.; Kahler, J.P.; Decuypere, J.P.; Gelders, G.; Lambie, E.; et al. ATP13A2 Deficiency Disrupts Lysosomal Polyamine Export. Nature 2020, 578, 419–424. [Google Scholar] [CrossRef]
  191. Si, J.; van den Haute, C.; Lobbestael, E.; Martin, S.; van Veen, S.; Vangheluwe, P.; Baekelandt, V. ATP13A2 Regulates Cellular α-Synuclein Multimerization, Membrane Association, and Externalization. Int. J. Mol. Sci. 2021, 22, 2689. [Google Scholar] [CrossRef]
  192. Russo, S.; Cogliati, F.; Cavalleri, F.; Cassitto, M.G.; Giglioli, R.; Toniolo, D.; Casari, G.; Larizza, L. Mapping to Distal Xq28 of Nonspecific X-Linked Mental Retardation MRX72: Linkage Analysis and Clinical Findings in a Three-Generation Sardinian Family. Am. J. Med. Genet. 2000, 94, 376–382. [Google Scholar] [CrossRef] [PubMed]
  193. Giannandrea, M.; Bianchi, V.; Mignogna, M.L.; Sirri, A.; Carrabino, S.; D’Elia, E.; Vecellio, M.; Russo, S.; Cogliati, F.; Larizza, L.; et al. Mutations in the Small GTPase Gene RAB39B Are Responsible for X-Linked Mental Retardation Associated with Autism, Epilepsy, and Macrocephaly. Am. J. Hum. Genet. 2010, 86, 185–195. [Google Scholar] [CrossRef] [Green Version]
  194. Lesage, S.; Bras, J.; Cormier-Dequaire, F.; Condroyer, C.; Nicolas, A.; Darwent, L.; Guerreiro, R.; Majounie, E.; Federoff, M.; Heutink, P.; et al. Loss-of-Function Mutations in RAB39B Are Associated with Typical Early-Onset Parkinson Disease. Neurol. Genet. 2015, 1, e9. [Google Scholar] [CrossRef] [Green Version]
  195. Mata, I.F.; Jang, Y.; Kim, C.H.; Hanna, D.S.; Dorschner, M.O.; Samii, A.; Agarwal, P.; Roberts, J.W.; Klepitskaya, O.; Shprecher, D.R.; et al. The RAB39B p.G192R Mutation Causes X-Linked Dominant Parkinson’s Disease. Mol. Neurodegener. 2015, 10, 4–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Shi, C.H.; Zhang, S.Y.; Yang, Z.H.; Yang, J.; Shang, D.D.; Mao, C.Y.; Liu, H.; Hou, H.M.; Shi, M.M.; Wu, J.; et al. A Novel RAB39B Gene Mutation in X-Linked Juvenile Parkinsonism with Basal Ganglia Calcification. Mov. Disord. 2016, 31, 1905–1909. [Google Scholar] [CrossRef] [PubMed]
  197. Güldner, M.; Schulte, C.; Hauser, A.K.; Gasser, T.; Brockmann, K. Broad Clinical Phenotype in Parkinsonism Associated with a Base Pair Deletion in RAB39B and Additional POLG Variant. Park. Relat. Disord. 2016, 31, 148–150. [Google Scholar] [CrossRef]
  198. Ciammola, A.; Carrera, P.; di Fonzo, A.; Sassone, J.; Villa, R.; Poletti, B.; Ferrari, M.; Girotti, F.; Monfrini, E.; Buongarzone, G.; et al. X-Linked Parkinsonism with Intellectual Disability Caused by Novel Mutations and Somatic Mosaicism in RAB39B Gene. Park. Relat. Disord. 2017, 44, 142–146. [Google Scholar] [CrossRef]
  199. Woodbury-Smith, M.; Deneault, E.; Yuen, R.K.C.; Walker, S.; Zarrei, M.; Pellecchia, G.; Howe, J.L.; Hoang, N.; Uddin, M.; Marshall, C.R.; et al. Mutations in RAB39B in Individuals with Intellectual Disability, Autism Spectrum Disorder, and Macrocephaly. Mol. Autism. 2017, 8, 59. [Google Scholar] [CrossRef] [Green Version]
  200. Puschmann, A. New Genes Causing Hereditary Parkinson’s Disease or Parkinsonism. Curr. Neurol. Neurosci. Rep. 2017, 17, 1–11. [Google Scholar] [CrossRef] [Green Version]
  201. Gao, Y.; Wilson, G.R.; Salce, N.; Romano, A.; Mellick, G.D.; Stephenson, S.E.M.; Lockhart, P.J. Genetic Analysis of RAB39B in an Early-Onset Parkinson’s Disease Cohort. Front. Neurol. 2020, 11, 523. [Google Scholar] [CrossRef] [PubMed]
  202. Mignogna, M.L.; Musardo, S.; Ranieri, G.; Gelmini, S.; Espinosa, P.; Marra, P.; Belloli, S.; Murtaj, V.; Moresco, R.M.; Bellone, C.; et al. RAB39B-Mediated Trafficking of the GluA2-AMPAR Subunit Controls Dendritic Spine Maturation and Intellectual Disability-Related Behaviour. Mol. Psychiatry 2021, 26, 6531–6549. [Google Scholar] [CrossRef]
  203. Tang, B.L. RAB39B’s Role in Membrane Traffic, Autophagy, and Associated Neuropathology. J. Cell Physiol. 2021, 236, 1579–1592. [Google Scholar] [CrossRef]
  204. El-Hattab, A.W.; Fang, P.; Jin, W.; Hughes, J.R.; Gibson, J.B.; Patel, G.S.; Grange, D.K.; Manwaring, L.P.; Patel, A.; Stankiewicz, P.; et al. Int22h-1/Int22h-2-Mediated Xq28 Rearrangements: Intellectual Disability Associated with Duplications and in Utero Male Lethality with Deletions. J. Med. Genet. 2011, 48, 840–850. [Google Scholar] [CrossRef]
  205. Vanmarsenille, L.; Giannandrea, M.; Fieremans, N.; Verbeeck, J.; Belet, S.; Raynaud, M.; Vogels, A.; Männik, K.; Õunap, K.; Jacqueline, V.; et al. Increased Dosage of RAB39B Affects Neuronal Development and Could Explain the Cognitive Impairment in Male Patients with Distal Xq28 Copy Number Gains. Hum. Mutat. 2014, 35, 377–383. [Google Scholar] [CrossRef] [PubMed]
  206. Andersen, E.F.; Baldwin, E.E.; Ellingwood, S.; Smith, R.; Lamb, A.N. Xq28 Duplication Overlapping the Int22h-1/Int22h-2 Region and Including RAB39B and CLIC2 in a Family with Intellectual and Developmental Disability. Am. J. Med. Genet. A 2014, 164A, 1795–1801. [Google Scholar] [CrossRef] [PubMed]
  207. El-Hattab, A.W.; Schaaf, C.P.; Fang, P.; Roeder, E.; Kimonis, V.E.; Church, J.A.; Patel, A.; Cheung, S.W. Clinical Characterization of Int22h1/Int22h2-Mediated Xq28 Duplication/Deletion: New Cases and Literature Review. BMC Med. Genet. 2015, 16, 12. [Google Scholar] [CrossRef] [Green Version]
  208. Ballout, R.A.; Dickerson, C.; Wick, M.J.; Al-Sweel, N.; Openshaw, A.S.; Srivastava, S.; Swanson, L.C.; Bramswig, N.C.; Kuechler, A.; Hong, B.; et al. Int22h1/Int22h2-Mediated Xq28 Duplication Syndrome: De Novo Duplications, Prenatal Diagnoses, and Additional Phenotypic Features. Hum. Mutat. 2020, 41, 1238–1249. [Google Scholar] [CrossRef] [PubMed]
  209. Zerial, M.; McBride, H. Rab Proteins as Membrane Organizers. Nat. Rev. Mol. Cell Biol. 2001, 2, 107–117. [Google Scholar] [CrossRef]
  210. Mignogna, M.L.; Giannandrea, M.; Gurgone, A.; Fanelli, F.; Raimondi, F.; Mapelli, L.; Bassani, S.; Fang, H.; van Anken, E.; Alessio, M.; et al. The Intellectual Disability Protein RAB39B Selectively Regulates GluA2 Trafficking to Determine Synaptic AMPAR Composition. Nat. Commun. 2015, 6, 6504. [Google Scholar] [CrossRef] [Green Version]
  211. Krebs, C.E.; Karkheiran, S.; Powell, J.C.; Cao, M.; Makarov, V.; Darvish, H.; di Paolo, G.; Walker, R.H.; Shahidi, G.A.; Buxbaum, J.D.; et al. The Sac1 Domain of SYNJ1 Identified Mutated in a Family with Early-Onset Progressive Parkinsonism with Generalized Seizures. Hum. Mutat. 2013, 34, 1200–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Dyment, D.A.; Smith, A.C.; Humphreys, P.; Schwartzentruber, J.; Beaulieu, C.L.; Bulman, D.E.; Majewski, J.; Woulfe, J.; Michaud, J.; Boycott, K.M. Homozygous Nonsense Mutation in SYNJ1 Associated with Intractable Epilepsy and Tau Pathology. Neurobiol. Aging 2015, 36, 1222.e1–1222.e5. [Google Scholar] [CrossRef] [PubMed]
  213. Kirola, L.; Behari, M.; Shishir, C.; Thelma, B.K. Identification of a Novel Homozygous Mutation Arg459Pro in SYNJ1 Gene of an Indian Family with Autosomal Recessive Juvenile Parkinsonism. Park. Relat. Disord. 2016, 31, 124–128. [Google Scholar] [CrossRef]
  214. Hardies, K.; Cai, Y.; Jardel, C.; Jansen, A.C.; Cao, M.; May, P.; Djémié, T.; Hachon Le Camus, C.; Keymolen, K.; Deconinck, T.; et al. Loss of SYNJ1 Dual Phosphatase Activity Leads to Early Onset Refractory Seizures and Progressive Neurological Decline. Brain 2016, 139, 2420–2430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Rauschendorf, M.A.; Jost, M.; Stock, F.; Zimmer, A.; Rösler, B.; Rijntjes, M.; Piroth, T.; Coenen, V.A.; Reinacher, P.C.; Meyer, P.T.; et al. Novel Compound Heterozygous Synaptojanin-1 Mutation Causes l-Dopa-Responsive Dystonia-Parkinsonism Syndrome. Mov. Disord. 2017, 32, 478–480. [Google Scholar] [CrossRef] [PubMed]
  216. Taghavi, S.; Chaouni, R.; Tafakhori, A.; Azcona, L.J.; Firouzabadi, S.G.; Omrani, M.D.; Jamshidi, J.; Emamalizadeh, B.; Shahidi, G.A.; Ahmadi, M.; et al. A Clinical and Molecular Genetic Study of 50 Families with Autosomal Recessive Parkinsonism Revealed Known and Novel Gene Mutations. Mol. Neurobiol. 2018, 55, 3477–3489. [Google Scholar] [CrossRef]
  217. ben Romdhan, S.; Sakka, S.; Farhat, N.; Triki, S.; Dammak, M.; Mhiri, C. A Novel SYNJ1 Mutation in a Tunisian Family with Juvenile Parkinson’s Disease Associated with Epilepsy. J. Mol. Neurosci. 2018, 66, 273–278. [Google Scholar] [CrossRef]
  218. al Zaabi, N.; al Menhali, N.; Al-Jasmi, F. SYNJ1 Gene Associated with Neonatal Onset of Neurodegenerative Disorder and Intractable Seizure. Mol. Genet. Genom. Med. 2018, 6, 109–113. [Google Scholar] [CrossRef] [Green Version]
  219. Hong, D.; Cong, L.; Zhong, S.; He, Y.; Xin, L.; Gao, X.; Zhang, J. Clonazepam Improves the Symptoms of Two Siblings with Novel Variants in the SYNJ1 Gene. Park. Relat. Disord. 2019, 62, 221–225. [Google Scholar] [CrossRef]
  220. Xie, F.; Chen, S.; Cen, Z.; Chen, Y.; Yang, D.; Wang, H.; Zhang, B.; Luo, W. A Novel Homozygous SYNJ1 Mutation in Two Siblings with Typical Parkinson’s Disease. Park. Relat. Disord. 2019, 69, 134–137. [Google Scholar] [CrossRef]
  221. Kumar, S.; Yadav, N.; Pandey, S.; Muthane, U.B.; Govindappa, S.T.; Abbas, M.M.; Behari, M.; Thelma, B.K. Novel and Reported Variants in Parkinson’s Disease Genes Confer High Disease Burden among Indians. Park. Relat. Disord. 2020, 78, 46–52. [Google Scholar] [CrossRef]
  222. Samanta, D.; Arya, K. Electroclinical Findings of SYNJ1 Epileptic Encephalopathy. J. Pediatr. Neurosci. 2020, 15, 29–33. [Google Scholar] [PubMed]
  223. Lesage, S.; Mangone, G.; Tesson, C.; Bertrand, H.; Benmahdjoub, M.; Kesraoui, S.; Arezki, M.; Singleton, A.; Corvol, J.C.; Brice, A. Clinical Variability of SYNJ1-Associated Early-Onset Parkinsonism. Front. Neurol. 2021, 12, 648457. [Google Scholar] [CrossRef]
  224. Maj, M.; Taylor, C.L.; Landau, K.; Toriello, H.V.; Li, D.; Bhoj, E.J.; Hakonarson, H.; Nelson, B.; Gluschitz, S.; Walker, R.H.; et al. A Novel SYNJ1 Homozygous Variant Causing Developmental and Epileptic Encephalopathy in an Afro-Caribbean Individual. Mol. Genet. Genom. Med. 2022, 11, e2064. [Google Scholar] [CrossRef]
  225. McPherson, P.S.; Garcia, E.P.; Slepnev, V.I.; David, C.; Zhang, X.; Grabs, D.; Sossin, W.S.; Bauerfeind, R.; Nemoto, Y.; de Camilli, P. A Presynaptic Inositol-5-Phosphatase. Nature 1996, 379, 353–357. [Google Scholar] [CrossRef]
  226. Guo, S.; Stolz, L.E.; Lemrow, S.M.; York, J.D. SAC1-like Domains of Yeast SAC1, INP52, and INP53 and of Human Synaptojanin Encode Polyphosphoinositide Phosphatases. J. Biol. Chem. 1999, 274, 12990–12995. [Google Scholar] [CrossRef] [Green Version]
  227. Hughes, W.E.; Cooke, F.T.; Parker, P.J. Sac Phosphatase Domain Proteins. Biochem. J. 2000, 350 Pt 2, 337–352. [Google Scholar] [PubMed]
  228. Cao, M.; Wu, Y.; Ashrafi, G.; McCartney, A.J.; Wheeler, H.; Bushong, E.A.; Boassa, D.; Ellisman, M.H.; Ryan, T.A.; de Camilli, P. Parkinson Sac Domain Mutation in Synaptojanin 1 Impairs Clathrin Uncoating at Synapses and Triggers Dystrophic Changes in Dopaminergic Axons. Neuron 2017, 93, 882–896.e5. [Google Scholar] [CrossRef] [Green Version]
  229. Fasano, D.; Parisi, S.; Pierantoni, G.M.; de Rosa, A.; Picillo, M.; Amodio, G.; Pellecchia, M.T.; Barone, P.; Moltedo, O.; Bonifati, V.; et al. Alteration of Endosomal Trafficking Is Associated with Early-Onset Parkinsonism Caused by SYNJ1 Mutations. Cell Death Dis. 2018, 9, 1–15. [Google Scholar] [CrossRef] [PubMed]
  230. Drouet, V.; Lesage, S. Synaptojanin 1 Mutation in Parkinson’s Disease Brings Further Insight into the Neuropathological Mechanisms. Biomed. Res. Int. 2014, 2014, 1–9. [Google Scholar] [CrossRef] [Green Version]
  231. Vanhauwaert, R.; Kuenen, S.; Masius, R.; Bademosi, A.; Manetsberger, J.; Schoovaerts, N.; Bounti, L.; Gontcharenko, S.; Swerts, J.; Vilain, S.; et al. The SAC1 Domain in Synaptojanin Is Required for Autophagosome Maturation at Presynaptic Terminals. EMBO J. 2017, 36, 1392–1411. [Google Scholar] [CrossRef] [PubMed]
  232. Choudhry, H.; Aggarwal, M.; Pan, P.-Y. Mini-Review: Synaptojanin 1 and Its Implications in Membrane Trafficking. Neurosci. Lett. 2021, 765, 136288. [Google Scholar] [CrossRef]
  233. Köroğlu, Ç.; Baysal, L.; Cetinkaya, M.; Karasoy, H.; Tolun, A. DNAJC6 Is Responsible for Juvenile Parkinsonism with Phenotypic Variability. Park. Relat. Disord. 2013, 19, 320–324. [Google Scholar] [CrossRef] [PubMed]
  234. Olgiati, S.; Quadri, M.; Fang, M.; Rood, J.P.M.A.; Saute, J.A.; Chien, H.F.; Bouwkamp, C.G.; Graafland, J.; Minneboo, M.; Breedveld, G.J.; et al. DNAJC6 Mutations Associated With Early-Onset Parkinson’s Disease. Ann. Neurol. 2016, 79, 244–256. [Google Scholar] [CrossRef] [PubMed]
  235. Elsayed, L.E.O.; Drouet, V.; Usenko, T.; Mohammed, I.N.; Hamed, A.A.A.; Elseed, M.A.; Salih, M.A.M.; Koko, M.E.; Mohamed, A.Y.O.; Siddig, R.A.; et al. A Novel Nonsense Mutation in DNAJC6 Expands the Phenotype of Autosomal-Recessive Juvenile-Onset Parkinson’s Disease. Ann. Neurol. 2016, 79, 335–337. [Google Scholar] [CrossRef] [PubMed]
  236. Ng, J.; Cortès-Saladelafont, E.; Abela, L.; Termsarasab, P.; Mankad, K.; Sudhakar, S.; Gorman, K.M.; Heales, S.J.R.; Pope, S.; Biassoni, L.; et al. DNAJC6 Mutations Disrupt Dopamine Homeostasis in Juvenile Parkinsonism-Dystonia. Mov. Disord. 2020, 35, 1357–1368. [Google Scholar] [CrossRef]
  237. Mittal, S.O. Levodopa Responsive-Generalized Dystonic Spells and Moaning in DNAJC6 Related Juvenile Parkinson’s Disease. Park. Relat. Disord. 2020, 81, 188–189. [Google Scholar] [CrossRef]
  238. Ray, S.; Padmanabha, H.; Mahale, R.; Mailankody, P.; Arunachal, G. DNAJC6 Mutation Causing Cranial-Onset Dystonia with Tremor Dominant Levodopa Non-Responsive Parkinsonism: A Novel Phenotype. Park. Relat. Disord. 2021, 89, 1–3. [Google Scholar] [CrossRef]
  239. Garza-Brambila, D.; Esparza-Hernández, C.N.; Ramirez-Zenteno, J.; Martinez-Ramirez, D. Juvenile Dystonia-Parkinsonism Due to DNAJC6 Mutation. Mov. Disord. Clin. Pract. 2021, 8, S26–S28. [Google Scholar] [CrossRef]
  240. Wittke, C.; Petkovic, S.; Dobricic, V.; Schaake, S.; Arzberger, T.; Compta, Y.; Englund, E.; Ferguson, L.W.; Gelpi, E.; Roeber, S.; et al. Genotype–Phenotype Relations for the Atypical Parkinsonism Genes: MDSGene Systematic Review. Mov. Disord. 2021, 36, 1499–1510. [Google Scholar] [CrossRef]
  241. Yim, Y.I.; Sun, T.; Wu, L.G.; Raimondi, A.; de Camilli, P.; Eisenberg, E.; Greene, L.E. Endocytosis and Clathrin-Uncoating Defects at Synapses of Auxilin Knockout Mice. Proc. Natl. Acad. Sci. USA 2010, 107, 4412–4417. [Google Scholar] [CrossRef] [Green Version]
  242. Vidyadhara, D.J.; Lee, J.E.; Chandra, S.S. Role of the Endolysosomal System in Parkinson’s Disease. J. Neurochem. 2019, 150, 487–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Wulansari, N.; Darsono, W.H.W.; Woo, H.; Chang, M.; Kim, J.; Bae, E.; Sun, W.; Lee, J.; Cho, I.; Shin, H.; et al. Neurodevelopmental Defects and Neurodegenerative Phenotypes in Human Brain Organoids Carrying Parkinson’s Disease-Linked DNAJC6 Mutations. Sci. Adv. 2021, 7, eabb1540. [Google Scholar] [CrossRef] [PubMed]
  244. Foo, J.N.; Chew, E.G.Y.; Chung, S.J.; Peng, R.; Blauwendraat, C.; Nalls, M.A.; Mok, K.Y.; Satake, W.; Toda, T.; Chao, Y.; et al. Identification of Risk Loci for Parkinson Disease in Asians and Comparison of Risk between Asians and Europeans: A Genome-Wide Association Study. JAMA Neurol. 2020, 77, 746–754. [Google Scholar] [CrossRef] [PubMed]
  245. Liu, H.; Dehestani, M.; Blauwendraat, C.; Makarious, M.B.; Leonard, H.; Kim, J.J.; Schulte, C.; Noyce, A.; Jacobs, B.M.; Foote, I.; et al. Polygenic Resilience Modulates the Penetrance of Parkinson Disease Genetic Risk Factors. Ann. Neurol. 2022, 92, 270–278. [Google Scholar] [CrossRef]
  246. Bandres-Ciga, S.; Saez-Atienzar, S.; Kim, J.J.; Makarious, M.B.; Faghri, F.; Diez-Fairen, M.; Iwaki, H.; Leonard, H.; Botia, J.; Ryten, M.; et al. Large-Scale Pathway Specific Polygenic Risk and Transcriptomic Community Network Analysis Identifies Novel Functional Pathways in Parkinson Disease. Acta Neuropathol. 2020, 140, 341–358. [Google Scholar] [CrossRef]
  247. Bottigliengo, D.; Foco, L.; Seibler, P.; Klein, C.; König, I.R.; Del Greco M, F. A Mendelian Randomization Study Investigating the Causal Role of Inflammation on Parkinson’s Disease. Brain 2022, 145, 3444–3453. [Google Scholar] [CrossRef]
  248. Sidransky, E. Gaucher Disease: Complexity in a “Simple” Disorder. Mol. Genet. Metab. 2004, 83, 6–15. [Google Scholar] [CrossRef]
  249. Mata, I.F.; Samii, A.; Schneer, S.H.; Roberts, J.W.; Griffith, A.; Leis, B.C.; Schellenberg, G.D.; Sidransky, E.; Bird, T.D.; Leverenz, J.B.; et al. Glucocerebrosidase Gene Mutations. Arch. Neurol. 2008, 65, 379–382. [Google Scholar] [CrossRef] [Green Version]
  250. Sidransky, E.; Nalls, M.A.; Aasly, J.O.; Aharon-Peretz, J.; Annesi, G.; Barbosa, E.R.; Bar-Shira, A.; Berg, D.; Bras, J.; Brice, A.; et al. Multicenter Analysis of Glucocerebrosidase Mutations in Parkinson’s Disease. New Engl. J. Med. 2009, 361, 1651–1661. [Google Scholar] [CrossRef] [Green Version]
  251. Rana, H.Q.; Balwani, M.; Bier, L.; Alcalay, R.N. Age-Specific Parkinson Disease Risk in GBA Mutation Carriers: Information for Genetic Counseling. Genet. Med. 2013, 15, 146–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Asselta, R.; Rimoldi, V.; Siri, C.; Cilia, R.; Guella, I.; Tesei, S.; Soldà, G.; Pezzoli, G.; Duga, S.; Goldwurm, S. Glucocerebrosidase Mutations in Primary Parkinsonism. Park. Relat. Disord. 2014, 20, 1215–1220. [Google Scholar] [CrossRef] [Green Version]
  253. Chen, J.; Li, W.; Zhang, T.; Wang, Y.J.; Jiang, X.J.; Xu, Z.Q.; Chiba-Falek, O. Glucocerebrosidase Gene Mutations Associated with Parkinson’s Disease: A Meta-Analysis in a Chinese Population. PLoS ONE 2014, 9, e115747. [Google Scholar] [CrossRef] [Green Version]
  254. Robak, L.A.; Jansen, I.E.; van Rooij, J.; Uitterlinden, A.G.; Kraaij, R.; Jankovic, J.; Heutink, P.; Shulman, J.M.; Nalls, M.A.; Plagnol, V.; et al. Excessive Burden of Lysosomal Storage Disorder Gene Variants in Parkinson’s Disease. Brain 2017, 140, 3191–3203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Balestrino, R.; Tunesi, S.; Tesei, S.; Lopiano, L.; Zecchinelli, A.L.; Goldwurm, S. Penetrance of Glucocerebrosidase (GBA) Mutations in Parkinson’s Disease: A Kin Cohort Study. Mov. Disord. 2020, 35, 2111–2114. [Google Scholar] [CrossRef]
  256. Smith, L.; Schapira, A.H.V. GBA Variants and Parkinson Disease: Mechanisms and Treatments. Cells 2022, 11, 1261. [Google Scholar] [CrossRef]
  257. Horowitz, M.; Wilder, S.; Horowitz, Z.; Reiner, O.; Gelbart, T.; Beutler, E. The Human Glucocerebrosidase Gene and Pseudogene: Structure and Evolution. Genomics 1989, 4, 87–96. [Google Scholar] [CrossRef]
  258. Dvir, H.; Harel, M.; McCarthy, A.A.; Toker, L.; Silman, I.; Futerman, A.H.; Sussman, J.L. X-Ray Structure of Human Acid-β-Glucosidase, the Defective Enzyme in Gaucher Disease. EMBO Rep. 2003, 4, 704–709. [Google Scholar]
  259. Futerman, A.H.; Zimran, A. Gaucher Disease; CRC Press: Boca Raton, FL, USA, 2006; ISBN 0429114559. [Google Scholar]
  260. McKeran, R.O.; Bradbury, P.; Taylor, D.; Stern, G. Neurological Involvement in Type 1 (Adult) Gaucher’s Disease. J. Neurol. Neurosurg. Psychiatry 1985, 48, 172–175. [Google Scholar] [CrossRef]
  261. Turpin, J.C.; Dubois, G.; Brice, A.; Masson, M.; Nadaud, M.C.; Boutry, J.M.; Schram, A.W.; Tager, J.M.; Baumann, N. Parkinsonian Symptomatology in a Patient with Type I (Adult) Gaucher’s Disease. In Proceedings of the Lipid Storage Disorders; Salvayre, R., Douste-Blazy, L., Gatt, S., Eds.; Springer: Boston, MA, USA, 1988; pp. 103–105. [Google Scholar]
  262. Neudorfer, O.; Giladi, N.; Elstein, D.; Abrahamov, A.; Turezkite, T.; Achai, E.; Reches, A.; Bembi, B.; Zimran, A. Occurrence of Parkinson’s Syndrome in Type I Gaucher Disease. Int. J. Med. 1996, 89, 691–694. [Google Scholar] [CrossRef]
  263. Machaczka, M.; Rucinska, M.; Skotnicki, A.B.; Jurczak, W. Parkinson’s Syndrome Preceding Clinical Manifestation of Gaucher’s Disease. Am. J. Hematol. 1999, 61, 216–217. [Google Scholar] [CrossRef]
  264. Tayebi, N.; Walker, J.; Stubblefield, B.; Orvisky, E.; LaMarca, M.E.; Wong, K.; Rosenbaum, H.; Schiffmann, R.; Bembi, B.; Sidransky, E. Gaucher Disease with Parkinsonian Manifestations: Does Glucocerebrosidase Deficiency Contribute to a Vulnerability to Parkinsonism? Mol. Genet. Metab. 2003, 79, 104–109. [Google Scholar] [CrossRef] [PubMed]
  265. Aharon-Peretz, J.; Badarny, S.; Rosenbaum, H.; Gershoni-Baruch, R. Mutations in the Glucocerebrosidase Gene and Parkinson Disease: Phenotype-Genotype Correlation. Neurology 2005, 65, 1460–1461. [Google Scholar] [CrossRef] [PubMed]
  266. Gan-Or, Z.; Bar-Shira, A.; Mirelman, A.; Gurevich, T.; Kedmi, M.; Giladi, N.; Orr-Urtreger, A. LRRK2 and GBA Mutations Differentially Affect the Initial Presentation of Parkinson Disease. Neurogenetics 2010, 11, 121–125. [Google Scholar] [CrossRef] [PubMed]
  267. Lesage, S.; Anheim, M.; Condroyer, C.; Pollak, P.; Durif, F.; Dupuits, C.; Viallet, F.; Lohmann, E.; Corvol, J.C.; Honoré, A.; et al. Large-Scale Screening of the Gaucher’s Disease-Related Glucocerebrosidase Gene in Europeans with Parkinson’s Disease. Hum. Mol. Genet. 2011, 20, 202–210. [Google Scholar] [CrossRef]
  268. Zhang, Y.; Sun, Q.Y.; Zhao, Y.W.; Shu, L.; Guo, J.F.; Xu, Q.; Yan, X.X.; Tang, B.S. Effect of GBA Mutations on Phenotype of Parkinson’s Disease: A Study on Chinese Population and a Meta-Analysis. Park. Dis. 2015, 2015, 916971. [Google Scholar] [CrossRef] [Green Version]
  269. Vieira, S.R.L.; Schapira, A.H.V. Glucocerebrosidase Mutations and Parkinson Disease. J. Neural. Transm. 2022, 129, 1105–1117. [Google Scholar] [CrossRef]
  270. Gan-Or, Z.; Giladi, N.; Rozovski, U.; Shifrin, C.; Rosner, S.; Gurevich, T.; Bar-Shira, A.; Orr-Urtreger, A. Genotype-Phenotype Correlations between Gba Mutations and Parkinson Disease Risk and Onsetsymbol. Neurology 2008, 70, 2277–2283. [Google Scholar] [CrossRef]
  271. Neumann, J.; Bras, J.; Deas, E.; O’sullivan, S.S.; Parkkinen, L.; Lachmann, R.H.; Li, A.; Holton, J.; Guerreiro, R.; Paudel, R.; et al. Glucocerebrosidase Mutations in Clinical and Pathologically Proven Parkinson’s Disease. Brain 2009, 132, 1783–1794. [Google Scholar] [CrossRef] [Green Version]
  272. Hu, F.Y.; Xi, J.; Guo, J.; Yu, L.H.; Liu, L.; He, X.H.; Liu, Z.L.; Zou, X.Y.; Xu, Y.M. Association of the Glucocerebrosidase N370S Allele with Parkinson’s Disease in Two Separate Chinese Han Populations of Mainland China. Eur. J. Neurol. 2010, 17, 1476–1478. [Google Scholar] [CrossRef]
  273. Winder-Rhodes, S.E.; Evans, J.R.; Ban, M.; Mason, S.L.; Williams-Gray, C.H.; Foltynie, T.; Duran, R.; Mencacci, N.E.; Sawcer, S.J.; Barker, R.A. Glucocerebrosidase Mutations Influence the Natural History of Parkinson’s Disease in a Community-Based Incident Cohort. Brain 2013, 136, 392–399. [Google Scholar] [CrossRef] [PubMed]
  274. Brockmann, K.; Srulijes, K.; Pflederer, S.; Hauser, A.K.; Schulte, C.; Maetzler, W.; Gasser, T.; Berg, D. GBA-Associated Parkinson’s Disease: Reduced Survival and More Rapid Progression in a Prospective Longitudinal Study. Mov. Disord. 2015, 30, 407–411. [Google Scholar] [CrossRef] [PubMed]
  275. Malek, N.; Weil, R.S.; Bresner, C.; Lawton, M.A.; Grosset, K.A.; Tan, M.; Bajaj, N.; Barker, R.A.; Burn, D.J.; Foltynie, T.; et al. Features of GBA-Associated Parkinson’s Disease at Presentation in the UK Tracking Parkinson’s Study. J. Neurol. Neurosurg. Psychiatry 2018, 89, 702–709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Zhang, Y.; Shu, L.; Zhou, X.; Pan, H.; Xu, Q.; Guo, J.; Tang, B.; Sun, Q. A Meta-Analysis of GBA-Related Clinical Symptoms in Parkinson’s Disease. Park. Dis. 2018, 2018, 3136415. [Google Scholar] [CrossRef] [Green Version]
  277. Brockmann, K.; Srulijes, K.; Hauser, A.K.; Schulte, C.; Csoti, I.; Gasser, T.; Berg, D. GBA-Associated PD Presents with Nonmotor Characteristics. Neurology 2011, 77, 276–280. [Google Scholar] [CrossRef]
  278. McNeill, A.; Duran, R.; Hughes, D.A.; Mehta, A.; Schapira, A.H.V. A Clinical and Family History Study of Parkinson’s Disease in Heterozygous Glucocerebrosidase Mutation Carriers. J. Neurol. Neurosurg. Psychiatry 2012, 83, 853–854. [Google Scholar] [CrossRef] [Green Version]
  279. Alcalay, R.N.; Caccappolo, E.; Mejia-Santana, H.; Tang, M.X.; Rosado, L.; Reilly, M.O.; Ruiz, D.; Ross, B.; Verbitsky, M.; Kisselev, S.; et al. Cognitive Performance of GBA Mutation Carriers with Early-Onset PD The CORE-PD Study. Neurology 2012, 78, 1434–1440. [Google Scholar] [CrossRef] [Green Version]
  280. Zokaei, N.; Mcneill, A.; Proukakis, C.; Beavan, M.; Jarman, P.; Korlipara, P.; Hughes, D.; Mehta, A.; Hu, M.T.M.; Schapira, A.H.V.; et al. Visual Short-Term Memory Deficits Associated with GBA Mutation and Parkinson’s Disease. Brain 2014, 137, 2303–2311. [Google Scholar] [CrossRef] [Green Version]
  281. Wang, C.; Cai, Y.; Gu, Z.; Ma, J.; Zheng, Z.; Tang, B.S.; Xu, Y.; Zhou, Y.; Feng, T.; Wang, T.; et al. Clinical Profiles of Parkinson’s Disease Associated with Common Leucine-Rich Repeat Kinase 2 and Glucocerebrosidase Genetic Variants in Chinese Individuals. Neurobiol. Aging 2014, 35, 725.e1–725.e6. [Google Scholar] [CrossRef]
  282. Petrucci, S.; Ginevrino, M.; Trezzi, I.; Monfrini, E.; Ricciardi, L.; Albanese, A.; Avenali, M.; Barone, P.; Bentivoglio, A.R.; Bonifati, V.; et al. GBA-Related Parkinson’s Disease: Dissection of Genotype–Phenotype Correlates in a Large Italian Cohort. Mov. Disord. 2020, 35, 2106–2111. [Google Scholar] [CrossRef]
  283. Leocadi, M.; Canu, E.; Donzuso, G.; Stojkovic, T.; Basaia, S.; Kresojević, N.; Stankovic, I.; Sarasso, E.; Piramide, N.; Tomic, A.; et al. Longitudinal Clinical, Cognitive, and Neuroanatomical Changes over 5 Years in GBA-Positive Parkinson’s Disease Patients. J. Neurol. 2022, 269, 1485–1500. [Google Scholar] [CrossRef] [PubMed]
  284. Wong, K.; Sidransky, E.; Verma, A.; Mixon, T.; Sandberg, G.D.; Wakefield, L.K.; Morrison, A.; Lwin, A.; Colegial, C.; Allman, J.M.; et al. Neuropathology Provides Clues to the Pathophysiology of Gaucher Disease. Mol. Genet. Metab. 2004, 82, 192–207. [Google Scholar] [CrossRef] [PubMed]
  285. Clark, L.N.; Kartsaklis, L.A.; Gilbert, R.W.; Dorado, B.; Ross, B.M.; Kisselev, S.; Verbitsky, M.; Mejia-Santana, H.; Cote, L.J.; Andrews, H.; et al. Association of Glucocerebrosidase Mutations with Dementia with Lewy Bodies. Arch. Neurol. 2009, 66, 578–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Westbroek, W.; Gustafson, A.M.; Sidransky, E. Exploring the Link between Glucocerebrosidase Mutations and Parkinsonism. Trends Mol. Med. 2011, 17, 485–493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Choi, J.H.; Stubblefield, B.; Cookson, M.R.; Goldin, E.; Velayati, A.; Tayebi, N.; Sidransky, E. Aggregation of α-Synuclein in Brain Samples from Subjects with Glucocerebrosidase Mutations. Mol. Genet. Metab. 2011, 104, 185–188. [Google Scholar] [CrossRef] [Green Version]
  288. Jesús, S.; Huertas, I.; Bernal-Bernal, I.; Bonilla-Toribio, M.; Cáceres-Redondo, M.T.; Vargas-González, L.; Gómez-Llamas, M.; Carrillo, F.; Calderón, E.; Carballo, M.; et al. GBA Variants Influence Motor and Non-Motor Features of Parkinson’s Disease. PLoS ONE 2016, 11, e0167749. [Google Scholar] [CrossRef] [Green Version]
  289. Winfield, S.L.; Tayebi, N.; Martin, B.M.; Ginns, E.I.; Sidransky, E. Identification of Three Additional Genes Contiguous to the Glucocerebrosidase Locus on Chromosome 1q21: Implications for Gaucher Disease. Genome. Res. 1997, 7, 1020–1026. [Google Scholar] [CrossRef] [Green Version]
  290. Chabás, A.; Gort, L.; Díaz-Font, A.; Montfort, M.; Santamaría, R.; Cidrás, M.; Grinberg, D.; Vilageliu, L. Perinatal Lethal Phenotype with Generalized Ichthyosis in a Type 2 Gaucher Disease Patient with the [L444P;E326K]/P182L Genotype: Effect of the E326K Change in Neonatal and Classic Forms of the Disease. Blood Cells Mol. Dis. 2005, 35, 253–258. [Google Scholar] [CrossRef]
  291. Hruska, K.S.; LaMarca, M.E.; Scott, C.R.; Sidransky, E. Gaucher Disease: Mutation and Polymorphism Spectrum in the Glucocerebrosidase Gene (GBA). Hum. Mutat. 2008, 29, 567–583. [Google Scholar] [CrossRef]
  292. Liou, B.; Grabowski, G.A. Is E326K Glucocerebrosidase a Polymorphic or Pathological Variant? Mol. Genet. Metab. 2012, 105, 528–529. [Google Scholar] [CrossRef]
  293. Duran, R.; Mencacci, N.E.; Angeli, A.V.; Shoai, M.; Deas, E.; Houlden, H.; Mehta, A.; Hughes, D.; Cox, T.M.; Deegan, P.; et al. The Glucocerobrosidase E326K Variant Predisposes to Parkinson’s Disease, but Does Not Cause Gaucher’s Disease. Mov. Disord. 2013, 28, 232–236. [Google Scholar] [CrossRef] [Green Version]
  294. Dandana, A.; ben Khelifa, S.; Chahed, H.; Miled, A.; Ferchichi, S. Gaucher Disease: Clinical, Biological and Therapeutic Aspects. Pathobiology 2016, 83, 13–23. [Google Scholar] [CrossRef] [PubMed]
  295. Cilia, R.; Tunesi, S.; Marotta, G.; Cereda, E.; Siri, C.; Tesei, S.; Zecchinelli, A.L.; Canesi, M.; Mariani, C.B.; Meucci, N.; et al. Survival and Dementia in GBA-Associated Parkinson’s Disease: The Mutation Matters. Ann. Neurol. 2016, 80, 662–673. [Google Scholar] [CrossRef]
  296. Gan-Or, Z.; Amshalom, I.; Kilarski, L.L.; Bar-Shira, A.; Gana-Weisz, M.; Mirelman, A.; Marder, K.; Bressman, S.; Giladi, N.; Orr-Urtreger, A. Differential Effects of Severe vs Mild GBA Mutations on Parkinson Disease. Neurology 2015, 84, 880–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Huh, Y.E.; Chiang, M.S.R.; Locascio, J.J.; Liao, Z.; Liu, G.; Choudhury, K.; Kuras, Y.I.; Tuncali, I.; Videnovic, A.; Hunt, A.L.; et al. β-Glucocerebrosidase Activity in GBA-Linked Parkinson Disease. Neurology 2020, 95, e685–e696. [Google Scholar] [CrossRef] [PubMed]
  298. Horowitz, M.; Pasmanik-Chor, M.; Ron, I.; Kolodny, E.H. The Enigma of the E326K Mutation in Acid β-Glucocerebrosidase. Mol. Genet. Metab. 2011, 104, 35–38. [Google Scholar] [CrossRef] [PubMed]
  299. Sidransky, E.; Lopez, G. The Link between the GBA Gene and Parkinsonism. Lancet Neurol. 2012, 11, 986–998. [Google Scholar] [CrossRef] [Green Version]
  300. Goker-Alpan, O.; Stubblefield, B.K.; Giasson, B.I.; Sidransky, E. Glucocerebrosidase Is Present in α-Synuclein Inclusions in Lewy Body Disorders. Acta Neuropathol. 2010, 120, 641–649. [Google Scholar] [CrossRef] [Green Version]
  301. Horowitz, M.; Braunstein, H.; Zimran, A.; Revel-Vilk, S.; Goker-Alpan, O. Lysosomal Functions and Dysfunctions: Molecular and Cellular Mechanisms Underlying Gaucher Disease and Its Association with Parkinson Disease. Adv. Drug Deliv. Rev. 2022, 187, 114402. [Google Scholar] [CrossRef]
  302. Cooper, A.A.; Gitler, A.D.; Cashikar, A.; Haynes, C.M.; Hill, K.J.; Bhullar, B.; Liu, K.; Xu, K.; Strathearn, K.E.; Liu, F.; et al. α-Synuclein Blocks ER-Golgi Traffic and Rab1 Rescues Neuron Loss in Parkinson’s Models. Science 2006, 313, 324–328. [Google Scholar] [CrossRef] [Green Version]
  303. Mazzulli, J.R.; Xu, Y.H.; Sun, Y.; Knight, A.L.; McLean, P.J.; Caldwell, G.A.; Sidransky, E.; Grabowski, G.A.; Krainc, D. Gaucher Disease Glucocerebrosidase and α-Synuclein Form a Bidirectional Pathogenic Loop in Synucleinopathies. Cell 2011, 146, 37–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Cullen, V.; Sardi, S.P.; Ng, J.; Xu, Y.H.; Sun, Y.; Tomlinson, J.J.; Kolodziej, P.; Kahn, I.; Saftig, P.; Woulfe, J.; et al. Acid β-Glucosidase Mutants Linked to Gaucher Disease, Parkinson Disease, and Lewy Body Dementia Alter α-Synuclein Processing. Ann. Neurol. 2011, 69, 940–953. [Google Scholar] [CrossRef]
  305. Kurzawa-Akanbi, M.; Hanson, P.S.; Blain, P.G.; Lett, D.J.; McKeith, I.G.; Chinnery, P.F.; Morris, C.M. Glucocerebrosidase Mutations Alter the Endoplasmic Reticulum and Lysosomes in Lewy Body Disease. J. Neurochem. 2012, 123, 298–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Cleeter, M.W.J.; Chau, K.Y.; Gluck, C.; Mehta, A.; Hughes, D.A.; Duchen, M.; Wood, N.W.; Hardy, J.; Mark Cooper, J.; Schapira, A.H. Glucocerebrosidase Inhibition Causes Mitochondrial Dysfunction and Free Radical Damage. Neurochem. Int. 2013, 62, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Rocha, E.M.; Smith, G.A.; Park, E.; Cao, H.; Graham, A.R.; Brown, E.; McLean, J.R.; Hayes, M.A.; Beagan, J.; Izen, S.C.; et al. Sustained Systemic Glucocerebrosidase Inhibition Induces Brain α-Synuclein Aggregation, Microglia and Complement C1q Activation in Mice. Antioxid. Redox Signal. 2015, 23, 550–564. [Google Scholar] [CrossRef] [Green Version]
  308. Chiasserini, D.; Paciotti, S.; Eusebi, P.; Persichetti, E.; Tasegian, A.; Kurzawa-Akanbi, M.; Chinnery, P.F.; Morris, C.M.; Calabresi, P.; Parnetti, L.; et al. Selective Loss of Glucocerebrosidase Activity in Sporadic Parkinson’s Disease and Dementia with Lewy Bodies. Mol. Neurodegener. 2015, 10, 4–9. [Google Scholar] [CrossRef] [Green Version]
  309. Fernandes, H.J.R.; Hartfield, E.M.; Christian, H.C.; Emmanoulidou, E.; Zheng, Y.; Booth, H.; Bogetofte, H.; Lang, C.; Ryan, B.J.; Sardi, S.P.; et al. ER Stress and Autophagic Perturbations Lead to Elevated Extracellular α-Synuclein in GBA-N370S Parkinson’s IPSC-Derived Dopamine Neurons. Stem Cell Rep. 2016, 6, 342–356. [Google Scholar] [CrossRef] [Green Version]
  310. Maor, G.; Cabasso, O.; Krivoruk, O.; Rodriguez, J.; Steller, H.; Segal, D.; Horowitz, M. The Contribution of Mutant GBA to the Development of Parkinson Disease in Drosophila. Hum. Mol. Genet. 2016, 25, ddw129. [Google Scholar] [CrossRef] [Green Version]
  311. Gündner, A.L.; Duran-Pacheco, G.; Zimmermann, S.; Ruf, I.; Moors, T.; Baumann, K.; Jagasia, R.; van de Berg, W.D.J.; Kremer, T. Path Mediation Analysis Reveals GBA Impacts Lewy Body Disease Status by Increasing α-Synuclein Levels. Neurobiol. Dis. 2019, 121, 205–213. [Google Scholar] [CrossRef]
  312. Brunialti, E.; Villa, A.; Mekhaeil, M.; Mornata, F.; Vegeto, E.; Maggi, A.; di Monte, D.A.; Ciana, P. Inhibition of Microglial β-Glucocerebrosidase Hampers the Microglia-Mediated Antioxidant and Protective Response in Neurons. J. Neuroinflam. 2021, 18, 220. [Google Scholar] [CrossRef]
  313. Lunghi, G.; Carsana, E.V.; Loberto, N.; Cioccarelli, L.; Prioni, S.; Mauri, L.; Bassi, R.; Duga, S.; Straniero, L.; Asselta, R.; et al. β-Glucocerebrosidase Deficiency Activates an Aberrant Lysosome-Plasma Membrane Axis Responsible for the Onset of Neurodegeneration. Cells 2022, 11, 2343. [Google Scholar] [CrossRef] [PubMed]
  314. Straniero, L.; Asselta, R.; Bonvegna, S.; Rimoldi, V.; Melistaccio, G.; Soldà, G.; Aureli, M.; della Porta, M.; Lucca, U.; di Fonzo, A.; et al. The SPID-GBA Study: Sex Distribution, Penetrance, Incidence, and Dementia in GBA-PD. Neurol. Genet. 2020, 6, e523. [Google Scholar] [CrossRef] [PubMed]
  315. Thurm, A.; Chlebowski, C.; Joseph, L.; Farmer, C.; Adedipe, D.; Weiss, M.; Wiggs, E.; Farhat, N.; Bianconi, S.; Berry-Kravis, E.; et al. Neurodevelopmental Characterization of Young Children Diagnosed with Niemann-Pick Disease, Type C1. J. Dev. Behav. Pediatr. 2020, 41, 388–396. [Google Scholar] [CrossRef]
  316. Eskes, E.C.B.; Sjouke, B.; Vaz, F.M.; Goorden, S.M.I.; van Kuilenburg, A.B.P.; Aerts, J.M.F.G.; Hollak, C.E.M. Biochemical and Imaging Parameters in Acid Sphingomyelinase Deficiency: Potential Utility as Biomarkers. Mol. Genet. Metab. 2020, 130, 16–26. [Google Scholar] [CrossRef]
  317. Bajwa, H.; Azhar, W. Niemann-Pick Disease; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  318. Gan-Or, Z.; Ozelius, L.J.; Bar-Shira, A.; Saunders-Pullman, R.; Mirelman, A.; Kornreich, R.; Gana-Weisz, M.; Raymond, D.; Rozenkrantz, L.; Deik, A.; et al. The p.L302P Mutation in the Lysosomal Enzyme Gene SMPD1 Is a Risk Factor for Parkinson Disease. Neurology 2013, 80, 1606–1610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Dagan, E.; Schlesinger, I.; Ayoub, M.; Mory, A.; Nassar, M.; Kurolap, A.; Peretz-Aharon, J.; Gershoni-Baruch, R. The Contribution of Niemann-Pick SMPD1 Mutations to Parkinson Disease in Ashkenazi Jews. Park. Relat. Disord. 2015, 21, 1067–1071. [Google Scholar] [CrossRef] [PubMed]
  320. Dagan, E.; Adir, V.; Schlesinger, I.; Borochowitz, Z.; Ayoub, M.; Mory, A.; Nassar, M.; Kurolap, A.; Aharon-Peretz, J.; Gershoni-Baruch, R. SMPD1 Mutations and Parkinson Disease. Park. Relat. Disord. 2015, 21, 1296–1297. [Google Scholar] [CrossRef] [PubMed]
  321. Gan-Or, Z.; Orr-Urtreger, A.; Alcalay, R.N.; Bressman, S.; Giladi, N.; Rouleau, G.A. The Emerging Role of SMPD1 Mutations in Parkinson’s Disease: Implications for Future Studies. Park. Relat. Disord. 2015, 21, 1294–1295. [Google Scholar] [CrossRef]
  322. Clark, L.N.; Chan, R.; Cheng, R.; Liu, X.; Park, N.; Parmalee, N.; Kisselev, S.; Cortes, E.; Torres, P.A.; Pastores, G.M.; et al. Gene-Wise Association of Variants in Four Lysosomal Storage Disorder Genes in Neuropathologically Confirmed Lewy Body Disease. PLoS ONE 2015, 10, e0125204. [Google Scholar] [CrossRef] [Green Version]
  323. Alcalay, R.N.; Mallett, V.; Vanderperre, B.; Tavassoly, O.; Dauvilliers, Y.; Wu, R.Y.J.; Ruskey, J.A.; Leblond, C.S.; Ambalavanan, A.; Laurent, S.B.; et al. SMPD1 Mutations, Activity, and α-Synuclein Accumulation in Parkinson’s Disease. Mov. Disord. 2019, 34, 526–535. [Google Scholar] [CrossRef]
  324. Foo, J.N.; Liany, H.; Bei, J.X.; Yu, X.Q.; Liu, J.; Au, W.L.; Prakash, K.M.; Tan, L.C.; Tan, E.K. A Rare Lysosomal Enzyme Gene SMPD1 Variant (p.R591C) Associates with Parkinson’s Disease. Neurobiol. Aging 2013, 34, 2890.e13–2890.e15. [Google Scholar] [CrossRef]
  325. Deng, S.; Deng, X.; Song, Z.; Xiu, X.; Guo, Y.; Xiao, J.; Deng, H. Systematic Genetic Analysis of the SMPD1 Gene in Chinese Patients with Parkinson’s Disease. Mol. Neurobiol. 2016, 53, 5025–5029. [Google Scholar] [CrossRef] [PubMed]
  326. Ylönen, S.; Siitonen, A.; Nalls, M.A.; Ylikotila, P.; Autere, J.; Eerola-Rautio, J.; Gibbs, R.; Hiltunen, M.; Tienari, P.J.; Soininen, H.; et al. Genetic Risk Factors in Finnish Patients with Parkinson’s Disease. Park. Relat. Disord. 2017, 45, 39–43. [Google Scholar] [CrossRef] [Green Version]
  327. Jinn, S.; Drolet, R.E.; Cramer, P.E.; Wong, A.H.K.; Toolan, D.M.; Gretzula, C.A.; Voleti, B.; Vassileva, G.; Disa, J.; Tadin-Strapps, M.; et al. TMEM175 Deficiency Impairs Lysosomal and Mitochondrial Function and Increases α-Synuclein Aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 2389–2394. [Google Scholar] [CrossRef] [Green Version]
  328. Perdigoto, C.N. Parkinson’s Disease Risk Protein TMEM175 Keeps Lysosomes Running on a Proton Leak. Nat. Struct. Mol. Biol. 2022, 29, 626. [Google Scholar] [CrossRef]
  329. Jinn, S.; Blauwendraat, C.; Toolan, D.; Gretzula, C.A.; Drolet, R.E.; Smith, S.; Nalls, M.A.; Marcus, J.; Singleton, A.B.; Stone, D.J. Functionalization of the TMEM175 p.M393T Variant as a Risk Factor for Parkinson Disease. Hum. Mol. Genet. 2019, 28, 3244–3254. [Google Scholar] [CrossRef] [Green Version]
  330. Krohn, L.; Öztürk, T.N.; Vanderperre, B.; Ouled Amar Bencheikh, B.; Ruskey, J.A.; Laurent, S.B.; Spiegelman, D.; Postuma, R.B.; Arnulf, I.; Hu, M.T.M.; et al. Genetic, Structural, and Functional Evidence Link TMEM175 to Synucleinopathies. Ann. Neurol. 2020, 87, 139–153. [Google Scholar] [CrossRef]
  331. Hu, M.; Li, P.; Wang, C.; Feng, X.; Geng, Q.; Chen, W.; Marthi, M.; Zhang, W.; Gao, C.; Reid, W.; et al. Parkinson’s Disease-Risk Protein TMEM175 Is a Proton-Activated Proton Channel in Lysosomes. Cell 2022, 185, 2292–2308.e20. [Google Scholar] [CrossRef]
  332. Qu, L.; Lin, B.; Zeng, W.; Fan, C.; Wu, H.; Ge, Y.; Li, Q.; Li, C.; Wei, Y.; Xin, J.; et al. Lysosomal K + Channel TMEM175 Promotes Apoptosis and Aggravates Symptoms of Parkinson’s Disease. EMBO Rep. 2022, 23, e53234. [Google Scholar] [CrossRef] [PubMed]
  333. Zhou, S.; Tian, Y.; Song, X.; Xiong, J.; Cheng, G. Brain Proteome-Wide and Transcriptome-Wide Association Studies, Bayesian Colocalization and Mendelian Randomization Analyses Revealed Causal Genes of Parkinson’s Disease. J. Gerontol. A Biol. Sci. Med. Sci. 2022. [Google Scholar] [CrossRef] [PubMed]
  334. Do, C.B.; Tung, J.Y.; Dorfman, E.; Kiefer, A.K.; Drabant, E.M.; Francke, U.; Mountain, J.L.; Goldman, S.M.; Tanner, C.M.; Langston, J.W.; et al. Web-Based Genome-Wide Association Study Identifies Two Novel Loci and a Substantial Genetic Component for Parkinson’s Disease. PLoS Genet. 2011, 7, e1002141. [Google Scholar] [CrossRef] [Green Version]
  335. Alcalay, R.N.; Levy, O.A.; Wolf, P.; Oliva, P.; Zhang, X.K.; Waters, C.H.; Fahn, S.; Kang, U.J.; Liong, C.; Ford, B.; et al. SCARB2 Variants and Glucocerebrosidase Activity in Parkinson’s Disease. NPJ Park. Dis. 2016, 2, 16004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Chen, S.; Zhang, Y.; Chen, W.; Wang, Y.; Liu, J.; Rong, T.-Y.; Ma, J.-F.; Wang, G.; Zhang, J.; Pan, J.; et al. Association Study of SCARB2 Rs6812193 Polymorphism with Parkinson’s Disease in Han Chinese. Neurosci. Lett. 2012, 516, 21–23. [Google Scholar] [CrossRef] [PubMed]
  337. Michelakakis, H.; Xiromerisiou, G.; Dardiotis, E.; Bozi, M.; Vassilatis, D.; Kountra, P.-M.; Patramani, G.; Moraitou, M.; Papadimitriou, D.; Stamboulis, E.; et al. Evidence of an Association between the Scavenger Receptor Class B Member 2 Gene and Parkinson’s Disease. Mov. Disord. 2012, 27, 400–405. [Google Scholar] [CrossRef]
  338. Hopfner, F.; Schulte, E.C.; Mollenhauer, B.; Bereznai, B.; Knauf, F.; Lichtner, P.; Zimprich, A.; Haubenberger, D.; Pirker, W.; Brücke, T.; et al. The Role of SCARB2 as Susceptibility Factor in Parkinson’s Disease. Mov. Disord. 2013, 28, 538–540. [Google Scholar] [CrossRef]
  339. Murphy, K.E.; Halliday, G.M. Glucocerebrosidase Deficits in Sporadic Parkinson Disease. Autophagy 2014, 10, 1350–1351. [Google Scholar] [CrossRef] [Green Version]
  340. Alcalay, R.N.; Levy, O.A.; Waters, C.C.; Fahn, S.; Ford, B.; Kuo, S.H.; Mazzoni, P.; Pauciulo, M.W.; Nichols, W.C.; Gan-Or, Z.; et al. Glucocerebrosidase Activity in Parkinson’s Disease with and without GBA Mutations. Brain 2015, 138, 2648–2658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Suzuki, K.; Suzuki, Y. Globoid Cell Leucodystrophy (Krabbe’s Disease): Deficiency of Galactocerebroside Beta-Galactosidase. Proc. Natl. Acad. Sci. USA 1970, 66, 302–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Suzuki, K. Globoid Cell Leukodystrophy (Krabbe’s Disease): Update. J. Child Neurol. 2003, 18, 595–603. [Google Scholar] [CrossRef]
  343. Senkevich, K.; Zorca, C.E.; Dworkind, A.; Rudakou, U.; Somerville, E.; Yu, E.; Ermolaev, A.; Nikanorova, D.; Ahmad, J.; Ruskey, J.A.; et al. GALC Variants Affect Galactosylceramidase Enzymatic Activity and Risk of Parkinson’s Disease. Brain 2022. [Google Scholar] [CrossRef]
  344. Tam, V.; Patel, N.; Turcotte, M.; Bossé, Y.; Paré, G.; Meyre, D. Benefits and Limitations of Genome-Wide Association Studies. Nat. Rev. Genet. 2019, 20, 467–484. [Google Scholar] [CrossRef]
  345. Dang, X.; Zhang, Z.; Luo, X.-J. Mendelian Randomization Study Using Dopaminergic Neuron-Specific EQTL Nominates Potential Causal Genes for Parkinson’s Disease. Mov. Disord. 2022, 37, 2451–2456. [Google Scholar] [CrossRef]
  346. Kia, D.A.; Zhang, D.; Guelfi, S.; Manzoni, C.; Hubbard, L.; Reynolds, R.H.; Botía, J.; Ryten, M.; Ferrari, R.; Lewis, P.A.; et al. Identification of Candidate Parkinson Disease Genes by Integrating Genome-Wide Association Study, Expression, and Epigenetic Data Sets. JAMA Neurol. 2021, 78, 464–472. [Google Scholar] [CrossRef] [PubMed]
  347. Cui, X.; Xu, C.; Zhang, L.; Wang, Y. Identification of Parkinson’s Disease-Causing Genes via Omics Data. Front. Genet. 2021, 12, 712164. [Google Scholar] [CrossRef]
  348. Yang, C.; Farias, F.H.G.; Ibanez, L.; Suhy, A.; Sadler, B.; Fernandez, M.V.; Wang, F.; Bradley, J.L.; Eiffert, B.; Bahena, J.A.; et al. Genomic Atlas of the Proteome from Brain, CSF and Plasma Prioritizes Proteins Implicated in Neurological Disorders. Nat. Neurosci. 2021, 24, 1302–1312. [Google Scholar] [CrossRef] [PubMed]
  349. Kaiser, S.; Zhang, L.; Mollenhauer, B.; Jacob, J.; Longerich, S.; Del-Aguila, J.; Marcus, J.; Raghavan, N.; Stone, D.; Fagboyegun, O.; et al. A Proteogenomic View of Parkinson’s Disease Causality and Heterogeneity. NPJ Park. Dis 2023, 9, 24. [Google Scholar] [CrossRef]
  350. Klein, A.D.; Mazzulli, J.R. Is Parkinson’s Disease a Lysosomal Disorder? Brain 2018, 141, 2255–2262. [Google Scholar] [CrossRef] [Green Version]
  351. Teixeira, M.; Sheta, R.; Idi, W.; Oueslati, A. Alpha-Synuclein and the Endolysosomal System in Parkinson’s Disease: Guilty by Association. Biomolecules 2021, 11, 1333. [Google Scholar] [CrossRef]
  352. Shahmoradian, S.H.; Lewis, A.J.; Genoud, C.; Hench, J.; Moors, T.E.; Navarro, P.P.; Castaño-Díez, D.; Schweighauser, G.; Graff-Meyer, A.; Goldie, K.N.; et al. Lewy Pathology in Parkinson’s Disease Consists of Crowded Organelles and Lipid Membranes. Nat. Neurosci. 2019, 22, 1099–1109. [Google Scholar] [CrossRef] [Green Version]
  353. Shippey, L.E.; Campbell, S.G.; Hill, A.F.; Smith, D.P. Propagation of Parkinson’s Disease by Extracellular Vesicle Production and Secretion. Biochem. Soc. Trans. 2022, 50, 1303–1314. [Google Scholar] [CrossRef]
  354. Xie, Y.X.; Naseri, N.N.; Fels, J.; Kharel, P.; Na, Y.; Lane, D.; Burré, J.; Sharma, M. Lysosomal Exocytosis Releases Pathogenic α-Synuclein Species from Neurons in Synucleinopathy Models. Nat. Commun. 2022, 13, 4918. [Google Scholar] [CrossRef]
  355. Espay, A.J.; Okun, M.S. Abandoning the Proteinopathy Paradigm in Parkinson Disease. JAMA Neurol. 2022, 64, 485–491. [Google Scholar] [CrossRef]
  356. Giguère, N.; Burke Nanni, S.; Trudeau, L.-E. On Cell Loss and Selective Vulnerability of Neuronal Populations in Parkinson’s Disease. Front. Neurol. 2018, 9, 455. [Google Scholar] [CrossRef] [PubMed]
  357. Cheng, H.-C.; Ulane, C.M.; Burke, R.E. Clinical Progression in Parkinson Disease and the Neurobiology of Axons. Ann. Neurol. 2010, 67, 715–725. [Google Scholar] [CrossRef] [PubMed]
  358. Chu, Y.; Morfini, G.A.; Langhamer, L.B.; He, Y.; Brady, S.T.; Kordower, J.H. Alterations in Axonal Transport Motor Proteins in Sporadic and Experimental Parkinson’s Disease. Brain 2012, 135, 2058–2073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  359. Burke, R.E.; O’Malley, K. Axon Degeneration in Parkinson’s Disease. Exp. Neurol. 2013, 246, 72–83. [Google Scholar] [CrossRef] [Green Version]
  360. Abe, T.; Kuwahara, T. Targeting of Lysosomal Pathway Genes for Parkinson’s Disease Modification: Insights From Cellular and Animal Models. Front. Neurol. 2021, 12, 681369. [Google Scholar] [CrossRef]
  361. Abeliovich, A.; Hefti, F.; Sevigny, J. Gene Therapy for Parkinson’s Disease Associated with GBA1 Mutations. J. Park. Dis. 2021, 11, S183–S188. [Google Scholar] [CrossRef]
  362. Pagan, F.L.; Wilmarth, B.; Torres-Yaghi, Y.; Hebron, M.L.; Mulki, S.; Ferrante, D.; Matar, S.; Ahn, J.; Moussa, C. Long-Term Safety and Clinical Effects of Nilotinib in Parkinson’s Disease. Mov. Disord. 2021, 36, 740–749. [Google Scholar] [CrossRef] [PubMed]
  363. Simuni, T.; Fiske, B.; Merchant, K.; Coffey, C.S.; Klingner, E.; Caspell-Garcia, C.; Lafontant, D.-E.; Matthews, H.; Wyse, R.K.; Brundin, P.; et al. Efficacy of Nilotinib in Patients With Moderately Advanced Parkinson Disease: A Randomized Clinical Trial. JAMA Neurol. 2021, 78, 312–320. [Google Scholar] [CrossRef]
  364. Jennings, D.; Huntwork-Rodriguez, S.; Henry, A.G.; Sasaki, J.C.; Meisner, R.; Diaz, D.; Solanoy, H.; Wang, X.; Negrou, E.; Bondar, V.V.; et al. Preclinical and Clinical Evaluation of the LRRK2 Inhibitor DNL201 for Parkinson’s Disease. Sci. Transl. Med. 2022, 14, eabj2658. [Google Scholar] [CrossRef]
  365. Cavallieri, F.; Cury, R.G.; Guimarães, T.; Fioravanti, V.; Grisanti, S.; Rossi, J.; Monfrini, E.; Zedde, M.; Di Fonzo, A.; Valzania, F.; et al. Recent Advances in the Treatment of Genetic Forms of Parkinson’s Disease: Hype or Hope? Cells 2023, 12, 764. [Google Scholar] [CrossRef]
Figure 1. Graphic representation of a dopaminergic neuron illustrating the relevant intracellular organelles, the involved monogenic PD genes (dark blue), and PD risk genes (light blue). In the represented model, a genetically determined impairment of the endolysosomal function in association with additional genetic and environmental pathogenic hits causes a defect in the degradation of misfolded proteins (including α-synuclein), which induces a compensatory autophagic response to address the disposal of the pathological proteins. However, a defective endolysosomal pathway implies an ineffective autophagic response, thus triggering a self-feeding vicious cycle of abnormal protein accumulation by dysfunctional engulfed lysosomes. As a consequence, the accumulation of misfolded aggregated proteins and engulfed lysosomes–autophagosomes leads to the progressive formation of pathological inclusions (Lewy bodies) and ultimately to neuronal distress and death. Created with BioRender.com.
Figure 1. Graphic representation of a dopaminergic neuron illustrating the relevant intracellular organelles, the involved monogenic PD genes (dark blue), and PD risk genes (light blue). In the represented model, a genetically determined impairment of the endolysosomal function in association with additional genetic and environmental pathogenic hits causes a defect in the degradation of misfolded proteins (including α-synuclein), which induces a compensatory autophagic response to address the disposal of the pathological proteins. However, a defective endolysosomal pathway implies an ineffective autophagic response, thus triggering a self-feeding vicious cycle of abnormal protein accumulation by dysfunctional engulfed lysosomes. As a consequence, the accumulation of misfolded aggregated proteins and engulfed lysosomes–autophagosomes leads to the progressive formation of pathological inclusions (Lewy bodies) and ultimately to neuronal distress and death. Created with BioRender.com.
Ijms 24 06338 g001
Figure 2. Phenotypic comparison between each monogenic form and idiopathic PD (represented by the dashed hexagon). The considered features are: age of onset (AOO), motor signs (bradykinesia, rigidity, rest tremor), cognitive problems (including neuropsychiatric symptoms and intellectual disability in the case of RAB39B), non-motor symptoms of PD including (dysautonomia, hyposmia, REM sleep behavior disorder), atypical signs (e.g., epilepsy, spasticity, dystonia, dysmorphisms), and therapy failure (poor response to levodopa treatment or early onset of LID and MF).
Figure 2. Phenotypic comparison between each monogenic form and idiopathic PD (represented by the dashed hexagon). The considered features are: age of onset (AOO), motor signs (bradykinesia, rigidity, rest tremor), cognitive problems (including neuropsychiatric symptoms and intellectual disability in the case of RAB39B), non-motor symptoms of PD including (dysautonomia, hyposmia, REM sleep behavior disorder), atypical signs (e.g., epilepsy, spasticity, dystonia, dysmorphisms), and therapy failure (poor response to levodopa treatment or early onset of LID and MF).
Ijms 24 06338 g002
Table 1. Monogenic causes of PD (including GBA1-PD).
Table 1. Monogenic causes of PD (including GBA1-PD).
Gene (Inheritance)BiologyPhenotypeLevodopa ResponseNeuropathology
LRRK2 (AD)Missense GOF disrupts vesicular transport and mitophagyClassical PD, cancerGood, with frequent LIDLBD of SN, occasional Tau pathology
SNCA (AD)Multiplications increase α-synuclein expression, missense impair degradationEarlier onset, dystonia, pyramidal signs, cerebellar signs, cognitive impairment, psychosis, depression, anxiety, orthostatic hypotension, urinary dysautonomiaGood, with frequent MF and LIDSN and cortical LBD, α-synuclein GCIs
VPS35 (AD)D620N impairs autophagosome formation (ATG9A), lysosomal function (cathepsin D), and mitophagy; increases Lrrk2 substrates phosphorylationClassical PD with earlier onset; cognitive and dysautonomic features are rareGood, with frequent MF and LIDNA
VPS13C (AR)LOF alters lysosomal homeostasis and promotes Pink1/Parkin mitophagyParkinsonism with dystonia, rare pyramidal signs, early cognitive decline, visual hallucinationsGood, with frequent MF and LIDDiffuse LBD
ATP13A2 (AR)LOF causes ER stress, lysosomal dysfunction (defective polyamine export), and autophagosome dysfunctionFrom EOPD (usually missense variants) to KRD (usually truncating mutations), possible striatal iron accumulationInitial response, MF and LID rapidly occurWidespread neuronal and glial lipofuscin accumulation without α-synuclein pathology
RAB39B (XLR)Altered α-synuclein homeostasis, abnormal synapse–endosome–Golgi trafficking and recyclingFrom classical PD to complex syndromes with parkinsonism, dystonia, ID, globi pallidi iron depositionGood, with frequent MF and LIDSN and cortical LBD, SN tau-positive neurofibrillary tangles
SYNJ1 (AR)LOF causes synaptic autophagy and transmission defectsFrom EOPD to atypical parkinsonism with oculomotion problems, dysarthria/anarthria, dysphagia, epilepsy, dystonia, pyramidal signs, and cognitive declineVariable, with different adverse effects including (LID, dystonia, postural hypotension)SN neuronal loss without LBD, neurofibrillary degeneration, tau protein staining in cell bodies and axonal hillocks
DNAJC6 (AR)LOF disrupts synaptic vesicle endocytosis and exacerbates α-synuclein overexpressionEOP, rapid progression with walking loss in 10 years, dystonia, pyramidal signs, postural tremor, dysarthria/anarthria, epilepsy, ID, and psychosisVariable, frequent LIDNA
GBA1 (AD risk factor)Proposed disease mechanisms include ER stress, ER-Golgi transport impairment, lysosomal dysfunction with impaired α-synuclein clearance (vicious cycle), autophagy dysfunction, lipid homeostasis disruption, mitochondrial dysfunction, neuroinflammation.Classical PD with earlier onset, faster progression, shorter survival, severe dysautonomia, depression, anxiety, and cognitive declineGood response, MF and LID appear earlier and more frequentlySN and cortical LBD
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yahya, V.; Di Fonzo, A.; Monfrini, E. Genetic Evidence for Endolysosomal Dysfunction in Parkinson’s Disease: A Critical Overview. Int. J. Mol. Sci. 2023, 24, 6338. https://doi.org/10.3390/ijms24076338

AMA Style

Yahya V, Di Fonzo A, Monfrini E. Genetic Evidence for Endolysosomal Dysfunction in Parkinson’s Disease: A Critical Overview. International Journal of Molecular Sciences. 2023; 24(7):6338. https://doi.org/10.3390/ijms24076338

Chicago/Turabian Style

Yahya, Vidal, Alessio Di Fonzo, and Edoardo Monfrini. 2023. "Genetic Evidence for Endolysosomal Dysfunction in Parkinson’s Disease: A Critical Overview" International Journal of Molecular Sciences 24, no. 7: 6338. https://doi.org/10.3390/ijms24076338

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