Next Article in Journal
Role of Astrocytes in Parkinson’s Disease Associated with Genetic Mutations and Neurotoxicants
Next Article in Special Issue
Syringin Prevents 6-Hydroxydopamine Neurotoxicity by Mediating the MiR-34a/SIRT1/Beclin-1 Pathway and Activating Autophagy in SH-SY5Y Cells and the Caenorhabditis elegans Model
Previous Article in Journal
Estrogen Actions in Placental Vascular Morphogenesis and Spiral Artery Remodeling: A Comparative View between Humans and Mice
Previous Article in Special Issue
Systematic Functional Analysis of PINK1 and PRKN Coding Variants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathogenic Aspects and Therapeutic Avenues of Autophagy in Parkinson’s Disease

Univ. de Bordeaux, CNRS, IMN, UMR 5293, F-33000 Bordeaux, France
*
Author to whom correspondence should be addressed.
Cells 2023, 12(4), 621; https://doi.org/10.3390/cells12040621
Submission received: 10 January 2023 / Revised: 7 February 2023 / Accepted: 11 February 2023 / Published: 15 February 2023
(This article belongs to the Special Issue Autophagy in Parkinson's Disease)

Abstract

:
The progressive aging of the population and the fact that Parkinson’s disease currently does not have any curative treatment turn out to be essential issues in the following years, where research has to play a critical role in developing therapy. Understanding this neurodegenerative disorder keeps advancing, proving the discovery of new pathogenesis-related genes through genome-wide association analysis. Furthermore, the understanding of its close link with the disruption of autophagy mechanisms in the last few years permits the elaboration of new animal models mimicking, through multiple pathways, different aspects of autophagic dysregulation, with the presence of pathological hallmarks, in brain regions affected by Parkinson’s disease. The synergic advances in these fields permit the elaboration of multiple therapeutic strategies for restoring autophagy activity. This review discusses the features of Parkinson’s disease, the autophagy mechanisms and their involvement in pathogenesis, and the current methods to correct this cellular pathway, from the development of animal models to the potentially curative treatments in the preclinical and clinical phase studies, which are the hope for patients who do not currently have any curative treatment.

1. Introduction

The threat of Parkinson’s disease (PD) has been studied since its first description by James Parkinson at the beginning of the 19th century [1]. Still, the global understanding of this neurodegenerative pathology has increased exponentially since the end of the 20th century. Indeed, a better understanding of symptoms from dopaminergic (DA) deficiency [2] allows for the development of the first symptomatic treatment with Levodopa, the precursor to dopamine, which is still one of the most effective drugs for Parkinson’s disease (PD) but is associated with significant side effects such as dyskinesia and habituation [3]. Furthermore, the discovery of the α-synuclein (α-syn) role [4] and a genetic role in pathogenesis by Polymeropoulos and colleagues [5] allows for the emergence of new fields of study, ranging from animal modeling to α-syn-based therapeutics.
Nowadays, the number of genes implicated is over 25 (for instance, the TMEM175 gene emerged in 2022 as a gene involved in PD); further understanding of the pathology and its link with the cellular clearance mechanism of autophagy are highlighted. Other drugs appeared, like monoamine oxidase type B [6] or catechol-O-methyl transferase [7] inhibitors. For example, the development of deep-brain stimulation [8] permits better nursing of PD patients, but these symptomatic treatments are not curative and are still accompanied by possible side effects.
In this review, we discuss the new insight into PD with the appearance of new genetic risk factors and the autophagy process and its implication in neurons and specifically in the pathology, making this mechanism an interesting way to develop curative treatments, some of which are presently in preclinical and clinical phases and represent an important hope for the cure of PD.

2. Parkinson’s Disease

Epidemiologically and globally, the burden of PD is approaching a threefold increase in 26 years, from 2.5 million patients in 1990 to 6.1 million patients in 2016. Over the next 30 years, the number of PD patients should reach more than 12 million worldwide by about 2050 [9,10,11]. Clinically, PD is characterized partly by its non-motor symptoms, such as constipation, sleep disorder, depression, pain, or also dementia [12]. However, it is primarily known for its motor symptoms, including bradykinesia, body rigidity, dysphagia, and tremor [13,14]. The loss of several neuronal populations characterizes PD. Still, the most characterized one is the extensive cell loss of DA neurons in the substantia nigra (SN), an essential input in the regulation of the basal ganglia motor loop through the nigrostriatal pathway [15,16].
Along with this degeneration, the pathological hallmark of PD is the presence of intraneuronal proteinaceous cytoplasmic inclusions, named Lewy bodies (LB), that invade the whole nervous system as the disease progresses [17]. Several fundamental discoveries have strongly implicated the protein α-syn in the pathogenesis of both familial and sporadic forms of PD (for in-depth review [18]). Other proteins [19], lipids [20], and organelles [21,22] also accumulate in LB [19], but α-syn is the main protein component [4].
PD pathogenesis can occur from environmental factors such as pesticide exposure, head trauma, or aging [23] and can also be the result of genetic mutations, with an increasing list of genes and mutations related to familial PD, comprising autosomal dominant forms like SNCA, LRRK2, or VPS35 and autosomal recessive genes like PRKN, PINK1, PARK7, and DJ1 mutations (Table 1).

3. Autophagy Mechanism

Autophagy is one of the two cellular processes in charge of cell clearance and maintaining cellular homeostasis. Disruption of autophagy is linked to multiple cellular malfunctions, starting with an accumulation of non-degraded elements [63], enhancement of reactive oxygen species (ROS) production [64], and neuronal bioenergetic imbalance [65]. Autophagy dysfunction thus appears as an essential component of PD pathogenesis [66,67]. Autophagy can be divided into three main pathways: macroautophagy (MA), chaperone-mediated autophagy (CMA), and microautophagy (Figure 1). Having all their particularities, they share the lysosome as a common organelle, permitting the elements’ degradation and recycling of their essential components. Lysosomes are tiny intracellular organelles with an average diameter of 500 nm [68] and contain over 50 different hydrolases [69], which are activated by the low lumen pH of the lysosome, which is around 4.5 [70]. This acidic pH, characteristic of a functional lysosome, is due to multiple membranous proton pumps, part of the more than 700 referenced lysosomal membrane proteins [71].

3.1. Macroautophagy

Macroautophagy (MA) is an autophagy-lysosomal pathway (ALP) mechanism that can be specific or not and is implicated in many cellular processes, like central nervous system development [72,73] and maintenance [74]. This mechanism for delivering cellular components to lysosomes for degradation is composed of multiple steps.
First, the initiation phase permits phagophore production and is negatively regulated by the mechanistic target of rapamycin (mTOR) complex 1 (mTORC1). Formation of this lipidic double membrane, assembled mainly from the endoplasmic reticulum and Golgi membranes [75,76,77,78], is triggered by the ULK1 (unc-51-like kinase 1) complex formed with ATG (autophagy-related protein) 13, ATG101, and FIP200 (focal adhesion kinase family interacting protein of 200 kDa) [79,80,81,82]. The ULK1 complex permits activation of a second kinase complex composed of Beclin-1, ATG14, and vacuolar protein sorting (VPS) 15 and 34, generating a cascade of reactions resulting in the formation and expansion of the phagophore at the autophagy initiation site [83]. Selective MA uses autophagy receptors to initiate the formation of the phagophore, like phosphorylation, ubiquitination, acetylation, and oligomerization [84], which are recognized by phagophore proteins ATG8, LC3, and GABARAP [85]. Once closed, the now-called autophagosome can enter a maturation phase, fusing with endosomes to add new material to be degraded [86]. Finally, the autophagosome fuses with lysosomes via GTPases, lipids, and SNARE regulation [87], allowing content degradation.

3.2. Chaperone-Mediated Autophagy

CMA is a specific ALP process with a role in cell metabolism [88] and cycle [89,90]. CMA uses a recognition complex composed of the cytosolic Hsc70 protein (heat shock cognate 70 kDa) and co-chaperones that bind to a pentapeptide CMA-specific recognition motif (the KFERQ motif). This motif is composed of a succession of charged residues, hydrophobic residues, and glutamine. It can be formed and exposed through the 3D folding of the protein [91] or by acetylation [92] and phosphorylation [93] post-transcriptional modifications. Once the Hcp70 complex recognizes the KFERQ-like motif, it translocates and unfolds the substrate needed to be degraded. Finally, the chaperone-substrate complex binds to the monomeric lysosomal-associated membrane protein-2A (LAMP-2A), which then forms a multimeric complex to permit the lysosomal entry of the substrate to allow proteasomal degradation.

3.3. Endosomal Microautophagy

The lysosome can also directly phagocytose and degrade different elements like mitochondria [94,95], peroxisomes [96], part of the endoplasmic reticulum [97], and the nucleus [98]. This process, known as “microautophagy”, can be selective or non-selective, and it uses invagination or protrusion of the lysosomal membrane to capture elements that need to be degraded. LC3, ESCRT (Endosomal Sorting Complex Required for Transport), and the ATPase VPS4A are involved in the invagination microautophagy mechanism [97]. Still, they are also implicated in late endosome (LE) membrane modulation in endosomal microautophagy (eMI). This process, similar to microautophagy, uses a late endosome as a degradative compartment. The substrate uptake can be direct or selective with Hsc70 complex binding to LE membrane phosphatidyl serine before phagocyted in the lumen in ESCRT-complex dependent manner [99]. eMI activation is enhanced in oxidative or genotoxic stress conditions and results in the diminution of MA activity [100].

4. Autophagy Disruption in Parkinson’s Disease

Many neurodegenerative diseases, including PD, are linked to autophagy defects by different pathways. Herein, we focus on some of the most prevalent PD-related mutations causing autophagy phenotypes.

4.1. Genetic Implication in Autophagy Dysfunction

Autophagy impairment can result from genetic mutations. Constitutive MA is essential for neuronal survival, as its selective genetic inactivation in neurons leads to the formation of ubiquitinated intracellular inclusions and neuron cell loss in mutant mice [101,102]. ALP dysfunctions are caused by multiple genes in Parkinson’s disease (Table 1). Human genetics studies indicated that lysosomal dysfunction may play an essential role in the pathogenesis of PD; a recent genome-wide association study (GWAS) discovered a significant burden of rare, likely damaging LSD gene variants in association with PD risk, such that 56% of PD cases have at least one putatively damaging variant in an LSD gene [103]. Additional studies confirmed previous results, identifying 18 and 6 new loci associated with PD [104,105,106]. At least 11 of these loci are either directly or indirectly (i.e., disruptible) linked to the ALP, implying that lysosomal dysfunction may play a primary pathogenic role in Parkinson’s disease.
The MA mechanism dysfunction, caused by different gene mutations, can be linked with PD, with SNCA gene mutations being the first reported cause of PD cases, with primary evidence in 1997 and the last discovery in 2021. The eight-point mutations in SNCA known are A30G [107], A30P [108], E46K [109], H50Q [110,111], G51D [112], A53E [113], A53T [5], and A53V [114]. These multiple point mutations and the multiplication of this gene [115] are linked to different phenotypes of PD patients according to the mutations, from typical pathology to early-onset dementia, hallucinations, myoclonus, or pyramidal signs. But SNCA-PD patients all share usual hallmarks such as LB presence, neuronal depletion, and autophagic flux problems in MA and CMA mechanisms [116].
The mitophagy subtype of the MA mechanism can be affected explicitly by PD-related mutations. PTEN-induced putative kinase 1 (PINK1) and Parkin, respectively encoded by PINK1 and PRKN genes, are two proteins interacting in the outer mitochondrial membrane and ubiquitinating it when the organelle is damaged, permitting its selective elimination [117]. PINK1 or PRKN mutations result in the difficulty of mitochondrial elimination by disrupting the mitophagy initiation step, leading to their cellular accumulation [118] and causing autosomal recessive juvenile Parkinsonism [119,120]. Mutations of the PARK7 gene encoding DJ1 are related to early-onset PD due to its interactions with PINK1 and the increased action of Parkin in wild-type conditions. The DJ1 inactivation induces a reduced steady-state level of PINK1, resulting in mitochondrial dysfunction [121].
LLRK2 mutations are the most common genetic cause of late-onset PD [25,122], contributing to specific MA and CMA dysfunction. Over 50 mutations of this gene lead to dysregulation of cytosolic LRRK2 (leucine-rich repeat protein kinase-2) activity, which affects the endocytic process of a synaptic vesicle or lysosomal maintenance [123]. Significant outcomes of LRRK2 mutations are an increase in lysosomal impairment and lysophagy due to LRRK2 enhanced activity [124], α-syn aggregation [125], autophagic vacuoles accumulation [26], and alteration of endolysosomal trafficking [126]. Heat shock proteins and co-chaperone mutations, playing an essential role in the CMA process, also dysregulate the machinery, like for DNAJC6 encoding Hsp40. This co-chaperone regulating Hsc70 complex recruitment leads to ALP impairment in DNAJC6 mutations [127].
The final and most common element of the ALP process, the lysosome, is also the target of different mutations, possibly modifying lysosomal pH or enzymatic activity and making it inactive. Glucocerebrosidase (GCase) is a lysosomal sphingolipid degrading enzyme, permitting the degradation of glucocerebroside in the lysosome. Mutations of GBA causing GCase lack leads to accumulation of substrate in the lysosomal lumen and α-syn aggregation [128]. ATP13A2 is another lysosomal protein that can be responsible for lysosomal malfunction. This metal cation transporter regulates lysosomal acidity and homeostasis by carrying Fe3+, Mn2+, Zn2+, and Ca2+. ATP13A2 and its gene PARK9 impairments are linked to a decrease in clearance in the lysosome due to significant alterations such as acidification malfunction and reduction of the proteolytic enzymatic mechanism [38]. Recently, the TMEM175 proton leak channel of lysosomes and endosomes regulating lumen pH through a negative feedback mechanism has been identified as a PD risk [129]. Indeed, this channel regulating lysosomal functions through proton-activating [62], or growth-factor activating manner [130], once mutated in its pore or luminal loops regions, decreases its H+ permeability, diminishing lysosomal pH below the physiological 4.5 value and predisposing neurons to stress-induced damage and accelerates the accumulation of pathological α-syn because of the inactivation of enzymes [130,131], resulting in apoptosis promotion and aggravated PD symptoms [132]. In addition to lysosomal impairments caused by TMEM175 deficiency, mitochondrial functions also appear degraded [61]. Firstly, as observed in other synucleinopathies [133], variants of TMEM175 were reported in 2023 in Italian PD patients [134].

4.2. Rodent Models Linked to ALP Dysfunctions

Different animal models are used in literature to understand better the ALP mechanisms and test strategies for restoring cellular clearance; herein, we discuss some of the novels or optimized rodent models based on ALP impairment involved in PD, proposed since 2018. ALP endosomal formation through the clathrin protein process is impaired by SYNJ1 mutations encoding synaptojanin-1, an auxilin-like protein. Heterozygous deletion of SYNJ1 in mice displays an age-dependent motor dysfunction phenotype in the rotarod test, with the histological observation of α-syn accumulation in SN and other brain regions, autophagy deficit, and DA neurodegeneration at 18 months old [135]. In 2021, the refinement of a mouse model with conditional knockout of autophagy-related gene 5 (Atg5) [136] was generated from a Lyz2Cre (lysosome 2 Cre) strain expressing Cre recombinase in myeloid cells, and Atg5f/f mice permitted Atg5 silencing only in Cre-positive cells. The silencing of this protein involved in the autophagosome formation and maturation processes permits this mouse strain to observe down-regulated autophagy activity, neuroinflammation, and neurodegeneration with approximately 50% fewer DA neurons in the SN and exacerbated locomotor deficits [137]. The retromer complex also maintains endosomal functions, regulating the recycling of cargo proteins, and its subunit VPS35 is associated with AD forms of PD. The conditional VPS35 D620N knock-in (KI) model obtained from VPS35 D620N KI mice characterized by Cataldi et al. [138] crossed with mice from the Sox2-Cre-delete line results in motor deficiency after 14 months in the open field and narrow beam tests, along with a significant decrease of DA neurons in the SN and fibers in the striatum compared to control. This neurodegeneration is accompanied by a LAMP2A level decrease, lysosomal accumulation, and twice as much lipofuscin in the SN, suggesting autophagic dysfunction in a PD-like manner [139]. Reproducing PD mutations clinically observed in an animal model was also used in LRRK2 G2019S mice strain expressing the mutated protein, only in DA neurons through regulation of LRRK2-mutated gene by TH promotor, refining a 2011 model [140]. This model shows neurodegeneration of DA neurons and a significantly higher presence of an insoluble fraction of p129S α-syn at 15 and 24 months old in the striatum and the ventral midbrain Indeed, overexpression of the LRRK2 G2019S mouse model combined with injection of the MPTP neurotoxin resulted in severe motor impairment, selective loss of DA neurons, and increased astrocyte activation, indicating that the combination had a synergistic effect [141].
Causing lysosomal defects to mimic ALP-PD-related dysfunction is a strategy used in animal models, like with the recently described PD-related gene TMEM175 KO mouse models. This model, obtained through CRISPR-Cas9 methods, leads to a decreased lysosomal pH at 4, reducing cathepsin B and D activity and a significative increase of phosphorylated α-syn presence compared to control mice. The accelerated fusion between autophagosome and lysosome was also reported, along with an accumulation of undigested autophagosomes in TMEM175 KO models [130,131].
Another model of overexpression targeting the lysosomal dysfunction of human tyrosinase via intracerebral injection of a viral vector in the SN of both rats and mice permitted an age-dependent production and accumulation of neuromelanin in DA neurons, progressively occupying all the cytoplasmic space of the cell. Due to the insoluble property of neuromelanin, which makes it non-degradable, this model shows motor impairment in the contralateral side from 2 months after injection and leads to the observation of 6 times fewer DA neurons in the ipsilateral SN 24 months after injection, intracellular inclusion body formation, and lysosomal and autophagy dysfunction in rats [142].
The increasing number of animal models linked to ALP dysfunction that mimic the PD phenotype is a key step in understanding this pathology and developing and testing possible future treatments. In addition to these latest findings, it is important to note the usefulness of non-animal models, particularly brain organoids, in the first steps of study due to their capacity to generate complex and multicellular systems from patients’ pluripotent stem cells [143].

5. Autophagy-Related Therapies for PD

We have reported here that ALP impairment plays a pivotal role in different aspects of PD pathophysiology. According to these new findings, therapeutic modulation of autophagy in the context of PD and lysosomal-related pathologies paves the way for the development of a panel of therapeutic strategies (Figure 2).

5.1. Pharmacological Treatments

5.1.1. mTOR-Dependent Drugs

The mechanistic target of rapamycin (mTOR) is a type of protein kinase involved in two complexes, the mTOR complex 2 regulating cell growth, proliferation, and protein synthesis, and the mTOR complex 1 (mTORC1) downregulating MA initiation, making it a therapeutic target objective. Early evidence suggests that rapamycin and derivative treatment reduces neurodegeneration and improves motor capacities and non-motor behavior in neurotoxin-MPTP or overexpressing α-syn mice models [144,145,146]. Other drugs are in the preclinical study phase to modulate the activity of mTOR and then increase autophagic activity. Caffeine also acts as an mTOR inhibitor, increases the number of autophagosomes, and reduces apoptosis, reestablishing autophagy activity [147,148]. Similarly, curcumin inhibits mTOR activity and permits autophagy enhancement [149]. AZD8055 shows interesting results as an mTOR inhibitor by increasing autophagy flux, the number of lysosomes, and acidified autolysosomes [150].
Similarly, corynoxine small-molecule was shown in 2021 to have neuroprotective effects, enhancing motor performances and decreasing α-syn aggregates in neurotoxic rotenone-induced mice PD-model [151]. Recently, small-molecule pyrazole derivatives invented by Kim and colleagues show remarkable mTORC1 inhibitory activity over 85% below 220 nM treatment without affecting mTORC2. Evaluated in mice, these molecules easily cross the blood-brain barrier (BBB) and decrease amyloid plaques, the Alzheimer’s disease hallmark [152]. Finally, new evidence for piperine as an autophagy enhancer emerged [153]. Its action to increase protein phosphatase 2A activity results in mTOR inhibition. It restores autophagy activity in a rotenone-induced PD mouse model [154] and attenuates olfactory and delayed motor deficits in transgenic mice overexpressing human SNCA [153]. The challenges of these treatments are to enhance autophagic flux targeting mTORC1 while avoiding disrupting cell survival, growth, or proliferation, in which mTORC2 plays a role [155], and find a good therapeutic dosage to induce average autophagic flux.

5.1.2. mTOR-Independent Drugs

Some components use different targets than mTOR to restore ALP processes. Multiple autophagic regulators are raised as an attractive potential target through modulation of these protein levels, as further described in Scrivo et al. [156]. Beclin-1 involves the phagophore nucleation/elongation process and has multiple known enhancer drugs. For example, KYP-2047, a prolyl oligopeptidase inhibitor, induces expression of Beclin-1 and is associated with increased α-syn fibril degradation, amelioration of autophagy, resulting in better cell viability in vitro [157,158], but also in SNCA-A30P transgenic mouse model in which is the increase of MA and striatal dopamine levels with a 28-days treatment was reported [159]. Glycyrrhizin inhibits the activity of a Beclin-1 inhibitor and acts as an autophagy enhancer by multiple pathways: it decreases α-syn and glucocorticoid levels and increases LC3 and Beclin-1 activity [160]. Santoro et al. show a dose-dependent effect of glycyrrhizin on DA neurodegeneration in the MPTP mouse model [161]. Trehalose may be the most exciting autophagy enhancer, targeting the TFEB pathway by dephosphorylating TFEB and increasing its nuclear translocation [162]. A 2022 study in a rotenone-induced PD mouse model shows the efficient action of trehalose treatment during the prodromal phase, resulting in improved non-motor parameters, motor performances, and the number of DA neurons of SN, but also decreased deposit of brain α-syn compared to non-treated animals [163]. Similar results were corroborated in a non-human primate model overexpressing the A53T α-syn [164]. Trehalose efficiency in autophagy and pathological condition enhancement is broadly more substantial when combined with sodium butyrate, as shown by Kakoty and colleagues in a PFF-rat model orally treated with 2 g/kg of trehalose and 150 mg/kg of sodium butyrate [165]. Cunha and colleagues report nanoparticles and nucleolipidic constructions of approximately 150 nm diameter, carrying trehalose, as safe and enhancing autophagy in vitro [166].

5.2. Gene Therapies

Targeting key elements of ALP by gene therapy is another promising avenue that is also being explored. First, evidence of LAMP2A overexpression in human mammary adenocarcinoma cell lines permitted enhanced CMA activity and increased cell survival [167]. One year later, the same results were observed in rats overexpressing α-syn in SN, with an injection of recombinant adenoviruses overexpressing LAMP2A, where amelioration of CMA mechanism, lower α-syn levels, and decrease of DA neurodegeneration [168]. A recent result reported the preventive effect of LAMP2A overexpression in the SNCA-induced fly model [169]. Viral-mediated overexpression of Beclin-1 has also proved to have therapeutic interest in rodent models overexpressing human α-syn. Beclin-1 overexpression is also linked to reduced apoptosis and enhanced autophagy [170]. Following Beclin-1 lentiviral injection in the temporal cortex and hippocampus of α-syn transgenic mice, a decrease in α-syn levels was observed through autophagy activation [171]. Finally, transcription factor EB (TFEB) has emerged as a master activator of the ALP machinery, promoting lysosomal biogenesis and enhancing autophagy [172]. Gene therapy using adeno-associated viral (AAV)-mediated TFEB overexpression in wild-type α-syn and human mutated A53T α-syn overexpressing rat models conferred a protective effect on rat midbrain neurons, associated with increased clearance of pathologic α-syn through autophagic recovery [173,174]. In the same study, Beclin-1 gene therapy overexpression shows similar results [174]. While TFEB neuronal expression was sufficient to prevent neurodegeneration in PD models, TFEB oligodendroglial overexpression leads to neuroprotective effects in the transgenic PLP α-syn mouse, as an MSA model [173]. Gene therapies appear as a unique and interesting treatment technology that may be patient-specific by regulating gene expression of specific genotype abnormalities in PD patients, and having a target-specific property depending on the serology type of the adeno-associated virus used is an important point to consider to avoid off-target gene regulation.

5.3. Biotechnological Strategy

Recently, new technologies to restore ALP have emerged, particularly by targeting lysosomal pH. Lysosomal acidification, which is defective in neurodegenerative diseases such as Alzheimer’s disease [175,176] and PD [177], is also essential for lysosomal function. Nanocarriers like nanoparticles or nucleolipids encapsulating drugs of interest are one of them [178]. Biocompatible organic acid-loaded nanoparticles (i.e., poly(DL-lactide-co-glycolide) (PLGA) or succinic diacid) targeting restoration of lysosome acidification have been tested and showed promising results in different pathological contexts [177,179,180,181,182,183,184,185,186]. Thus, in vitro studies in several models of lysosomal impairment, including ATP13A2-mutant cells, GBA-mutant fibroblasts from PD patients, and neuroblastoma cells treated with lysosomotropic agents pathologically enhancing lysosomal pH, demonstrate that PLGA acidic nanoparticles can enhance cell survival and restore the physiological pH of the lysosome lumen [177].
Furthermore, those acid-loaded nanoparticles are neuroprotective in two different mouse models of PD: the MPTP mouse model and the LB-seeding mouse model of PD. The same PLGA acidic nanoparticles were assessed in an atherosclerosis mouse model apoe-/- reporting a significant increase in lysosome and macrophage numbers and the rescue of lysosomal acidity in macrophages [187]. Strategies enhancing or restoring lysosomal-mediated degradation thus appear as tantalizing neuroprotective/disease-modifying therapeutic strategies. Restoration of functional lysosomes would be of significant therapeutic interest for PD, in which nanotechnologies could play a key role because of their interesting properties to easily cross the BBB and because they may be target-specific to avoid adverse effects and deliver active compounds directly at the desired point.

5.4. Drugs Currently in Clinical Trials

In recent years, multiple treatments to restore autophagy have entered clinical testing phases [188]. Ambroxol may be a promising one; firstly, used as a cough syrup, it can enhance GCase activity, permit better lysosomal function, and decrease reactive oxygen species (ROS) apparition [189]. A clinical phase II study is ongoing until 2023 in a single-center, double-blind, randomized, and placebo-controlled manner (NCT02914366) [190]. Based on the promising results of the AIM-PD phase 2 clinical trial (NCT02941822), the University College London is starting a phase clinical trial in 2023, gathering 330 people with PD across approximately 12 clinical centers in the United Kingdom [191]. Another GCase activator drug is in the clinical phase study under the LTI-291 name. This small molecule forms an active complex with GCase and has shown 130% increased activity of GCase in vitro. Phase I study performed in 2021 shows good tolerability and no adverse effects of LTI-291 administration [192]. Venglustat is also a drug that participates in GCase activity amelioration by inhibiting glucosylceramide synthase. Phases I and II performed by Peterschimtt and colleagues resulted in good safety of oral administration [193] and improved glucosylceramide levels in GBA-PD patients [194], without observation of severe adverse events.
As described previously, mTOR inhibition through rapamycin capacities can be an interesting strategy for Sirolimus medication. This drug, clinically tried in multiple system atrophy patients, does not show any benefit on pathological rating scales, neuroimaging, or blood biomarkers compared to controls, and adverse events were more frequent for treated patients [195]. Currently, no clinical trial with Sirolimus is planned for PD patients, but it opened the way for mTOR inhibition through rapamycin in neurodegenerative disorders.
Nilotinib or Tasigna® drug is a tyrosine kinase inhibitor enhancing autophagy activity [196], showing interesting results on the reduction, with a rate twice as low, of PD hallmark α-syn in the A53T mouse model [197]. Tested on clinical phase I and phase II studies, it proves its safety and tolerability, with no adverse effects observed and a decrease in oligomeric forms of the α-syn in PD patients [198,199]. However, a recent meta-analysis concluded that it had no effect on motor outcomes [200]. Pagan and colleagues support starting a phase 3 trial project to evaluate nilotinib 300 mg in a more extensive multicenter study [199].

6. Conclusions

At the same time, the growing knowledge of PD, the ALP mechanisms, and the link between this disease and this important cellular mechanism has allowed for the generation of animal models, promising new research and knowledge. Presently, multiple potential treatments to restore the critical mechanism of autophagy in PD are being clinically tested [188] and will soon provide answers on their capacity to act as disease modifiers in a pathology for which there is no curative treatment yet. In addition, the future should see the development of drugs targeting the mTOR pathway or other autophagy modulators, like TFEB and Beclin-1. The strategy of regulating autophagy to treat neurodegenerative disorders is an interesting avenue, with multiple and diverse possible solutions, from drugs to nanotechnologies by way of gene therapies. Of interest, TFEB-based therapy in the AD context has proven beneficial [201,202]. Beyond that, all these diseases have a dysfunction of the ALP pathway in common. As a result, any treatment developed for one of the neurodegenerative diseases may have an impact on the others. The growing interest in and improved knowledge of this mechanism in recent years should accelerate future discoveries, increasing the chances of rapidly seeing possible therapeutics for PD patients in the following years.

Author Contributions

R.K. did the literature search and drafted the manuscript; R.K. prepared the tables; B.D. reviewed the manuscript; all authors read and approved the submitted final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

R.K. is a recipient of a Clément Fayat Foundation fellowship (France). INSERM, CNRS, and the University of Bordeaux provided financial and infrastructural support. This study received financial support from the French government in the framework of the University of Bordeaux’s IdEx “Investments for the Future” program (GPR BRAIN_2030).

Acknowledgments

We apologize to the authors of several high-quality scientific articles that contributed significantly to the development of the field but could not be cited due to space limitations.

Conflicts of Interest

The authors declared no conflict of interest.

References

  1. Parkinson, J. An essay on the shaking palsy. 1817. J. Neuropsychiatry Clin. Neurosci. 2002, 14, 223–236; discussion 222. [Google Scholar] [CrossRef] [PubMed]
  2. Ehringer, H.; Hornykiewicz, O. Distribution of noradrenaline and dopamine (3-hydroxytyramine) in the human brain and their behavior in diseases of the extrapyramidal system. Klin. Wochenschr. 1960, 38, 1236–1239. [Google Scholar] [CrossRef] [PubMed]
  3. Weiss, J.L.; Ng, L.K.; Chase, T.N. Long-lasting dyskinesia induced by levodopa. Lancet 1971, 1, 1016–1017. [Google Scholar] [CrossRef]
  4. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. Alpha-synuclein in Lewy bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef] [PubMed]
  5. 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 alpha-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [Green Version]
  6. Tan, Y.Y.; Jenner, P.; Chen, S.D. Monoamine Oxidase-B Inhibitors for the Treatment of Parkinson’s Disease: Past, Present, and Future. J. Parkinsons Dis. 2022, 12, 477–493. [Google Scholar] [CrossRef]
  7. Fabbri, M.; Ferreira, J.J.; Rascol, O. COMT Inhibitors in the Management of Parkinson’s Disease. CNS Drugs 2022, 36, 261–282. [Google Scholar] [CrossRef]
  8. Limousin, P.; Foltynie, T. Long-term outcomes of deep brain stimulation in Parkinson disease. Nat. Rev. Neurol. 2019, 15, 234–242. [Google Scholar] [CrossRef] [Green Version]
  9. GBD 2016 Parkinson’s Disease Collaborators. 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] [Green Version]
  10. Rocca, W.A. The burden of Parkinson’s disease: A worldwide perspective. Lancet Neurol. 2018, 17, 928–929. [Google Scholar] [CrossRef]
  11. Dorsey, E.R.; Sherer, T.; Okun, M.S.; Bloem, B.R. The Emerging Evidence of the Parkinson Pandemic. J. Parkinsons Dis. 2018, 8, S3–S8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Aubignat, M.; Tir, M.; Krystkowiak, P. Non-motor symptoms of Parkinson’s disease from pathophysiology to early diagnosis. Rev. Med. Interne 2021, 42, 251–257. [Google Scholar] [CrossRef] [PubMed]
  13. Blesa, J.; Foffani, G.; Dehay, B.; Bezard, E.; Obeso, J.A. Motor and non-motor circuit disturbances in early Parkinson disease: Which happens first? Nat. Rev. Neurosci. 2022, 23, 115–128. [Google Scholar] [CrossRef]
  14. Kalia, L.V.; Lang, A.E. Parkinson’s disease. Lancet 2015, 386, 896–912. [Google Scholar] [CrossRef] [PubMed]
  15. Arber, S.; Costa, R.M. Networking brainstem and basal ganglia circuits for movement. Nat. Rev. Neurosci. 2022, 23, 342–360. [Google Scholar] [CrossRef]
  16. McGregor, M.M.; Nelson, A.B. Circuit Mechanisms of Parkinson’s Disease. Neuron 2019, 101, 1042–1056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Braak, H.; Del Tredici, K.; Rub, U.; de Vos, R.A.; Jansen Steur, E.N.; Braak, E. Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol. Aging 2003, 24, 197–211. [Google Scholar] [CrossRef]
  18. Dehay, B.; Bourdenx, M.; Gorry, P.; Przedborski, S.; Vila, M.; Hunot, S.; Singleton, A.; Olanow, C.W.; Merchant, K.M.; Bezard, E.; et al. Targeting alpha-synuclein for treatment of Parkinson’s disease: Mechanistic and therapeutic considerations. Lancet Neurol. 2015, 14, 855–866. [Google Scholar] [CrossRef] [Green Version]
  19. Wakabayashi, K.; Tanji, K.; Odagiri, S.; Miki, Y.; Mori, F.; Takahashi, H. The Lewy body in Parkinson’s disease and related neurodegenerative disorders. Mol. Neurobiol. 2013, 47, 495–508. [Google Scholar] [CrossRef]
  20. Gai, W.P.; Yuan, H.X.; Li, X.Q.; Power, J.T.; Blumbergs, P.C.; Jensen, P.H. In situ and in vitro study of colocalization and segregation of alpha-synuclein, ubiquitin, and lipids in Lewy bodies. Exp. Neurol. 2000, 166, 324–333. [Google Scholar] [CrossRef]
  21. Fares, M.B.; Jagannath, S.; Lashuel, H.A. Reverse engineering Lewy bodies: How far have we come and how far can we go? Nat. Rev. Neurosci. 2021, 22, 111–131. [Google Scholar] [CrossRef] [PubMed]
  22. Forno, L.S. Neuropathology of Parkinson’s disease. J. Neuropathol. Exp. Neurol. 1996, 55, 259–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Emamzadeh, F.N.; Surguchov, A. Parkinson’s Disease: Biomarkers, Treatment, and Risk Factors. Front. Neurosci. 2018, 12, 612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. 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]
  25. 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]
  26. Plowey, E.D.; Cherra, S.J., 3rd; Liu, Y.J.; Chu, C.T. Role of autophagy in G2019S-LRRK2-associated neurite shortening in differentiated SH-SY5Y cells. J. Neurochem. 2008, 105, 1048–1056. [Google Scholar] [CrossRef] [Green Version]
  27. Zavodszky, E.; Seaman, M.N.; 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]
  28. Takao, M.; Aoyama, M.; Ishikawa, K.; Sakiyama, Y.; Yomono, H.; Saito, Y.; Kurisaki, H.; Mihara, B.; Murayama, S. Spinocerebellar ataxia type 2 is associated with Parkinsonism and Lewy body pathology. BMJ Case Rep. 2011, 2011. [Google Scholar] [CrossRef] [Green Version]
  29. Marcelo, A.; Afonso, I.T.; Afonso-Reis, R.; Brito, D.V.C.; Costa, R.G.; Rosa, A.; Alves-Cruzeiro, J.; Ferreira, B.; Henriques, C.; Nobre, R.J.; et al. Autophagy in Spinocerebellar ataxia type 2, a dysregulated pathway, and a target for therapy. Cell Death Dis. 2021, 12, 1117. [Google Scholar] [CrossRef]
  30. Grotzsch, H.; Pizzolato, G.P.; Ghika, J.; Schorderet, D.; Vingerhoets, F.J.; Landis, T.; Burkhard, P.R. Neuropathology of a case of dopa-responsive dystonia associated with a new genetic locus, DYT14. Neurology 2002, 58, 1839–1842. [Google Scholar] [CrossRef]
  31. Cookson, M.R.; Lockhart, P.J.; McLendon, C.; O’Farrell, C.; Schlossmacher, M.; Farrer, M.J. RING finger 1 mutations in Parkin produce altered localization of the protein. Hum. Mol. Genet. 2003, 12, 2957–2965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Narendra, D.; Tanaka, A.; Suen, D.F.; Youle, R.J. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J. Cell Biol. 2008, 183, 795–803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Gandhi, S.; Muqit, M.M.; Stanyer, L.; Healy, D.G.; Abou-Sleiman, P.M.; Hargreaves, I.; Heales, S.; Ganguly, M.; Parsons, L.; Lees, A.J.; et al. PINK1 protein in normal human brain and Parkinson’s disease. Brain 2006, 129, 1720–1731. [Google Scholar] [CrossRef] [PubMed]
  34. Dagda, R.K.; Cherra, S.J., 3rd; Kulich, S.M.; Tandon, A.; Park, D.; Chu, C.T. Loss of PINK1 function promotes mitophagy through effects on oxidative stress and mitochondrial fission. J. Biol. Chem. 2009, 284, 13843–13855. [Google Scholar] [CrossRef] [Green Version]
  35. Taipa, R.; Pereira, C.; Reis, I.; Alonso, I.; Bastos-Lima, A.; Melo-Pires, M.; Magalhaes, M. DJ-1 linked parkinsonism (PARK7) is associated with Lewy body pathology. Brain 2016, 139, 1680–1687. [Google Scholar] [CrossRef]
  36. Nash, Y.; Schmukler, E.; Trudler, D.; Pinkas-Kramarski, R.; Frenkel, D. DJ-1 deficiency impairs autophagy and reduces alpha-synuclein phagocytosis by microglia. J. Neurochem. 2017, 143, 584–594. [Google Scholar] [CrossRef] [Green Version]
  37. 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]
  38. Dehay, B.; Ramirez, A.; Martinez-Vicente, M.; Perier, C.; Canron, M.H.; Doudnikoff, E.; Vital, A.; Vila, M.; Klein, C.; Bezard, E. Loss of P-type ATPase ATP13A2/PARK9 function induces general lysosomal deficiency and leads to Parkinson disease neurodegeneration. Proc. Natl. Acad. Sci. USA 2012, 109, 9611–9616. [Google Scholar] [CrossRef] [Green Version]
  39. Tsuboi, Y.; Mishima, T.; Fujioka, S. Perry Disease: Concept of a New Disease and Clinical Diagnostic Criteria. J. Mov. Disord. 2021, 14, 1–9. [Google Scholar] [CrossRef]
  40. Ishikawa, K.I.; Saiki, S.; Furuya, N.; Imamichi, Y.; Tsuboi, Y.; Hattori, N. p150(glued) deficiency impairs effective fusion between autophagosomes and lysosomes due to their redistribution to the cell periphery. Neurosci. Lett. 2019, 690, 181–187. [Google Scholar] [CrossRef]
  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, e36458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Vilarino-Guell, C.; Rajput, A.; Milnerwood, A.J.; Shah, B.; Szu-Tu, C.; Trinh, J.; Yu, I.; Encarnacion, M.; Munsie, L.N.; Tapia, L.; et al. DNAJC13 mutations in Parkinson disease. Hum. Mol. Genet. 2014, 23, 1794–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chartier-Harlin, M.C.; Dachsel, J.C.; Vilarino-Guell, C.; Lincoln, S.J.; Lepretre, F.; Hulihan, M.M.; Kachergus, J.; Milnerwood, A.J.; Tapia, L.; Song, M.S.; et al. Translation initiator EIF4G1 mutations in familial Parkinson disease. Am. J. Hum. Genet. 2011, 89, 398–406. [Google Scholar] [CrossRef] [Green Version]
  44. Ramirez-Valle, F.; Braunstein, S.; Zavadil, J.; Formenti, S.C.; Schneider, R.J. eIF4GI links nutrient sensing by mTOR to cell proliferation and inhibition of autophagy. J. Cell Biol. 2008, 181, 293–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Burchell, V.S.; Nelson, D.E.; Sanchez-Martinez, A.; Delgado-Camprubi, M.; Ivatt, R.M.; Pogson, J.H.; Randle, S.J.; Wray, S.; Lewis, P.A.; Houlden, H.; et al. The Parkinson’s disease-linked proteins Fbxo7 and Parkin interact to mediate mitophagy. Nat. Neurosci. 2013, 16, 1257–1265. [Google Scholar] [CrossRef] [Green Version]
  46. Strauss, K.M.; Martins, L.M.; Plun-Favreau, H.; Marx, F.P.; Kautzmann, S.; Berg, D.; Gasser, T.; Wszolek, Z.; Muller, T.; Bornemann, A.; et al. Loss of function mutations in the gene encoding Omi/HtrA2 in Parkinson’s disease. Hum. Mol. Genet. 2005, 14, 2099–2111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Rodrigue-Gervais, I.G.; Doiron, K.; Champagne, C.; Mayes, L.; Leiva-Torres, G.A.; Vanie, P., Jr.; Douglas, T.; Vidal, S.M.; Alnemri, E.S.; Saleh, M. The mitochondrial protease HtrA2 restricts the NLRP3 and AIM2 inflammasomes. Sci. Rep. 2018, 8, 8446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Magrinelli, F.; Mehta, S.; Di Lazzaro, G.; Latorre, A.; Edwards, M.J.; Balint, B.; Basu, P.; Kobylecki, C.; Groppa, S.; Hegde, A.; et al. Dissecting the Phenotype and Genotype of PLA2G6-Related Parkinsonism. Mov. Disord. 2022, 37, 148–161. [Google Scholar] [CrossRef]
  49. Zhou, Q.; Yen, A.; Rymarczyk, G.; Asai, H.; Trengrove, C.; Aziz, N.; Kirber, M.T.; Mostoslavsky, G.; Ikezu, T.; Wolozin, B.; et al. Impairment of PARK14-dependent Ca(2+) signalling is a novel determinant of Parkinson’s disease. Nat. Commun. 2016, 7, 10332. [Google Scholar] [CrossRef] [Green Version]
  50. 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, 385. [Google Scholar] [CrossRef]
  51. Kuru, S.; Yoshida, M.; Tatsumi, S.; Mimuro, M. Immunohistochemical localization of spatacsin in alpha-synucleinopathies. Neuropathology 2014, 34, 135–139. [Google Scholar] [CrossRef] [PubMed]
  52. Varga, R.E.; Khundadze, M.; Damme, M.; Nietzsche, S.; Hoffmann, B.; Stauber, T.; Koch, N.; Hennings, J.C.; Franzka, P.; Huebner, A.K.; et al. In Vivo Evidence for Lysosome Depletion and Impaired Autophagic Clearance in Hereditary Spastic Paraplegia Type SPG11. PLoS Genet. 2015, 11, e1005454. [Google Scholar] [CrossRef] [PubMed]
  53. Ikeda, A.; Nishioka, K.; Meng, H.; Takanashi, M.; Hasegawa, I.; Inoshita, T.; Shiba-Fukushima, K.; Li, Y.; Yoshino, H.; Mori, A.; et al. Mutations in CHCHD2 cause alpha-synuclein aggregation. Hum. Mol. Genet. 2019, 28, 3895–3911. [Google Scholar] [CrossRef]
  54. Quadri, M.; Mandemakers, W.; Grochowska, M.M.; Masius, R.; Geut, H.; Fabrizio, E.; Breedveld, G.J.; Kuipers, D.; Minneboo, M.; Vergouw, L.J.M.; et al. LRP10 genetic variants in familial Parkinson’s disease and dementia with Lewy bodies: A genome-wide linkage and sequencing study. Lancet Neurol. 2018, 17, 597–608. [Google Scholar] [CrossRef] [PubMed]
  55. Wilson, G.R.; Sim, J.C.; McLean, C.; Giannandrea, M.; Galea, C.A.; Riseley, J.R.; Stephenson, S.E.; Fitzpatrick, E.; Haas, S.A.; Pope, K.; et al. Mutations in RAB39B cause X-linked intellectual disability and early-onset Parkinson disease with alpha-synuclein pathology. Am. J. Hum. Genet. 2014, 95, 729–735. [Google Scholar] [CrossRef] [Green Version]
  56. Niu, M.; Zheng, N.; Wang, Z.; Gao, Y.; Luo, X.; Chen, Z.; Fu, X.; Wang, Y.; Wang, T.; Liu, M.; et al. RAB39B Deficiency Impairs Learning and Memory Partially Through Compromising Autophagy. Front. Cell Dev. Biol. 2020, 8, 598622. [Google Scholar] [CrossRef] [PubMed]
  57. Arasaratnam, C.J.; Singh-Bains, M.K.; Waldvogel, H.J.; Faull, R.L.M. Neuroimaging and neuropathology studies of X-linked dystonia parkinsonism. Neurobiol. Dis. 2021, 148, 105186. [Google Scholar] [CrossRef]
  58. Deng, H.X.; Shi, Y.; Yang, Y.; Ahmeti, K.B.; Miller, N.; Huang, C.; Cheng, L.; Zhai, H.; Deng, S.; Nuytemans, K.; et al. Identification of TMEM230 mutations in familial Parkinson’s disease. Nat. Genet. 2016, 48, 733–739. [Google Scholar] [CrossRef]
  59. Kim, M.J.; Deng, H.X.; Wong, Y.C.; Siddique, T.; Krainc, D. The Parkinson’s disease-linked protein TMEM230 is required for Rab8a-mediated secretory vesicle trafficking and retromer trafficking. Hum. Mol. Genet. 2017, 26, 729–741. [Google Scholar] [CrossRef] [Green Version]
  60. 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]
  61. Jinn, S.; Drolet, R.E.; Cramer, P.E.; Wong, A.H.; 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 alpha-synuclein aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 2389–2394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Cang, C.; Aranda, K.; Seo, Y.J.; Gasnier, B.; Ren, D. TMEM175 Is an Organelle K(+) Channel Regulating Lysosomal Function. Cell 2015, 162, 1101–1112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Moors, T.; Paciotti, S.; Chiasserini, D.; Calabresi, P.; Parnetti, L.; Beccari, T.; van de Berg, W.D. Lysosomal Dysfunction and alpha-Synuclein Aggregation in Parkinson’s Disease: Diagnostic Links. Mov. Disord. 2016, 31, 791–801. [Google Scholar] [CrossRef]
  64. Kiffin, R.; Bandyopadhyay, U.; Cuervo, A.M. Oxidative stress and autophagy. Antioxid. Redox Signal. 2006, 8, 152–162. [Google Scholar] [CrossRef] [PubMed]
  65. Scherz-Shouval, R.; Elazar, Z. ROS, mitochondria and the regulation of autophagy. Trends Cell Biol. 2007, 17, 422–427. [Google Scholar] [CrossRef]
  66. Arotcarena, M.L.; Teil, M.; Dehay, B. Autophagy in Synucleinopathy: The Overwhelmed and Defective Machinery. Cells 2019, 8, 565. [Google Scholar] [CrossRef] [Green Version]
  67. Bourdenx, M.; Dehay, B. What lysosomes actually tell us about Parkinson’s disease? Ageing Res. Rev. 2016, 32, 140–149. [Google Scholar] [CrossRef]
  68. Klionsky, D.J.; Eskelinen, E.L. The vacuole versus the lysosome: When size matters. Autophagy 2014, 10, 185–187. [Google Scholar] [CrossRef] [Green Version]
  69. Braulke, T.; Bonifacino, J.S. Sorting of lysosomal proteins. Biochim. Biophys. Acta 2009, 1793, 605–614. [Google Scholar] [CrossRef] [Green Version]
  70. Yang, C.; Wang, X. Lysosome biogenesis: Regulation and functions. J. Cell Biol. 2021, 220, e202102001. [Google Scholar] [CrossRef]
  71. Chapel, A.; Kieffer-Jaquinod, S.; Sagne, C.; Verdon, Q.; Ivaldi, C.; Mellal, M.; Thirion, J.; Jadot, M.; Bruley, C.; Garin, J.; et al. An extended proteome map of the lysosomal membrane reveals novel potential transporters. Mol. Cell Proteom. 2013, 12, 1572–1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Collier, J.J.; Guissart, C.; Olahova, M.; Sasorith, S.; Piron-Prunier, F.; Suomi, F.; Zhang, D.; Martinez-Lopez, N.; Leboucq, N.; Bahr, A.; et al. Developmental Consequences of Defective ATG7-Mediated Autophagy in Humans. N. Engl. J. Med. 2021, 384, 2406–2417. [Google Scholar] [CrossRef]
  73. Kim, M.; Sandford, E.; Gatica, D.; Qiu, Y.; Liu, X.; Zheng, Y.; Schulman, B.A.; Xu, J.; Semple, I.; Ro, S.H.; et al. Mutation in ATG5 reduces autophagy and leads to ataxia with developmental delay. eLife 2016, 5, e12245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Komatsu, M.; Wang, Q.J.; Holstein, G.R.; Friedrich, V.L., Jr.; Iwata, J.; Kominami, E.; Chait, B.T.; Tanaka, K.; Yue, Z. Essential role for autophagy protein Atg7 in the maintenance of axonal homeostasis and the prevention of axonal degeneration. Proc. Natl. Acad. Sci. USA 2007, 104, 14489–14494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Axe, E.L.; Walker, S.A.; Manifava, M.; Chandra, P.; Roderick, H.L.; Habermann, A.; Griffiths, G.; Ktistakis, N.T. Autophagosome formation from membrane compartments enriched in phosphatidylinositol 3-phosphate and dynamically connected to the endoplasmic reticulum. J. Cell Biol. 2008, 182, 685–701. [Google Scholar] [CrossRef] [Green Version]
  76. Ge, L.; Melville, D.; Zhang, M.; Schekman, R. The ER-Golgi intermediate compartment is a key membrane source for the LC3 lipidation step of autophagosome biogenesis. eLife 2013, 2, e00947. [Google Scholar] [CrossRef]
  77. Puri, C.; Renna, M.; Bento, C.F.; Moreau, K.; Rubinsztein, D.C. Diverse autophagosome membrane sources coalesce in recycling endosomes. Cell 2013, 154, 1285–1299. [Google Scholar] [CrossRef] [Green Version]
  78. van der Vaart, A.; Griffith, J.; Reggiori, F. Exit from the Golgi is required for the expansion of the autophagosomal phagophore in yeast Saccharomyces cerevisiae. Mol. Biol. Cell 2010, 21, 2270–2284. [Google Scholar] [CrossRef] [Green Version]
  79. Hosokawa, N.; Hara, T.; Kaizuka, T.; Kishi, C.; Takamura, A.; Miura, Y.; Iemura, S.; Natsume, T.; Takehana, K.; Yamada, N.; et al. Nutrient-dependent mTORC1 association with the ULK1-Atg13-FIP200 complex required for autophagy. Mol. Biol. Cell 2009, 20, 1981–1991. [Google Scholar] [CrossRef] [Green Version]
  80. Hosokawa, N.; Sasaki, T.; Iemura, S.; Natsume, T.; Hara, T.; Mizushima, N. Atg101, a novel mammalian autophagy protein interacting with Atg13. Autophagy 2009, 5, 973–979. [Google Scholar] [CrossRef]
  81. Jung, C.H.; Jun, C.B.; Ro, S.H.; Kim, Y.M.; Otto, N.M.; Cao, J.; Kundu, M.; Kim, D.H. ULK-Atg13-FIP200 complexes mediate mTOR signaling to the autophagy machinery. Mol. Biol. Cell 2009, 20, 1992–2003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Mercer, C.A.; Kaliappan, A.; Dennis, P.B. A novel, human Atg13 binding protein, Atg101, interacts with ULK1 and is essential for macroautophagy. Autophagy 2009, 5, 649–662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Russell, R.C.; Tian, Y.; Yuan, H.; Park, H.W.; Chang, Y.Y.; Kim, J.; Kim, H.; Neufeld, T.P.; Dillin, A.; Guan, K.L. ULK1 induces autophagy by phosphorylating Beclin-1 and activating VPS34 lipid kinase. Nat. Cell Biol. 2013, 15, 741–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Gubas, A.; Dikic, I. A guide to the regulation of selective autophagy receptors. FEBS J. 2022, 289, 75–89. [Google Scholar] [CrossRef]
  85. Galluzzi, L.; Baehrecke, E.H.; Ballabio, A.; Boya, P.; Bravo-San Pedro, J.M.; Cecconi, F.; Choi, A.M.; Chu, C.T.; Codogno, P.; Colombo, M.I.; et al. Molecular definitions of autophagy and related processes. EMBO J. 2017, 36, 1811–1836. [Google Scholar] [CrossRef]
  86. Maday, S.; Wallace, K.E.; Holzbaur, E.L. Autophagosomes initiate distally and mature during transport toward the cell soma in primary neurons. J. Cell Biol. 2012, 196, 407–417. [Google Scholar] [CrossRef]
  87. Lorincz, P.; Juhasz, G. Autophagosome-Lysosome Fusion. J. Mol. Biol. 2020, 432, 2462–2482. [Google Scholar] [CrossRef]
  88. Schneider, J.L.; Suh, Y.; Cuervo, A.M. Deficient chaperone-mediated autophagy in liver leads to metabolic dysregulation. Cell Metab. 2014, 20, 417–432. [Google Scholar] [CrossRef] [Green Version]
  89. Ferreira, J.V.; Fofo, H.; Bejarano, E.; Bento, C.F.; Ramalho, J.S.; Girao, H.; Pereira, P. STUB1/CHIP is required for HIF1A degradation by chaperone-mediated autophagy. Autophagy 2013, 9, 1349–1366. [Google Scholar] [CrossRef] [Green Version]
  90. Park, C.; Suh, Y.; Cuervo, A.M. Regulated degradation of Chk1 by chaperone-mediated autophagy in response to DNA damage. Nat. Commun. 2015, 6, 6823. [Google Scholar] [CrossRef]
  91. Kirchner, P.; Bourdenx, M.; Madrigal-Matute, J.; Tiano, S.; Diaz, A.; Bartholdy, B.A.; Will, B.; Cuervo, A.M. Proteome-wide analysis of chaperone-mediated autophagy targeting motifs. PLoS Biol. 2019, 17, e3000301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Bonhoure, A.; Vallentin, A.; Martin, M.; Senff-Ribeiro, A.; Amson, R.; Telerman, A.; Vidal, M. Acetylation of translationally controlled tumor protein promotes its degradation through chaperone-mediated autophagy. Eur. J. Cell Biol. 2017, 96, 83–98. [Google Scholar] [CrossRef] [PubMed]
  93. Kaushik, S.; Cuervo, A.M. AMPK-dependent phosphorylation of lipid droplet protein PLIN2 triggers its degradation by CMA. Autophagy 2016, 12, 432–438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Lemasters, J.J. Variants of mitochondrial autophagy: Types 1 and 2 mitophagy and micromitophagy (Type 3). Redox Biol. 2014, 2, 749–754. [Google Scholar] [CrossRef] [Green Version]
  95. Soubannier, V.; McLelland, G.L.; Zunino, R.; Braschi, E.; Rippstein, P.; Fon, E.A.; McBride, H.M. A vesicular transport pathway shuttles cargo from mitochondria to lysosomes. Curr. Biol. 2012, 22, 135–141. [Google Scholar] [CrossRef] [Green Version]
  96. Manjithaya, R.; Nazarko, T.Y.; Farre, J.C.; Subramani, S. Molecular mechanism and physiological role of pexophagy. FEBS Lett. 2010, 584, 1367–1373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Loi, M.; Raimondi, A.; Morone, D.; Molinari, M. ESCRT-III-driven piecemeal micro-ER-phagy remodels the ER during recovery from ER stress. Nat. Commun. 2019, 10, 5058. [Google Scholar] [CrossRef] [Green Version]
  98. Otto, F.B.; Thumm, M. Mechanistic dissection of macro- and micronucleophagy. Autophagy 2021, 17, 626–639. [Google Scholar] [CrossRef]
  99. Tekirdag, K.; Cuervo, A.M. Chaperone-mediated autophagy and endosomal microautophagy: Joint by a chaperone. J. Biol. Chem. 2018, 293, 5414–5424. [Google Scholar] [CrossRef] [Green Version]
  100. Mesquita, A.; Glenn, J.; Jenny, A. Differential activation of eMI by distinct forms of cellular stress. Autophagy 2021, 17, 1828–1840. [Google Scholar] [CrossRef]
  101. Hara, T.; Nakamura, K.; Matsui, M.; Yamamoto, A.; Nakahara, Y.; Suzuki-Migishima, R.; Yokoyama, M.; Mishima, K.; Saito, I.; Okano, H.; et al. Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature 2006, 441, 885–889. [Google Scholar] [CrossRef]
  102. Komatsu, M.; Waguri, S.; Chiba, T.; Murata, S.; Iwata, J.; Tanida, I.; Ueno, T.; Koike, M.; Uchiyama, Y.; Kominami, E.; et al. Loss of autophagy in the central nervous system causes neurodegeneration in mice. Nature 2006, 441, 880–884. [Google Scholar] [CrossRef] [PubMed]
  103. Robak, L.A.; Jansen, I.E.; van Rooij, J.; Uitterlinden, A.G.; Kraaij, R.; Jankovic, J.; International Parkinson’s Disease Genomics Consortium; Heutink, P.; Shulman, J.M. Excessive burden of lysosomal storage disorder gene variants in Parkinson’s disease. Brain 2017, 140, 3191–3203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. 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]
  105. Martin, S.; Smolders, S.; Van den Haute, C.; Heeman, B.; van Veen, S.; Crosiers, D.; Beletchi, I.; Verstraeten, A.; Gossye, H.; Gelders, G.; et al. Mutated ATP10B increases Parkinson’s disease risk by compromising lysosomal glucosylceramide export. Acta Neuropathol. 2020, 139, 1001–1024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. 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]
  107. Liu, H.; Koros, C.; Strohaker, T.; Schulte, C.; Bozi, M.; Varvaresos, S.; Ibanez de Opakua, A.; Simitsi, A.M.; Bougea, A.; Voumvourakis, K.; et al. A Novel SNCA A30G Mutation Causes Familial Parkinson’s Disease. Mov. Disord. 2021, 36, 1624–1633. [Google Scholar] [CrossRef]
  108. Kruger, R.; Kuhn, W.; Muller, T.; Woitalla, D.; Graeber, M.; Kosel, S.; Przuntek, H.; Epplen, J.T.; Schols, L.; Riess, O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef]
  109. Zarranz, J.J.; Alegre, J.; Gomez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atares, B.; et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef]
  110. 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]
  111. Proukakis, C.; Dudzik, C.G.; Brier, T.; MacKay, D.S.; Cooper, J.M.; Millhauser, G.L.; Houlden, H.; Schapira, A.H. A novel alpha-synuclein missense mutation in Parkinson disease. Neurology 2013, 80, 1062–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Lesage, S.; Anheim, M.; Letournel, F.; Bousset, L.; Honore, A.; Rozas, N.; Pieri, L.; Madiona, K.; Durr, A.; Melki, R.; et al. G51D alpha-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Ann. Neurol. 2013, 73, 459–471. [Google Scholar] [CrossRef] [PubMed]
  113. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Poyhonen, M.; Paetau, A. Novel alpha-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]
  114. Yoshino, H.; Hirano, M.; Stoessl, A.J.; Imamichi, Y.; Ikeda, A.; Li, Y.; Funayama, M.; Yamada, I.; Nakamura, Y.; Sossi, V.; et al. Homozygous alpha-synuclein p.A53V in familial Parkinson’s disease. Neurobiol. Aging 2017, 57, 248.e7–248.e12. [Google Scholar] [CrossRef] [PubMed]
  115. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. alpha-Synuclein locus triplication causes Parkinson’s disease. Science 2003, 302, 841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Lei, Z.; Cao, G.; Wei, G. A30P mutant alpha-synuclein impairs autophagic flux by inactivating JNK signaling to enhance ZKSCAN3 activity in midbrain dopaminergic neurons. Cell Death Dis. 2019, 10, 133. [Google Scholar] [CrossRef] [Green Version]
  117. Matsuda, N.; Sato, S.; Shiba, K.; Okatsu, K.; Saisho, K.; Gautier, C.A.; Sou, Y.S.; Saiki, S.; Kawajiri, S.; Sato, F.; et al. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J. Cell Biol. 2010, 189, 211–221. [Google Scholar] [CrossRef]
  118. Geisler, S.; Holmstrom, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef]
  119. 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]
  120. Valente, E.M.; Abou-Sleiman, P.M.; Caputo, V.; Muqit, M.M.; Harvey, K.; Gispert, S.; Ali, Z.; Del Turco, D.; Bentivoglio, A.R.; Healy, D.G.; et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science 2004, 304, 1158–1160. [Google Scholar] [CrossRef]
  121. Tang, B.; Xiong, H.; Sun, P.; Zhang, Y.; Wang, D.; Hu, Z.; Zhu, Z.; Ma, H.; Pan, Q.; Xia, J.H.; et al. Association of PINK1 and DJ-1 confers digenic inheritance of early-onset Parkinson’s disease. Hum. Mol. Genet. 2006, 15, 1816–1825. [Google Scholar] [CrossRef] [PubMed]
  122. Simon, D.K.; Tanner, C.M.; Brundin, P. Parkinson Disease Epidemiology, Pathology, Genetics, and Pathophysiology. Clin. Geriatr. Med. 2020, 36, 1–12. [Google Scholar] [CrossRef] [PubMed]
  123. Mamais, A.; Kluss, J.H.; Bonet-Ponce, L.; Landeck, N.; Langston, R.G.; Smith, N.; Beilina, A.; Kaganovich, A.; Ghosh, M.C.; Pellegrini, L.; et al. Mutations in LRRK2 linked to Parkinson disease sequester Rab8a to damaged lysosomes and regulate transferrin-mediated iron uptake in microglia. PLoS Biol. 2021, 19, e3001480. [Google Scholar] [CrossRef] [PubMed]
  124. Herbst, S.; Campbell, P.; Harvey, J.; Bernard, E.M.; Papayannopoulos, V.; Wood, N.W.; Morris, H.R.; Gutierrez, M.G. LRRK2 activation controls the repair of damaged endomembranes in macrophages. EMBO J. 2020, 39, e104494. [Google Scholar] [CrossRef] [PubMed]
  125. Singleton, A.B. Altered alpha-synuclein homeostasis causing Parkinson’s disease: The potential roles of dardarin. Trends Neurosci. 2005, 28, 416–421. [Google Scholar] [CrossRef] [PubMed]
  126. Rivero-Rios, P.; Romo-Lozano, M.; Madero-Perez, J.; Thomas, A.P.; Biosa, A.; Greggio, E.; Hilfiker, S. The G2019S variant of leucine-rich repeat kinase 2 (LRRK2) alters endolysosomal trafficking by impairing the function of the GTPase RAB8A. J. Biol. Chem. 2019, 294, 4738–4758. [Google Scholar] [CrossRef] [Green Version]
  127. Wulansari, N.; Darsono, W.H.W.; Woo, H.J.; Chang, M.Y.; Kim, J.; Bae, E.J.; Sun, W.; Lee, J.H.; Cho, I.J.; 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]
  128. Smith, L.; Schapira, A.H.V. GBA Variants and Parkinson Disease: Mechanisms and Treatments. Cells 2022, 11, 1261. [Google Scholar] [CrossRef]
  129. 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]
  130. Wie, J.; Liu, Z.; Song, H.; Tropea, T.F.; Yang, L.; Wang, H.; Liang, Y.; Cang, C.; Aranda, K.; Lohmann, J.; et al. A growth-factor-activated lysosomal K(+) channel regulates Parkinson’s pathology. Nature 2021, 591, 431–437. [Google Scholar] [CrossRef]
  131. 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.e2220. [Google Scholar] [CrossRef] [PubMed]
  132. 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]
  133. Krohn, L.; Ozturk, 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] [PubMed]
  134. Palomba, N.P.; Fortunato, G.; Pepe, G.; Modugno, N.; Pietracupa, S.; Damiano, I.; Mascio, G.; Carrillo, F.; Di Giovannantonio, L.G.; Ianiro, L.; et al. Common and Rare Variants in TMEM175 Gene Concur to the Pathogenesis of Parkinson’s Disease in Italian Patients. Mol. Neurobiol. 2023. [Google Scholar] [CrossRef]
  135. Pan, P.Y.; Sheehan, P.; Wang, Q.; Zhu, X.; Zhang, Y.; Choi, I.; Li, X.; Saenz, J.; Zhu, J.; Wang, J.; et al. Synj1 haploinsufficiency causes dopamine neuron vulnerability and alpha-synuclein accumulation in mice. Hum. Mol. Genet. 2020, 29, 2300–2312. [Google Scholar] [CrossRef]
  136. Nishiyama, J.; Miura, E.; Mizushima, N.; Watanabe, M.; Yuzaki, M. Aberrant membranes and double-membrane structures accumulate in the axons of Atg5-null Purkinje cells before neuronal death. Autophagy 2007, 3, 591–596. [Google Scholar] [CrossRef] [Green Version]
  137. Tu, H.Y.; Yuan, B.S.; Hou, X.O.; Zhang, X.J.; Pei, C.S.; Ma, Y.T.; Yang, Y.P.; Fan, Y.; Qin, Z.H.; Liu, C.F.; et al. alpha-synuclein suppresses microglial autophagy and promotes neurodegeneration in a mouse model of Parkinson’s disease. Aging Cell 2021, 20, e13522. [Google Scholar] [CrossRef]
  138. Cataldi, S.; Follett, J.; Fox, J.D.; Tatarnikov, I.; Kadgien, C.; Gustavsson, E.K.; Khinda, J.; Milnerwood, A.J.; Farrer, M.J. Altered dopamine release and monoamine transporters in Vps35 p.D620N knock-in mice. NPJ Parkinsons Dis. 2018, 4, 27. [Google Scholar] [CrossRef] [Green Version]
  139. Niu, M.; Zhao, F.; Bondelid, K.; Siedlak, S.L.; Torres, S.; Fujioka, H.; Wang, W.; Liu, J.; Zhu, X. VPS35 D620N knockin mice recapitulate cardinal features of Parkinson’s disease. Aging Cell 2021, 20, e13347. [Google Scholar] [CrossRef]
  140. Ramonet, D.; Daher, J.P.; Lin, B.M.; Stafa, K.; Kim, J.; Banerjee, R.; Westerlund, M.; Pletnikova, O.; Glauser, L.; Yang, L.; et al. Dopaminergic neuronal loss, reduced neurite complexity and autophagic abnormalities in transgenic mice expressing G2019S mutant LRRK2. PLoS ONE 2011, 6, e18568. [Google Scholar] [CrossRef]
  141. Arbez, N.; He, X.; Huang, Y.; Ren, M.; Liang, Y.; Nucifora, F.C.; Wang, X.; Pei, Z.; Tessarolo, L.; Smith, W.W.; et al. G2019S-LRRK2 mutation enhances MPTP-linked Parkinsonism in mice. Hum. Mol. Genet. 2020, 29, 580–590. [Google Scholar] [CrossRef] [PubMed]
  142. Carballo-Carbajal, I.; Laguna, A.; Romero-Gimenez, J.; Cuadros, T.; Bove, J.; Martinez-Vicente, M.; Parent, A.; Gonzalez-Sepulveda, M.; Penuelas, N.; Torra, A.; et al. Brain tyrosinase overexpression implicates age-dependent neuromelanin production in Parkinson’s disease pathogenesis. Nat. Commun. 2019, 10, 973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Wray, S. Modelling neurodegenerative disease using brain organoids. Semin. Cell Dev. Biol. 2021, 111, 60–66. [Google Scholar] [CrossRef] [PubMed]
  144. Bai, X.; Wey, M.C.; Fernandez, E.; Hart, M.J.; Gelfond, J.; Bokov, A.F.; Rani, S.; Strong, R. Rapamycin improves motor function, reduces 4-hydroxynonenal adducted protein in brain, and attenuates synaptic injury in a mouse model of synucleinopathy. Pathobiol. Aging Age Relat. Dis. 2015, 5, 28743. [Google Scholar] [CrossRef] [Green Version]
  145. Malagelada, C.; Jin, Z.H.; Jackson-Lewis, V.; Przedborski, S.; Greene, L.A. Rapamycin protects against neuron death in in vitro and in vivo models of Parkinson’s disease. J. Neurosci. 2010, 30, 1166–1175. [Google Scholar] [CrossRef] [Green Version]
  146. Masini, D.; Bonito-Oliva, A.; Bertho, M.; Fisone, G. Inhibition of mTORC1 Signaling Reverts Cognitive and Affective Deficits in a Mouse Model of Parkinson’s Disease. Front. Neurol. 2018, 9, 208. [Google Scholar] [CrossRef] [Green Version]
  147. Saiki, S.; Sasazawa, Y.; Imamichi, Y.; Kawajiri, S.; Fujimaki, T.; Tanida, I.; Kobayashi, H.; Sato, F.; Sato, S.; Ishikawa, K.; et al. Caffeine induces apoptosis by enhancement of autophagy via PI3K/Akt/mTOR/p70S6K inhibition. Autophagy 2011, 7, 176–187. [Google Scholar] [CrossRef] [Green Version]
  148. Luan, Y.; Ren, X.; Zheng, W.; Zeng, Z.; Guo, Y.; Hou, Z.; Guo, W.; Chen, X.; Li, F.; Chen, J.F. Chronic Caffeine Treatment Protects Against alpha-Synucleinopathy by Reestablishing Autophagy Activity in the Mouse Striatum. Front. Neurosci. 2018, 12, 301. [Google Scholar] [CrossRef] [Green Version]
  149. Zia, A.; Farkhondeh, T.; Pourbagher-Shahri, A.M.; Samarghandian, S. The role of curcumin in aging and senescence: Molecular mechanisms. Biomed. Pharmacother. 2021, 134, 111119. [Google Scholar] [CrossRef]
  150. Lee, J.H.; Rao, M.V.; Yang, D.S.; Stavrides, P.; Im, E.; Pensalfini, A.; Huo, C.; Sarkar, P.; Yoshimori, T.; Nixon, R.A. Transgenic expression of a ratiometric autophagy probe specifically in neurons enables the interrogation of brain autophagy in vivo. Autophagy 2019, 15, 543–557. [Google Scholar] [CrossRef]
  151. Chen, L.; Huang, Y.; Yu, X.; Lu, J.; Jia, W.; Song, J.; Liu, L.; Wang, Y.; Huang, Y.; Xie, J.; et al. Corynoxine Protects Dopaminergic Neurons Through Inducing Autophagy and Diminishing Neuroinflammation in Rotenone-Induced Animal Models of Parkinson’s Disease. Front. Pharmacol. 2021, 12, 642900. [Google Scholar] [CrossRef] [PubMed]
  152. Oleksak, P.; Nepovimova, E.; Chrienova, Z.; Musilek, K.; Patocka, J.; Kuca, K. Contemporary mTOR inhibitor scaffolds to diseases breakdown: A patent review (2015–2021). Eur. J. Med. Chem. 2022, 238, 114498. [Google Scholar] [CrossRef] [PubMed]
  153. Li, R.; Lu, Y.; Zhang, Q.; Liu, W.; Yang, R.; Jiao, J.; Liu, J.; Gao, G.; Yang, H. Piperine promotes autophagy flux by P2RX4 activation in SNCA/alpha-synuclein-induced Parkinson disease model. Autophagy 2022, 18, 559–575. [Google Scholar] [CrossRef] [PubMed]
  154. Liu, J.; Chen, M.; Wang, X.; Wang, Y.; Duan, C.; Gao, G.; Lu, L.; Wu, X.; Wang, X.; Yang, H. Piperine induces autophagy by enhancing protein phosphotase 2A activity in a rotenone-induced Parkinson’s disease model. Oncotarget 2016, 7, 60823–60843. [Google Scholar] [CrossRef] [Green Version]
  155. Yu, J.S.; Cui, W. Proliferation, survival and metabolism: The role of PI3K/AKT/mTOR signalling in pluripotency and cell fate determination. Development 2016, 143, 3050–3060. [Google Scholar] [CrossRef] [Green Version]
  156. Scrivo, A.; Bourdenx, M.; Pampliega, O.; Cuervo, A.M. Selective autophagy as a potential therapeutic target for neurodegenerative disorders. Lancet Neurol. 2018, 17, 802–815. [Google Scholar] [CrossRef]
  157. Cui, H.; Kilpelainen, T.; Zouzoula, L.; Auno, S.; Trontti, K.; Kurvonen, S.; Norrbacka, S.; Hovatta, I.; Jensen, P.H.; Myohanen, T.T. Prolyl oligopeptidase inhibition reduces alpha-synuclein aggregation in a cellular model of multiple system atrophy. J. Cell Mol. Med. 2021, 25, 9634–9646. [Google Scholar] [CrossRef]
  158. Rostami, J.; Jantti, M.; Cui, H.; Rinne, M.K.; Kukkonen, J.P.; Falk, A.; Erlandsson, A.; Myohanen, T. Prolyl oligopeptidase inhibition by KYP-2407 increases alpha-synuclein fibril degradation in neuron-like cells. Biomed. Pharmacother. 2020, 131, 110788. [Google Scholar] [CrossRef]
  159. Savolainen, M.H.; Richie, C.T.; Harvey, B.K.; Mannisto, P.T.; Maguire-Zeiss, K.A.; Myohanen, T.T. The beneficial effect of a prolyl oligopeptidase inhibitor, KYP-2047, on alpha-synuclein clearance and autophagy in A30P transgenic mouse. Neurobiol. Dis. 2014, 68, 1–15. [Google Scholar] [CrossRef]
  160. Yang, G.; Li, J.; Cai, Y.; Yang, Z.; Li, R.; Fu, W. Glycyrrhizic Acid Alleviates 6-Hydroxydopamine and Corticosterone-Induced Neurotoxicity in SH-SY5Y Cells Through Modulating Autophagy. Neurochem. Res. 2018, 43, 1914–1926. [Google Scholar] [CrossRef]
  161. Santoro, M.; Maetzler, W.; Stathakos, P.; Martin, H.L.; Hobert, M.A.; Rattay, T.W.; Gasser, T.; Forrester, J.V.; Berg, D.; Tracey, K.J.; et al. In-vivo evidence that high mobility group box 1 exerts deleterious effects in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model and Parkinson’s disease which can be attenuated by glycyrrhizin. Neurobiol. Dis. 2016, 91, 59–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Rusmini, P.; Cortese, K.; Crippa, V.; Cristofani, R.; Cicardi, M.E.; Ferrari, V.; Vezzoli, G.; Tedesco, B.; Meroni, M.; Messi, E.; et al. Trehalose induces autophagy via lysosomal-mediated TFEB activation in models of motoneuron degeneration. Autophagy 2019, 15, 631–651. [Google Scholar] [CrossRef] [PubMed]
  163. Moon, S.H.; Kwon, Y.; Huh, Y.E.; Choi, H.J. Trehalose ameliorates prodromal non-motor deficits and aberrant protein accumulation in a rotenone-induced mouse model of Parkinson’s disease. Arch. Pharm. Res. 2022, 45, 417–432. [Google Scholar] [CrossRef]
  164. Howson, P.A.; Johnston, T.H.; Ravenscroft, P.; Hill, M.P.; Su, J.; Brotchie, J.M.; Koprich, J.B. Beneficial Effects of Trehalose on Striatal Dopaminergic Deficits in Rodent and Primate Models of Synucleinopathy in Parkinson’s Disease. J. Pharmacol. Exp. Ther. 2019, 369, 364–374. [Google Scholar] [CrossRef] [PubMed]
  165. Kakoty, V.; K, C.S.; Dubey, S.K.; Yang, C.H.; Taliyan, R. Neuroprotective Effects of Trehalose and Sodium Butyrate on Preformed Fibrillar Form of alpha-Synuclein-Induced Rat Model of Parkinson’s Disease. ACS Chem. Neurosci. 2021, 12, 2643–2660. [Google Scholar] [CrossRef] [PubMed]
  166. Cunha, A.; Gaubert, A.; Verget, J.; Thiolat, M.L.; Barthelemy, P.; Latxague, L.; Dehay, B. Trehalose-Based Nucleolipids as Nanocarriers for Autophagy Modulation: An In Vitro Study. Pharmaceutics 2022, 14, 857. [Google Scholar] [CrossRef]
  167. Saha, T. LAMP2A overexpression in breast tumors promotes cancer cell survival via chaperone-mediated autophagy. Autophagy 2012, 8, 1643–1656. [Google Scholar] [CrossRef] [Green Version]
  168. Xilouri, M.; Brekk, O.R.; Landeck, N.; Pitychoutis, P.M.; Papasilekas, T.; Papadopoulou-Daifoti, Z.; Kirik, D.; Stefanis, L. Boosting chaperone-mediated autophagy in vivo mitigates alpha-synuclein-induced neurodegeneration. Brain 2013, 136, 2130–2146. [Google Scholar] [CrossRef] [Green Version]
  169. Issa, A.R.; Sun, J.; Petitgas, C.; Mesquita, A.; Dulac, A.; Robin, M.; Mollereau, B.; Jenny, A.; Cherif-Zahar, B.; Birman, S. The lysosomal membrane protein LAMP2A promotes autophagic flux and prevents SNCA-induced Parkinson disease-like symptoms in the Drosophila brain. Autophagy 2018, 14, 1898–1910. [Google Scholar] [CrossRef] [Green Version]
  170. Qian, X.; Li, X.; Cai, Q.; Zhang, C.; Yu, Q.; Jiang, Y.; Lee, J.H.; Hawke, D.; Wang, Y.; Xia, Y.; et al. Phosphoglycerate Kinase 1 Phosphorylates Beclin1 to Induce Autophagy. Mol. Cell 2017, 65, 917–931.e916. [Google Scholar] [CrossRef]
  171. Spencer, B.; Potkar, R.; Trejo, M.; Rockenstein, E.; Patrick, C.; Gindi, R.; Adame, A.; Wyss-Coray, T.; Masliah, E. Beclin 1 gene transfer activates autophagy and ameliorates the neurodegenerative pathology in alpha-synuclein models of Parkinson’s and Lewy body diseases. J. Neurosci. 2009, 29, 13578–13588. [Google Scholar] [CrossRef] [Green Version]
  172. Settembre, C.; Di Malta, C.; Polito, V.A.; Garcia Arencibia, M.; Vetrini, F.; Erdin, S.; Erdin, S.U.; Huynh, T.; Medina, D.; Colella, P.; et al. TFEB links autophagy to lysosomal biogenesis. Science 2011, 332, 1429–1433. [Google Scholar] [CrossRef] [Green Version]
  173. Arotcarena, M.L.; Bourdenx, M.; Dutheil, N.; Thiolat, M.L.; Doudnikoff, E.; Dovero, S.; Ballabio, A.; Fernagut, P.O.; Meissner, W.G.; Bezard, E.; et al. Transcription factor EB overexpression prevents neurodegeneration in experimental synucleinopathies. JCI Insight 2019, 4, e129719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Decressac, M.; Mattsson, B.; Weikop, P.; Lundblad, M.; Jakobsson, J.; Bjorklund, A. TFEB-mediated autophagy rescues midbrain dopamine neurons from alpha-synuclein toxicity. Proc. Natl. Acad. Sci. USA 2013, 110, E1817–E1826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Lee, J.H.; McBrayer, M.K.; Wolfe, D.M.; Haslett, L.J.; Kumar, A.; Sato, Y.; Lie, P.P.; Mohan, P.; Coffey, E.E.; Kompella, U.; et al. Presenilin 1 Maintains Lysosomal Ca(2+) Homeostasis via TRPML1 by Regulating vATPase-Mediated Lysosome Acidification. Cell Rep. 2015, 12, 1430–1444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Lee, J.H.; Yang, D.S.; Goulbourne, C.N.; Im, E.; Stavrides, P.; Pensalfini, A.; Chan, H.; Bouchet-Marquis, C.; Bleiwas, C.; Berg, M.J.; et al. Faulty autolysosome acidification in Alzheimer’s disease mouse models induces autophagic build-up of Abeta in neurons, yielding senile plaques. Nat. Neurosci. 2022, 25, 688–701. [Google Scholar] [CrossRef]
  177. Bourdenx, M.; Daniel, J.; Genin, E.; Soria, F.N.; Blanchard-Desce, M.; Bezard, E.; Dehay, B. Nanoparticles restore lysosomal acidification defects: Implications for Parkinson and other lysosomal-related diseases. Autophagy 2016, 12, 472–483. [Google Scholar] [CrossRef] [Green Version]
  178. Cunha, A.; Gaubert, A.; Latxague, L.; Dehay, B. PLGA-Based Nanoparticles for Neuroprotective Drug Delivery in Neurodegenerative Diseases. Pharmaceutics 2021, 13, 1042. [Google Scholar] [CrossRef]
  179. Brouillard, M.; Barthelemy, P.; Dehay, B.; Crauste-Manciet, S.; Desvergnes, V. Nucleolipid Acid-Based Nanocarriers Restore Neuronal Lysosomal Acidification Defects. Front. Chem. 2021, 9, 736554. [Google Scholar] [CrossRef]
  180. Baltazar, G.C.; Guha, S.; Lu, W.; Lim, J.; Boesze-Battaglia, K.; Laties, A.M.; Tyagi, P.; Kompella, U.B.; Mitchell, C.H. Acidic nanoparticles are trafficked to lysosomes and restore an acidic lysosomal pH and degradative function to compromised ARPE-19 cells. PLoS ONE 2012, 7, e49635. [Google Scholar] [CrossRef]
  181. Lee, H.J.; Han, J.; Jang, Y.; Kim, S.J.; Park, J.H.; Seo, K.S.; Jeong, S.; Shin, S.; Lim, K.; Heo, J.Y.; et al. Docosahexaenoic acid prevents paraquat-induced reactive oxygen species production in dopaminergic neurons via enhancement of glutathione homeostasis. Biochem. Biophys. Res. Commun. 2015, 457, 95–100. [Google Scholar] [CrossRef] [PubMed]
  182. Prevot, G.; Soria, F.N.; Thiolat, M.L.; Daniel, J.; Verlhac, J.B.; Blanchard-Desce, M.; Bezard, E.; Barthelemy, P.; Crauste-Manciet, S.; Dehay, B. Harnessing Lysosomal pH through PLGA Nanoemulsion as a Treatment of Lysosomal-Related Neurodegenerative Diseases. Bioconjug. Chem. 2018, 29, 4083–4089. [Google Scholar] [CrossRef] [PubMed]
  183. Senturk, M.; Cakir, M.; Tekin, A.; Kucukkartallar, T.; Yildirim, M.A.; Alkan, S.; Findik, S. Comparison of primary repair and repair with polyglycolic acid coated tube in recurrent laryngeal nerve cuts (an experimental study). Am. J. Surg. 2020, 219, 632–636. [Google Scholar] [CrossRef] [PubMed]
  184. 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]
  185. Guha, S.; Liu, J.; Baltazar, G.; Laties, A.M.; Mitchell, C.H. Rescue of compromised lysosomes enhances degradation of photoreceptor outer segments and reduces lipofuscin-like autofluorescence in retinal pigmented epithelial cells. Adv. Exp. Med. Biol. 2014, 801, 105–111. [Google Scholar] [CrossRef] [Green Version]
  186. Xue, X.; Wang, L.R.; Sato, Y.; Jiang, Y.; Berg, M.; Yang, D.S.; Nixon, R.A.; Liang, X.J. Single-walled carbon nanotubes alleviate autophagic/lysosomal defects in primary glia from a mouse model of Alzheimer’s disease. Nano Lett. 2014, 14, 5110–5117. [Google Scholar] [CrossRef] [Green Version]
  187. Zhang, X.; Misra, S.K.; Moitra, P.; Zhang, X.; Jeong, S.J.; Stitham, J.; Rodriguez-Velez, A.; Park, A.; Yeh, Y.S.; Gillanders, W.E.; et al. Use of acidic nanoparticles to rescue macrophage lysosomal dysfunction in atherosclerosis. Autophagy 2022, 1–18. [Google Scholar] [CrossRef]
  188. Sanchez-Mirasierra, I.; Ghimire, S.; Hernandez-Diaz, S.; Soukup, S.F. Targeting Macroautophagy as a Therapeutic Opportunity to Treat Parkinson’s Disease. Front. Cell Dev. Biol. 2022, 10, 1–18. [Google Scholar] [CrossRef]
  189. Maegawa, G.H.; Tropak, M.B.; Buttner, J.D.; Rigat, B.A.; Fuller, M.; Pandit, D.; Tang, L.; Kornhaber, G.J.; Hamuro, Y.; Clarke, J.T.; et al. Identification and characterization of ambroxol as an enzyme enhancement agent for Gaucher disease. J. Biol. Chem. 2009, 284, 23502–23516. [Google Scholar] [CrossRef] [Green Version]
  190. Silveira, C.R.A.; MacKinley, J.; Coleman, K.; Li, Z.; Finger, E.; Bartha, R.; Morrow, S.A.; Wells, J.; Borrie, M.; Tirona, R.G.; et al. Ambroxol as a novel disease-modifying treatment for Parkinson’s disease dementia: Protocol for a single-centre, randomized, double-blind, placebo-controlled trial. BMC Neurol. 2019, 19, 20. [Google Scholar] [CrossRef]
  191. Mullin, S.; Smith, L.; Lee, K.; D’Souza, G.; Woodgate, P.; Elflein, J.; Hallqvist, J.; Toffoli, M.; Streeter, A.; Hosking, J.; et al. Ambroxol for the Treatment of Patients With Parkinson Disease With and Without Glucocerebrosidase Gene Mutations: A Nonrandomized, Noncontrolled Trial. JAMA Neurol. 2020, 77, 427–434. [Google Scholar] [CrossRef] [Green Version]
  192. den Heijer, J.M.; Kruithof, A.C.; van Amerongen, G.; de Kam, M.L.; Thijssen, E.; Grievink, H.W.; Moerland, M.; Walker, M.; Been, K.; Skerlj, R.; et al. A randomized single and multiple ascending dose study in healthy volunteers of LTI-291, a centrally penetrant glucocerebrosidase activator. Br. J. Clin. Pharmacol. 2021, 87, 3561–3573. [Google Scholar] [CrossRef] [PubMed]
  193. Peterschmitt, M.J.; Crawford, N.P.S.; Gaemers, S.J.M.; Ji, A.J.; Sharma, J.; Pham, T.T. Pharmacokinetics, Pharmacodynamics, Safety, and Tolerability of Oral Venglustat in Healthy Volunteers. Clin. Pharmacol. Drug Dev. 2021, 10, 86–98. [Google Scholar] [CrossRef] [PubMed]
  194. Peterschmitt, M.J.; Saiki, H.; Hatano, T.; Gasser, T.; Isaacson, S.H.; Gaemers, S.J.M.; Minini, P.; Saubadu, S.; Sharma, J.; Walbillic, S.; et al. Safety, Pharmacokinetics, and Pharmacodynamics of Oral Venglustat in Patients with Parkinson’s Disease and a GBA Mutation: Results from Part 1 of the Randomized, Double-Blinded, Placebo-Controlled MOVES-PD Trial. J. Parkinsons Dis. 2022, 12, 557–570. [Google Scholar] [CrossRef] [PubMed]
  195. Palma, J.A.; Martinez, J.; Millar Vernetti, P.; Ma, T.; Perez, M.A.; Zhong, J.; Qian, Y.; Dutta, S.; Maina, K.N.; Siddique, I.; et al. mTOR Inhibition with Sirolimus in Multiple System Atrophy: A Randomized, Double-Blind, Placebo-Controlled Futility Trial and 1-Year Biomarker Longitudinal Analysis. Mov. Disord. 2022, 37, 778–789. [Google Scholar] [CrossRef]
  196. Fu, J.; Yang, Y.; Zhu, L.; Chen, Y.; Liu, B. Unraveling the Roles of Protein Kinases in Autophagy: An Update on Small-Molecule Compounds for Targeted Therapy. J. Med. Chem. 2022, 65, 5870–5885. [Google Scholar] [CrossRef]
  197. Hebron, M.L.; Lonskaya, I.; Olopade, P.; Selby, S.T.; Pagan, F.; Moussa, C.E. Tyrosine Kinase Inhibition Regulates Early Systemic Immune Changes and Modulates the Neuroimmune Response in alpha-Synucleinopathy. J. Clin. Cell Immunol. 2014, 5, 259. [Google Scholar] [CrossRef] [Green Version]
  198. Pagan, F.L.; Hebron, M.L.; Wilmarth, B.; Torres-Yaghi, Y.; Lawler, A.; Mundel, E.E.; Yusuf, N.; Starr, N.J.; Arellano, J.; Howard, H.H.; et al. Pharmacokinetics and pharmacodynamics of a single dose Nilotinib in individuals with Parkinson’s disease. Pharmacol. Res. Perspect. 2019, 7, e00470. [Google Scholar] [CrossRef] [Green Version]
  199. 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]
  200. Xie, X.; Yuan, P.; Kou, L.; Chen, X.; Li, J.; Li, Y. Nilotinib in Parkinson’s disease: A systematic review and meta-analysis. Front. Aging Neurosci. 2022, 14, 996217. [Google Scholar] [CrossRef]
  201. Xu, Y.; Du, S.; Marsh, J.A.; Horie, K.; Sato, C.; Ballabio, A.; Karch, C.M.; Holtzman, D.M.; Zheng, H. TFEB regulates lysosomal exocytosis of tau and its loss of function exacerbates tau pathology and spreading. Mol. Psychiatry 2021, 26, 5925–5939. [Google Scholar] [CrossRef] [PubMed]
  202. Chun, Y.S.; Kim, M.Y.; Lee, S.Y.; Kim, M.J.; Hong, T.J.; Jeon, J.K.; Ganbat, D.; Kim, H.T.; Kim, S.S.; Kam, T.I.; et al. MEK1/2 inhibition rescues neurodegeneration by TFEB-mediated activation of autophagic lysosomal function in a model of Alzheimer’s Disease. Mol. Psychiatry 2022, 27, 4770–4780. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of different autophagic pathways. Macroautophagy is composed of (1) the initiation phase, (2) the elongation phase, (3) the separation and enclosure of the phagophore, (4) the maturation step, and (5) the fusion with the lysosome. Chaperone-mediated autophagy starts with (1) the recognition of the KFERQ motif by the Hsc70 complex, (2) leading the cargo to LAMP2A, which (3) translocates, unfolds, and (4) degrades the protein. Endo/microautophagy is characterized by cargo capture directly by the endosome or lysosome before lysosomal degradation.
Figure 1. Schematic representation of different autophagic pathways. Macroautophagy is composed of (1) the initiation phase, (2) the elongation phase, (3) the separation and enclosure of the phagophore, (4) the maturation step, and (5) the fusion with the lysosome. Chaperone-mediated autophagy starts with (1) the recognition of the KFERQ motif by the Hsc70 complex, (2) leading the cargo to LAMP2A, which (3) translocates, unfolds, and (4) degrades the protein. Endo/microautophagy is characterized by cargo capture directly by the endosome or lysosome before lysosomal degradation.
Cells 12 00621 g001
Figure 2. Schematic representation of the impact of mutated proteins on the autophagy mechanism (left panel) and the targets of different treatments on autophagy (right panel).
Figure 2. Schematic representation of the impact of mutated proteins on the autophagy mechanism (left panel) and the targets of different treatments on autophagy (right panel).
Cells 12 00621 g002
Table 1. Summary of familial PD-associated genes. This table gathers the names of genes for which mutations are reported in familial PD cases, their associated protein name, their inheritance type (AD: autosomal dominant; AR: autosomal recessive; XL-D: X-linked dominant), and their involvement in LB hallmark and ALP.
Table 1. Summary of familial PD-associated genes. This table gathers the names of genes for which mutations are reported in familial PD cases, their associated protein name, their inheritance type (AD: autosomal dominant; AR: autosomal recessive; XL-D: X-linked dominant), and their involvement in LB hallmark and ALP.
GeneProteinInheritanceLB ALP Involvement
SNCAα-synucleinADYes [5]Yes [24]
LRRK2Leucine-rich repeat kinase 2AD/ARYes [25]Yes [26]
VPS35Vacuolar protein sorting ortholog 35ADUnknownYes [27]
ATXN2Ataxin 2ADYes [28]Yes [29]
GCH1GTP cyckihydrolase 1ADNo [30]Unknown
PRKNParkin RBR E3 ubiquitin protein ligaseARNo [31]Yes [32]
PINK1PTEN-induced putative kinase 1ARYes [33]Yes [34]
PARK7Parkinsonism-associated deglycase/DJ1ARYes [35]Yes [36]
ATP13A2ATPase cation transporting 13A2ARNo [37]Yes [38]
DCTN1Dynactin subunit 1ADNo/few [39]Yes [40]
DNAJC6DnaJ Heat Shock Protein Family (Hsp40) Member C6ARUnknownYes [41]
DNAJC13DnaJ Heat Shock Protein Family (Hsp40) Member C13ADYes [42]Yes [42]
EIF4G1Eukaryotic translation initiation factor 4G1ADYes [43]Yes [44]
FBXO7F-Box Protein 7ARUnknownYes [45]
HTRA2HTRA serine peptidase 2ADYes [46]Yes [47]
PLA2G6Phospholipase A2 group 6ARYes [48]Yes [49]
SYNJ1Synaptojanin 1ARUnknownYes [50]
SPG11SpatacsinARYes [51]Yes [52]
CHCHD2Coiled-Coil-Helix-Coiled-Coil-Helix Domain Containing 2ADYes [53]Yes [53]
LRP10LDL Receptor-Related Protein 10ADYes [54]Unknown
RAB39BRas-related protein Rab-39BXL-DYes [55]Yes [56]
TAF1TATA-box binding protein associated factor 1XL-DNo [57]Unknown
TMEM230Transmembrane Protein 230ADYes [58]Yes [59]
UQCRC1Ubiquinol-Cytochrome C Reductase Core Protein 1ADUnknownUnknown
VPS13CVacuolar Protein Sorting 13 Homolog CARYes [60]Yes [60]
TMEM175Endosomal/lysosomal proton channel TMEM175UnknownYes [61]Yes [62]
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

Kinet, R.; Dehay, B. Pathogenic Aspects and Therapeutic Avenues of Autophagy in Parkinson’s Disease. Cells 2023, 12, 621. https://doi.org/10.3390/cells12040621

AMA Style

Kinet R, Dehay B. Pathogenic Aspects and Therapeutic Avenues of Autophagy in Parkinson’s Disease. Cells. 2023; 12(4):621. https://doi.org/10.3390/cells12040621

Chicago/Turabian Style

Kinet, Rémi, and Benjamin Dehay. 2023. "Pathogenic Aspects and Therapeutic Avenues of Autophagy in Parkinson’s Disease" Cells 12, no. 4: 621. https://doi.org/10.3390/cells12040621

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