Next Article in Journal
Phylogenetic Analysis Based on DNA Barcoding and Genetic Diversity Assessment of Morinda officinalis How in Vietnam Inferred by Microsatellites
Next Article in Special Issue
Special Issue “Parkinson’s Disease: Genetics and Pathogenesis”
Previous Article in Journal
Variation among the Complete Chloroplast Genomes of the Sumac Species Rhus chinensis: Reannotation and Comparative Analysis
Previous Article in Special Issue
Relationship between Substantia Nigra Neuromelanin Imaging and Dual Alpha-Synuclein Labeling of Labial Minor in Salivary Glands in Isolated Rapid Eye Movement Sleep Behavior Disorder and Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Human Induced Pluripotent Stem Cell Phenotyping and Preclinical Modeling of Familial Parkinson’s Disease

1
Southwest National Primate Research Center, Texas Biomedical Research Institute, San Antonio, TX 78227, USA
2
Cell Systems and Anatomy, San Antonio, TX 78229, USA
3
Department of Radiology, Long School of Medicine, University of Texas Health at San Antonio, San Antonio, TX 78229, USA
*
Author to whom correspondence should be addressed.
Genes 2022, 13(11), 1937; https://doi.org/10.3390/genes13111937
Submission received: 14 August 2022 / Revised: 13 October 2022 / Accepted: 18 October 2022 / Published: 25 October 2022
(This article belongs to the Special Issue Parkinson's Disease: Genetics and Pathogenesis)

Abstract

:
Parkinson’s disease (PD) is primarily idiopathic and a highly heterogenous neurodegenerative disease with patients experiencing a wide array of motor and non-motor symptoms. A major challenge for understanding susceptibility to PD is to determine the genetic and environmental factors that influence the mechanisms underlying the variations in disease-associated traits. The pathological hallmark of PD is the degeneration of dopaminergic neurons in the substantia nigra pars compacta region of the brain and post-mortem Lewy pathology, which leads to the loss of projecting axons innervating the striatum and to impaired motor and cognitive functions. While the cause of PD is still largely unknown, genome-wide association studies provide evidence that numerous polymorphic variants in various genes contribute to sporadic PD, and 10 to 15% of all cases are linked to some form of hereditary mutations, either autosomal dominant or recessive. Among the most common mutations observed in PD patients are in the genes LRRK2, SNCA, GBA1, PINK1, PRKN, and PARK7/DJ-1. In this review, we cover these PD-related mutations, the use of induced pluripotent stem cells as a disease in a dish model, and genetic animal models to better understand the diversity in the pathogenesis and long-term outcomes seen in PD patients.

1. Parkinson’s Disease Pathology, Symptoms, Treatments

Parkinson’s disease (PD) is a chronic and progressive neurodegenerative disorder that arises from the loss of dopaminergic neurons in the substantia nigra pars compacta (SNpc) of the midbrain and other organ systems [1,2]. It was first characterized by Dr. James Parkinson in 1817 in his “An essay on the Shaking Palsy”, where he described patients with trembling extremities, hunched posture, and a shuffling gait [3]. It is the second most prevalent neurodegenerative disease that affects 1% of the population older than 60 years of age [4,5]. The number of PD patients has doubled from 1990 to 2015 and is projected to double again by 2040 [6]. At the onset of the disease, 30–70% of midbrain dopaminergic neurons are lost [7] yet the disease progresses slowly over years with a prodromal stage that may last 10–15 years before the presentation of motor deficits and proper diagnosis [8,9].
The loss of dopaminergic neurons in the SNpc severs the nigrostriatal pathway, which is one of four major dopamine pathways in the brain. When the connection between the SNpc and the dorsal striatum is lost, the ability to produce controlled movement becomes impaired. The symptoms of the disease include uncontrollable resting tremors, postural imbalance, cogwheel rigidity, bradykinesia, akinesia, and cognitive impairment [10,11,12,13,14,15,16]. Depression, dementia, and hallucinations are also observed [17,18,19,20,21,22,23,24,25]. Available treatments only provide symptomatic therapy. Levodopa or dopamine agonists are used to replace lost dopamine. However, long-term use of levodopa may lead to adverse side effects, such as dyskinesia [26]. Monoamine oxidase B (MAO-B) and catechol-O-methyl transferase (COMT) inhibitors are provided in conjunction with levodopa to inactivate dopamine metabolism and degradation [27]. When the patient no longer responds positively to dopamine replacement treatment, a neurosurgical procedure known as deep brain stimulation can be performed to alleviate symptoms [28,29,30,31,32]. This procedure utilizes an electrode that is surgically implanted into the ventral intermediate nucleus of the thalamus, globus pallidus, and subthalamic nucleus to reduce severe motor complications [33,34]. To date, there is no curative treatment available for PD.
For the vast majority of patients, the cause of PD remains unknown, and it is primarily an idiopathic disease. Nevertheless, there are many environmental risk factors that are known contributors to PD, including metals such as iron, copper, manganese, lead, and mercury, as well as toxins such as 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), rotenone, paraquat, dieldrin, hexachlorohexanes, and 2-4-dichlorophenoxyacetic acid [35]. The use of amphetamines and methamphetamines also increases risk [36]. However, age remains to be the strongest risk factor for PD [37].
At least 5% of cases are linked to specific genetic mutations [38]. Mutations in genes, such as leucine-rich repeat kinase 2 (LRRK2), α-synuclein (SNCA), glucosylceramidase β 1 (GBA1), phosphatase and tensing homolog-induced kinase 1 (PINK1), parkin (PRKN)and PARK7 (also referred to as DJ-1), are associated with increased PD risk [39]. Much of what we know about these mutations was discovered by using in vivo transgenic animal models and in vitro stem cell models [40,41,42,43]. The consequences of these mutations lead to aberrations in oxidative stress, mitochondrial dysfunction, perturbed protein quality control, protein aggregation, and altered kinase activity. Some mutations may also lead to the early onset of PD.

2. Leucine-Rich-Repeat Kinase 2 (LRRK2)

The LRRK2 protein has Roc-COR-kinase domains and exhibits profound kinase and GTPase activity [44,45,46]. It exists as an oligomeric structure with minimal kinase activity. Once bound to guanosine triphosphate (GTP), it dissociates to form an intermediate structure whereupon it autophosphorylates to form a homodimer kinase [47]. It has functions linked to transcription, translation, autophagy, mitochondrial function, cytoskeletal remodeling vesicular transport, dopamine homeostasis, and synaptogenesis [48,49,50,51,52,53]. It has also been shown in human brains to be constitutively expressed in neurons and glia [54]. LRRK2 interacts with 14-3-3 proteins via phosphorylated Ser910 and Ser935 [55]. The binding of 14-3-3 to LRRK2 is thought to be disrupted with PD-related mutations [56]. Knockout of wild-type LRRK2 causes impairment of protein degradation pathways, accumulation of α-synuclein (α-syn), and apoptotic cell death in aged mice, signifying LRRK2 is essential in those pathways [57]. Furthermore, increased LRRK2 kinase activity was observed in post-mortem brain tissue and immune cells of idiopathic PD patients, providing clinical significance of LRRK2 as a therapeutic target [58,59,60].
LRRK2 is known to be phosphorylated at these sites: Ser910, Ser935, Ser955, Ser973, Ser1292, and Thr826 [61]. Specifically, the Ser1292 residue is the site of autophosphorylation [62]. Autophosphorylation is significantly increased in disease-causing variants [63]. The kinase activity is also required for the cytotoxicity of LRRK2 mutants [64,65]. Autophosphorylation was observed to have decreased when LRRK2 was chemically inhibited [55]. Pharmacological inhibition was reported to protect against the toxic phenotype of hyper-kinase activity [66]. However, LRRK2 kinase inhibition was shown to have off-target effects in peripheral tissues in animal models [67,68,69].
LRRK2 was reported to phosphorylate cytoskeletal proteins, such as tau, microtubule affinity regulating kinase 1 (MARK1), tubulins, rho guanine nucleotide exchange factor 7 (ARHGEF7), and ezrin-radixin-moesin (ERM) [70,71,72,73,74,75,76]. LRRK2 also exhibits physical interaction with F-actin and microtubules [66,77]. Pathogenic LRRK2 is believed to alter cytoskeletal scaffolding leading to deficits in neurite outgrowth and axonal transport [78,79]. In fact, phosphorylation of ezrin, radixin, and moesin (ERM) family proteins by pathogenic LRRK2 promoted the cytoskeletal rearrangement [76]. Rab10 is another phosphorylation target of LRRK2, specifically at Thr73 [80]. LRRK2 variants augmented Rab10 phosphorylation which affected vesicular transport [81]. Increased Rab10 phosphorylation was observed in dopaminergic neurons in the SNpc of idiopathic PD patients, which makes this an excellent and indirect measure for the LRRK2 kinase activity [82]. LRRK2 was also shown to control the vesicle trafficking [83,84]. Snapin, EndophilinA, and Rab5a were identified as possible LRRK2 substrates involved in synaptic vesicle exo- and endocytosis [85,86,87,88]. Evidence suggests that LRRK2 could inhibit autophagy, causing accumulation of autophagic organelles, that is reversed through kinase inhibition [89,90,91]. Hyperphosphorylation also affected 4E-BP1, leading to excessive protein translation and neurodegeneration in Drosophila model [50]. However, Drosophila only possesses LRRK1 and not LRRK2 and such findings were not strongly translated to mammalian models [92]. Although, LRRK2 phosphorylates ribosomal protein S15, regulating protein synthesis in the G2019S LRRK2 transgenic Drosophila and human dopaminergic neurons [93].

3. Common LRRK2 Mutations That Lead to Parkinson’s Disease

The most prevalent genetic mutation has been identified in the LRRK2 gene [94,95]. This particular mutation, known as the G2019S, accounts for at least 1% of all PD cases and 5–15% of familial PD cases [96]. Clinically, symptoms of LRRK2 mutant carriers are indistinguishable from idiopathic PD [97]. LRRK2 variants either increase or decrease the risk of developing PD. The common variants in Asia associated with increased risk of PD are A419V, R1628P, and G2385R [98,99,100]. These missense variants increase the risk for disease by about two-fold [101]. Interestingly, the variants N551K, R1398H, and K1423K appear to reduce the risk of PD [102,103]. There are eight isoforms of LRRK2 that lead to autosomal dominant familial PD: R1441C/H/G/S, Y1699C, N1437H, I2020T, and G2019S [104,105,106,107]. Each of these isoforms is a single amino acid substitution. The R1441C/H/G/S, N1437H, and Y1699C variants are located in the GTPase domain and reduce the GTP hydrolysis [108,109,110]. The variants in the GTPase and kinase domains are linked to neurotoxicity [52,63]. The I2020T and G2019S mutations are found on the kinase domain and lead to increased LRRK2 kinase activity [63,64,65,107,111,112]. Although the effect on kinase activity is not as pronounced in I2020T as it is in the G2019S variant [113], it has been well established that the mutations within the kinase domain augment kinase function and lead to Parkinsonism.

4. The G2019S LRRK2 Mutation

The most prevalent of all LRRK2 mutations is the G2019S point mutation located in the kinase domain of the protein [95,114]. This mutation exhibits high prevalence in Ashkenazi Jewish, North African Berber, and Arab PD patients [115,116,117]. Interestingly, it is not as common among Asian PD patients [39,118,119,120,121,122,123]. It is an autosomal dominant point mutation that resides in the kinase domain. This in turn leads to the hyper-kinase activity of LRRK2 exhibited in enhanced autophosphorylation at S1292, up to four-fold increase [63,64,65,124]. However, the mutation does not alter the gene expression of LRRK2. The symptoms that arise in disease-manifesting carriers closely resemble those associated with idiopathic PD [125,126]. Kinase inhibition blocks neurotoxicity in vitro and in vivo [66]. Interestingly, there is no marked phenotypic difference between heterozygous and homozygous G2019S mutant carriers [127]. Dopaminergic neurons derived from induced pluripotent stem cells (iPSC) harboring the G2019S mutation demonstrate increased susceptibility to oxidative stress and early neuritic branching defects [128]. G2019S LRRK2 is also linked to an upregulation of the p53-p21 pathway, contributing to cellular senescence and accumulation of α-syn [129]. However, the full impact of G2019S mutation is still largely unknown [97].

5. G2019S LRRK2 Induced Pluripotent Stem Cell Models

In vitro disease modeling has been revolutionized with the use of iPSCs [130,131,132,133,134]. To generate iPSCs, somatic cells are reprogrammed with the transfection of pluripotent factors octamer-binding protein 3/4 (Oct3/4), SRY-box transcription factor 2 (Sox2), c-Myc, and Krüppel-like factor 4 (Klf4) [135]. The reprogramming of adult somatic cells allows for the creation of patient-specific iPSCs that harbor known pathological genetic mutations [136]. The reprogrammed cells express embryonic stem cell markers and are capable of differentiating into any cell type of the three germ layers. Patient iPSCs can be genetically modified to correct mutations with CRISPR/Cas9, transcription activator-like effector nucleases (TALENS), or zinc-finger nuclease [137,138,139,140,141,142]. iPSCs can then be differentiated into a cell type of choice and sequenced to identify differentially expressed genes. Patient-derived iPSCs are a powerful tool to study a disease in a dish, and indeed many groups have used neurons derived from PD patient iPSCs to advance our understanding of the cause of neuronal degeneration.
The first LRRK2 iPSCs were generated from the PD patient fibroblasts of a 63-year-old male homozygous carrier and a 42-year-old male heterozygous for G2019S LRRK2 [143]. Accumulation of α-syn and neurite retraction have been observed in LRRK2 G2019S cells [128,144,145]. Endogenous α-syn accumulation also occurred when exogenous α-syn in the form of recombinant human preformed fibrils was cultured in iPSC-derived neurons [146]. Neural stem cells derived from G2019S iPSCs exhibited increased susceptibility to proteasomal stress and passage-dependent deficiencies in nuclear-envelope organization, clonal expansion, and neuronal differentiation [143]. Specifically, neuronal differentiation was affected as dopaminergic neurons showed early neuritic branching defects [147]. Sensory neurons derived from iPSCs also showed shortened neurites as well as microtubule-rich axon aggregation and altered calcium dynamics [148]. Furthermore, G2019S iPSC-derived dopaminergic neurons also displayed lower baseline ER-Ca2+ levels with Ca2+ influx increased, and Ca2+ buffering capacity decreased after membrane depolarization [149].
Single-cell RNA-sequencing (scRNA-seq) of G2019S PD patient iPSC-derived neural stem cells revealed affected genes involved in mitochondrial function, DNA repair, protein degradation, oxidative stress, lysosome biogenesis, ubiquitin–proteasome system, endosome function, autophagy, and mitochondrial quality control [150]. It was also recently shown in another scRNA-seq experiment that neuroepithelial stem cells (NESC) derived from G2019S iPSCs exhibited mitochondrial defects [151]. They found that the G2019S NESCs exhibited fragmented mitochondria, impaired mitochondria function, and impaired autophagosomal-lysosomal pathway, and the cells were more prone to release reactive oxygen species (ROS). They later expanded on their scRNA-seq data to show that G2019S LRRK2 NESCs initiated early cell-cycle exit and earlier neural differentiation than wild type (WT), leading to increased cell apoptosis [152]. This was in part due to the downregulation of the core regulatory circuit transcription factor Nuclear Receptor Subfamily 2 Group F Member 1 (NR2F1) and altered distal super-enhancer activity. Importantly, using a single-cell longitudinal imaging platform, early studies demonstrated that Nrf2 expression in neurons directly mitigates toxicity induced by α-synuclein and mutations in LRRK2 and that this effect is time-dependent [153]. Furthermore, mitochondrial genome damage and mitochondrial transport-related PD pathogenesis were also present [154,155]. It is also believed that G2019S LRRK2 disrupts the interaction between the mitochondrial transport protein, Miro, to PINK1 and Parkin, arresting the movement of damaged mitochondria along the cytoskeleton and delaying mitophagy [156]. Dopaminergic neurons derived from G2019S iPSCs show this abnormal mitochondrial trafficking and distribution as well [157]. This group also revealed that despite high levels of sirtuin, there was a reduction of sirtuin deacetylase activity, nicotinamide adenine dinucleotide (NAD+), and protein lysine deacetylase activity, leading to bioenergy deficits.
In addition to mitochondrial deficits, hyper-kinase activity also disrupts the endocytosis of synaptic vesicles in iPSC-derived ventral midbrain neurons [158]. Transcriptomics and proteomics revealed that clathrin-mediated endocytosis was disrupted [159]. Specifically, endothelial cytokines I–III, dynamin-1, and various Rab proteins were significantly downregulated. Evidence suggests that the synaptic defects are due to LRRK2 phosphorylation of auxilin (DNAJC6) which causes differential clathrin binding and disrupts endocytosis [160]. This impaired endocytosis led to the accumulation of oxidized dopamine and caused reduced glucocerebrosidase activity and increased α-syn accumulation.
As mentioned before, LRRK2 G2019S mutant neurons were found to exhibit increased accumulation and release of α-syn [161]. Mutant iPSC-derived midbrain dopaminergic neurons also showed higher basal levels of LC3 II. This aberrant autophagy was reversed with the inhibition of mitochondrial fission with fission dynamin-related protein 1 (DRP1) peptide inhibitor p110 [162]. Additionally, LRRK2 phosphorylation of leucyl-tRNA synthetase (LRS) was shown to reduce leucine binding and impair autophagy, leading to protein misfolding and endoplasmic reticulum stress [163].
LRRK2 further plays a role in the innate and acquired immunity of the peripheral and central nervous system [164]. Interestingly, LRRK2 is highly expressed in macrophages and microglia [164]. Evidence suggests that LRRK2 controls the secretion of inflammatory mediators [164,165,166]. LRRK2 GTPase function is also implicated in the inflammatory response [167,168,169]. Indeed, LRRK2 mutant microglia and astrocytes exhibit increased inflammatory cytokine and chemokine production [170,171,172,173,174]. Knockout or pharmacological inhibition of LRRK2 alleviates this inflammatory response [175,176,177,178,179]. Furthermore, LRRK2 mutations that lead to α-syn accumulation also impair nuclear factor kappa B (NF-κB) signaling in iPSC-derived neurons [180]. G2019S iPSC-derived astrocytes exhibited downregulation of matrix metalloproteinase 2 (mmp2) and transforming growth factor β1 (TGFβ1) [181]. Finally, LRRK2 may also play a role in hematopoiesis as G2019S LRRK2 iPSC-derived monocytes undergo accelerated production while CD14+CD16+ monocytes are reduced [182]. These mutant monocytes also exhibit migratory deficits.

6. G2019S LRRK2 Animal Models

LRRK2 animal models have provided insights into the regulation of protein translation, vesicle trafficking, neurite outgrowth, autophagy, and cytoskeletal dynamics [93,183,184]. Different models exhibit different clinical aberrations, such as degeneration of midbrain dopaminergic neurons, accumulation of α-syn, abnormal dopamine secretion, and behavioral deficits. LRRK2 overexpression in Drosophila led to the age-dependent loss of dopaminergic neurons and reduction of locomotor activity [185,186,187]. G2019S LRRK2 induced loss of photoreceptors and the impaired visual system was also observed [188]. The mutation also induced the mislocalization of tau in dendrites, causing degeneration [189]. Overexpression of G2019S LRRK2 in C. elegans also resulted in dopaminergic neuron degeneration, enhanced vulnerability to mitochondrial dysfunction, and inhibition of autophagy [190,191]. Transgenic mice expressing G2019S mutations exhibited motor deficits but with minimal evidence of neurodegeneration [192]. Only two groups have demonstrated the G2019S LRRK2 model of age-dependent loss of dopaminergic neurons in the SNpc [193,194]. Transgenic rats expressing the G2019S mutation exhibited oxidative stress in the striatum and SNpc with increased inducible nitric oxide synthase expression and abnormal morphology of SNpc dopaminergic neurons [195,196]. Unfortunately, genetic animal models display inconsistent phenotypes and do not fully replicate the human condition (neurodegeneration, Lewy body formation, and significant motor deficits), thus better models are needed [197].

7. Synuclein α (SNCA)

The SNCA gene encodes the presynaptic protein α-Synuclein (α-syn), which has been found to be localized in the nuclear membrane and synaptic vesicles [198]. The function of this protein is not well understood but current evidence suggests it participates in the axonal transport of synaptic vesicles by binding and transporting fatty acids [199,200,201]. It is also believed to participate in the differentiation and survival of dopaminergic neuron progenitor cells of mice and humans [202,203]. In fact, α-syn is expressed in the SNpc, especially in neurons containing neuromelanin, an insoluble granular pigment [204].
The misfolding of α-syn can lead to aberrant aggregation in the form of insoluble filaments and deposits in nerve cells [205]. These α-syn aggregates are also a major component of Lewy bodies, cellular inclusions in the neuronal cytoplasm that may lead to PD pathogenesis [206]. Lewy bodies impair neuronal communication and may even spread to healthy neurons [207,208]. They are also known to increase oxidative stress, disrupt axonal transport for neurotransmitter vesicles, and contribute to transcriptional dysregulation, protein sequestration, mitochondrial and synaptic dysfunction, and inhibition of the ubiquitin–proteasome system [207,208,209,210,211,212,213]. Accumulation of α-syn also affects the lysosomal clearance of protein aggregates, thus resulting in a vicious cycle perpetuating the toxic effects of α-syn aggregates [214]. It has been shown that the majority of α-syn in Lewy bodies in postmortem PD brain tissue appears as phosphorylated S129 α-syn [215,216]. Evidence suggests that S129 phosphorylation promotes α-syn aggregation and neurotoxicity [217]. Aggregated forms of α-syn are in fact more prone to S129 phosphorylation and accumulation during disease progression [218,219,220]. Polo-like kinases, casein kinases, and G protein-coupled receptor kinases have been shown to modulate the phosphorylation of α-syn at S129 [220,221,222,223,224,225]. Interestingly, numerous reports suggest that LRRK2 may be involved in phosphorylating S129 and α-syn aggregation [226,227]. Ultimately, it is still uncertain whether S129 α-syn is indeed neurotoxic [228].
Braak et al. hypothesized that neurodegeneration occurs in a predetermined sequence caused by an unknown pathogen in the gut or nasal cavity initiating sporadic PD [229,230]. This is associated with a specific α-syn spreading pattern, which may be why PD patients exhibit gastrointestinal and olfactory problems [231,232,233]. Indeed, Lewy body pathology has been confirmed in neurons of the olfactory tract and enteric nervous system [234,235,236,237]. Therefore, it is possible that α-syn propagates in a prion-like fashion [238]. In fact, the cell-to-cell transmission of α-syn has been observed after transgenic α-syn overexpression or exposure to preformed fibrils of α-syn and homogenates from postmortem PD patients [239,240,241,242,243,244]. Therefore, the pathological aggregation of misfolded α-syn is hypothesized to be critical in PD pathogenesis.
An increased risk of developing PD in humans can result due to an overexpression of the SNCA gene because of locus triplication [245]. Alternatively, SNCA missense mutations have also been shown to increase PD risk. Mutations in the SNCA gene were first identified as causing autosomal dominant PD in a large Italian family known as the Contursi kindred [246]. These patients carried the A53T point mutation and exhibited hallmark characteristics of PD including Lewy body pathology and positive response to L-dopa treatment. However, early onset and rapid disease progression were also observed. Other missense mutations include A30P [247,248,249], E46K [250], A53E [251,252], A53V [253], G51D [254,255], H50Q [256], and A18T and A29S [257]. Copy number variations have also been reported in PD patients [245]. Carriers of the A30P mutation were also associated with early onset, but a milder disease progression was observed compared to those with the A53T mutation [247]. The A53T and A30P mutations increase the likelihood of α-syn protein oligomerization instead of fibrillation [258], which is believed to accelerate α-syn aggregation [259]. Patients with A53T and E46K developed dementia consistent with Lewy body dementia [250,260] as the E46K mutation also promotes α-syn aggregation similar to A53T [261]. Interestingly, patients with the E46K mutation have also experienced visual hallucinations [250]. The A18T and A29S missense mutations were found in patients with sporadic PD [257].

8. SNCA Induced Pluripotent Stem Cell Models

Dopaminergic neurons derived from iPSCs of a PD patient with SNCA triplication have been studied. The level of α-syn protein in these dopaminergic neurons was twice the amount compared to those derived from normal iPSCs [262]. These neurons exhibited changes in growth, viability, cellular energy metabolism, and stress resistance when challenged with starvation or toxins [263], and increased oxidative stress [264]. SNCA triplication also led to a reduced capacity for iPSCs to differentiate into neurons, decreased neurite outgrowth, and lower neuronal activity compared to normal neurons [265]. The mRNA levels of nuclear receptor-related 1 protein (NURR1), G-protein-regulated inward-rectifier potassium channel 2 (GIRK-2), and tyrosine hydroxylase (TH) were also significantly reduced. Furthermore, lysosomal dysfunction has also been also induced by the α-syn accumulation [266]. Aggregates appeared to interact with ATP synthase and lead to premature mitochondrial permeability transition pore opening, making neurons more vulnerable to cell death [267]. Interestingly, SNCA triplication also affects non-neuronal cells, such as microglia. Microglia derived from SNCA triplication iPSCs exhibited impaired phagocytosis compared to isogenic controls [268].
An isogenic gene-corrected iPSC line for A53T was generated with zinc-finger nuclease-mediated genomic editing [140]. Gene correction reversed nitrosative stress and endoplasmic reticulum stress in iPSC-derived neurons [269]. A53T mutation increased apoptotic cell death in iPSC-derived midbrain dopaminergic neurons by increasing S-nitrosylation of MEF2C, affecting the transcriptional regulation of PGC1a, a master regulator of mitochondrial biogenesis [270]. A53T iPSC-derived neurons also displayed irregular protein aggregation, compromised neuritic outgrowth, contorted or fragmented axons with varicosities containing α-syn and Tau, and disrupted synaptic connectivity [271]. Interestingly, A53T midbrain dopaminergic neurons contained higher concentrations of α-syn monomers relative to tetramers when compared to isogenic controls, ultimately decreasing solubility α-syn [272]. Oligomeric α-syn has emerged as the key mediator for α-syn accumulation [273]. In fact, A53T and SNCA triplication iPSC-derived dopaminergic neurons exhibited increased α-syn oligomerization in a proximity ligation assay [274]. Increased sensitivity to mitochondrial toxins and nitrosative stress-induced neuronal loss were also observed in A53T dopaminergic neurons [275]. Furthermore, transcriptomic analysis of A53T and SNCA triplication iPSC-derived dopaminergic neurons revealed perturbations in the expression of genes likened to mitochondrial function, which was consistent with a reduction in mitochondrial respiration, impaired mitochondrial membrane potential, aberrant mitochondrial morphology, and decreased levels of phosphorylated DPR1 Ser616 [274]. They also observed increased endoplasmic reticulum stress and impaired cholesterol and lipid homeostasis. Single-cell transcriptomic analysis of A53T iPSC-derived dopaminergic neurons compared to an isogenic gene-corrected counterpart revealed perturbations in glycolysis, cholesterol metabolism, synaptic signaling, and ubiquitin–proteasomal degradation [276]. Apart from neuronal lineage cells, O4+ oligodendrocyte linage cells derived from A53T iPSCs also exhibited impaired maturation compared to controls [277].

9. SNCA Animal Models

Transgenic models with SNCA mutations have typically failed to display clear dopaminergic neurodegeneration or parkinsonian motor deficits [278]. However, there has been apparent α-syn aggregation and altered neuronal functions in these models. For example, transgenic mice expressing truncated α-syn show a reduced number of nigro-striatal neurons due to cell loss during early development [279]. However, the clinical relevance of these transgenic models remains questionable as the mutated α-syn protein is affecting early developmental stages rather than the later onset of neurodegeneration as seen in human patients [280].
SNCA overexpression models in rodents have been shown to affect the development and maintenance of dopaminergic neurons [281]. Overexpression has led to the formation of α-syn aggregates in the brain causing motor and olfactory deficits but not dopaminergic neurodegeneration [282]. E46K SNCA rats also display α-syn aggregation, altered metabolism of striatal dopamine, and increased oxidative stress but again not dopaminergic neurodegeneration [283].
Viral vectors such as adeno-associated virus (AAV) and lentivirus (LV) have also been used for transfecting rodents with SNCA. The AAV6 serotype generated an 80% loss of dopaminergic neurons and profound motor deficits [284]. This model also showed progressive neurodegeneration over a 2-to-4-month period. Using an LV vector led to α-syn aggregation but with no apparent loss of dopaminergic neurons nor behavioral changes [285]. Alternatively, AAV2/7-α-syn transduction in mouse SNpc produced dose-dependent dopaminergic neurodegeneration and motor deficits [286]. AAV2/2 was used to deliver WT and A53T α-syn into marmosets, leading to 30–60% nigral dopaminergic neurodegeneration and subsequent striatal dopamine depletion with mild motor deficits [287]. AAV1/2-A53T α-syn in macaques displayed 30% nigrostriatal dopaminergic neurodegeneration, 50% dopamine depletion, and 40% DAT reduction [288].

10. Glucocerebrosidase 1 (GBA1)

GBA1 encodes a lysosomal protein β-glucocerebrosidase (GCase). Mutations in this gene result in the accumulation of glycolipid substrates in lysosomes which disrupts lysosomal function and can lead to an autosomal recessive lysosomal storage disorder known as Gaucher’s disease (GD) [289]. Although a small minority of GBA1 mutation carriers develop PD [290,291], mutations in GBA1 increase the risk of developing PD [292,293]. In fact, GBA1 mutations are common genetic risk factors for PD, where 7–10% of patients with PD are carriers of a GBA1 mutation [291,294]. The L444P point mutation has been identified as being associated with PD and GD patients [291,295]. Furthermore, the L444P mutation appears to have a higher risk of developing PD compared to other GBA1 point mutations [296,297,298]. Interestingly, there is no difference in risk between patients with either homozygous or heterozygous mutations, although homozygous carriers tend to have PD onset 6 to 11 years earlier than heterozygous counterparts [290,294,298,299]. The clinical features of GBA1-associated PD are similar to those associated with idiopathic PD, including olfactory deficits and sleep disturbance but with earlier onset and accelerated autonomic, cognitive, and motor decline [298]. These patients also present with a more pronounced loss of nigrostriatal dopaminergic neurons and greater Lewy body pathology compared to those with idiopathic PD. GCase deficiency and lysosomal dysfunction as a result of GBA1 mutations are thought to be important pathogenic mechanisms for PD [300]. GBA1 mutations appear to impair α-syn degradation in the lysosome due to perturbed GCase activity [301,302]. Interestingly, GCase activity is also reduced in postmortem brain tissue of PD patients without GBA1 mutations [303,304]. It is possible that the GlcCer substrate accumulation may promote the pathogenic conversion of α-syn into its insoluble form [305,306]. Indeed, the GlcCer substrate also stabilizes a-syn oligomeric intermediates and induces rapid polymerization of fibrils [302].

11. GBA1 Induced Pluripotent Stem Cell Models

Interestingly, WT GCase activity is reduced in brain tissue and iPSC-derived neurons of idiopathic PD patients and other genetic forms of PD without GBA1 mutation [160,303,304,307,308]. It was noted that mitochondrial oxidative stress leads to the accumulation of oxidized dopamine, resulting in reduced GCase activity, lysosomal dysfunction, and α-syn accumulation [309]. The iPSCs were generated from patients with GD and PD harboring GBA1 mutations and differentiated into midbrain DA neurons. The iPSC-derived neurons exhibited reduced GCase activity and protein levels, increased glucosylceramide and α-syn levels as well as autophagic and lysosomal defects [310]. The mutant neurons also exhibited dysregulation of calcium homeostasis and increased vulnerability to stress responses involving the elevation of cystolic calcium. Gene correction of the mutation rescues the observed pathological phenotypes [310]. N370S iPSC-derived dopaminergic neurons exhibited disruption of the autophagy pathway and ER stress leading to elevated extracellular a-syn [311]. GBA-PD patient-derived dopaminergic neurons with heterozygous N370S and L444P mutations display stress responses, sphingolipid accumulation, mitochondrial dysfunction, increased mitochondrial ROS, and changes in NAD+ metabolism, ameliorated with NAD+ precursor nicotinamide riboside [312]. Heterozygous-null GBA1 iPSC-derived cortical neurons and astrocytes exhibit reduced lysosome number, increased lysosomal pH, reduced lysosomal cathepsin protease activity, and increased accumulation of soluble and insoluble α-syn without changes in α-syn mRNA levels [313,314].
GCase chaperones were able to recover GCase activity and reduce α-syn levels in iPSC-derived DA neurons and mouse models [308]. GBA1 iPSC-derived neurons exhibited prolonged mitochondria–lysosome contacts due to defective modulation of the untethering protein TBC1D15, which mediates Rab7 GTP hydrolysis for contact untethering, ultimately leading to disrupted mitochondrial distribution and function [315]. This defect was rescued with a GCase modulator, indicating deficits were due to a lack of GCase activity. Another GCase chaperone S-181 was tested on 84GG GBA1 patient-derived dopaminergic neurons partially restored lysosomal function and lowered accumulation of oxidized dopamine, glucosylceramide, and α-syn [307].
Interestingly, G2019S LRRK2 iPSC-derived dopaminergic neurons also exhibited reduced GCase activity [316]. In fact, pharmacological inhibition of LRRK2 kinase activity increased GCase activity in both iPSC-derived dopaminergic neurons carrying LRRK2 and GBA1 mutations. The increase in GCase activity was sufficient to partially rescue the accumulation of oxidized dopamine and α-syn. Heterozygous-null GBA-1 iPSC-derived cortical neurons did not exhibit any differences in WT LRRK2 kinase activity [314]. However, LRRK2 inhibition rescued lysosomal number and Cathepsin L activity, and partial lysosome re-acidification. Interestingly, they did not see any change in GCase activity with LRRK2 inhibition, contradicting previous findings [316]. The possible interplay between LRRK2 and GBA1 is fascinating and needs to be further explored.

12. GBA1 Animal Models

There are over 200 known GD-associated mutations; therefore, it is difficult to determine which mutations are particularly responsible for PD susceptibility [295]. Homozygous L444P mice generated partial gene duplication and either died soon after birth due to compromised epidermal permeability barrier caused by defective glucosylceramide metabolism or exhibited systemic inflammation [317,318]. L444P conditional knock-in mice lived longer and exhibited increased striatal α-syn levels and astrogliosis at 1 year of age, although motor performance was not assessed [319]. Heterozygous L444P mice demonstrated impaired neuronal autophagy and mitophagy as well as mitochondrial dysfunction [320,321]. Heterozygous L444P mice also demonstrated impaired α-syn degradation and increased α-syn levels but interestingly did not form α-syn aggregates [320,321,322]. Furthermore, heterozygous L444P mice did not exhibit nigrostriatal neurodegeneration, neuroinflammation, or impairments in olfaction, coordination, and cognition [321,322,323]. The D409H mutation is a rare but severe mutation found in GBA1 PD patients [324]. α-Syn aggregation was observed in the cerebellum and brainstem of homozygous D409H mice [325]. However, heterozygous and homozygous D409H mice did not exhibit nigrostriatal neurodegeneration, neuroinflammation, or motor deficits [326]. Treatment with a GCase chaperone resulted in the activation of WT GCase and reduction of GCase lipid substrates and α-syn in brain tissue [307]. Homozygous V394L mice have 27% of WT GCase activity but do not exhibit α-syn aggregation [325]. Homozygous R643C mice exhibited increased nigrostriatal α-syn and neuroinflammation [319]. Homozygous N370S mutation has been neonatal lethal despite mild phenotype in patients [327]. Furthermore, several GBA1-PD models combine GBA1 mutation with overexpression of α-syn mutations to induce α-syn aggregation and pronounced PD symptoms despite the fact that these two mutations are not reported in PD patients [323,326,328].

13. PTEN-Induced Kinase 1 (PINK1)

Phosphatase and tensin homolog (PTEN)-induced putative kinase 1 (PINK1) is a mitochondria-targeted Ser/Thr protein kinase. In normal physiological conditions, PINK1 is imported into the mitochondria via the translocase of the outer membrane and translocase of the inner membrane [329]. PINK1 is cleaved in the transmembrane segment by the mitochondrial intramembrane protease PARL, where it is then retranslocated to the cytosol for proteasomal degradation [330]. Once the mitochondrial inner membrane becomes depolarized, PINK1 mitochondrial transport is arrested and phosphorylates Parkin on the ubiquitin-like domain and activates the E3 ligase activity of Parkin [331,332]. Parkin ubiquitinates mitochondrial outer membrane proteins and induces autophagic clearance of depolarized mitochondria [333,334,335]. PINK1 is localized on the mitochondria and may exhibit a protective effect, but as a result of a mutation that protection is lost, resulting in increased susceptibility to cellular stress [336]. It is also known to phosphorylate TNF receptor-associated protein 1 (TRAP1), 5-hydroxytryptamine receptor 2a (5-HT2A), and Parkin [337,338,339]. PINK1 mutations are the second most common autosomal-recessive form of early onset PD [329,340,341]. Furthermore, Pink1 mutation carriers may develop psychiatric comorbidity alongside gait disturbances [342,343].

14. PINK1 Induced Pluripotent Stem Cell Models

PINK1 kinase activity was significantly reduced in G411S PINK1 mutant iPSC-derived neurons [344]. In iPSC-derived neurons, PINK1 mutations reduced complex I activity, which lead to a reduction in mitochondrial membrane potential [345]. PINK1 PD patient iPSC-derived neurons treated with valinomycin, which triggers rapid loss of mitochondrial membrane potential, impaired recruitment of overexpressed Parkin to mitochondria [346,347]. Interestingly, loss of PINK1 abolished the degradation of mitochondrial protein only in fibroblasts, but not in isogenic iPSC-derived neurons, which do not exhibit significant mitophagy even with parkin overexpression and valinomycin treatment [346]. Phosphorylation of Ser250 in NdufA10 regulates the activity of ubiquinone reductase in mitochondrial complex I. A phosphomimetic mutant of NdufA10 reversed the deficiencies in complex I activity and ATP synthesis in PINK1 iPSC-derived neurons [345]. Furthermore, direct supplementation of cardiolipin, a mitochondrial inner membrane-specific lipid, to isolated mitochondria rescues PINK1-induced complex I defects [348]. Additionally, antioxidant treatment with coenzyme Q10 can rescue the cellular vulnerability associated with mitochondrial dysfunction in iPSC-derived neurons from PINK1 PD patients [349]. Apart from abnormal mitochondrial morphology, PINK1 PD patient iPSC-derived midbrain dopaminergic neurons exhibited α-syn accumulation and increased cytosolic dopamine levels [350]. Interestingly, increased expression of LRRK2 mRNA and protein was observed in PINK1 iPSC-derived neurons [351]. Moreover, transient overexpression of WT PINK1 can downregulate LRRK2 expression. The implication of this study suggests a convergent pathway between these two genes in PD pathogenesis.

15. PINK1 Animal Models

PINK1 knockout mice appear to have an age-dependent and moderate reduction in striatal dopamine levels accompanied by low locomotor activity due to deficient mitochondrial respiration and increased sensitivity to oxidative stress [352,353]. This model does not exhibit significant neurodegeneration, Lewy body formation, or loss of dopamine. Another PINK1 knockout mouse model has been reported to show olfactory and gait disturbances, similar to prodromal symptoms of human PD patients [354]. Pink1 deficient mice also have impaired dopamine release [355] and exhibited impaired complex I function, mitochondrial depolarization reduced ATP synthesis, and increased sensitivity to apoptotic stress [356]. As such, PINK1 knockout mice may be useful as a model for prodromal PD [357,358]. On the other hand, PINK1 null mice with an exon 4–5 deletion showed progressive loss of striatal dopamine, but nigrostriatal neurodegeneration was not observed [359].
The phenotype for rats differs from mice, with a closer resemblance to PD pathology. PINK1 knockout rats exhibit age-dependent loss of nigral dopaminergic neurons beginning at age 6–8 months [360,361]. Motor deficits including reduced rearing frequency and distance traveled in an open field, reduced hind limb grip strength, and increased foot slips and traversal time on a tapered balance beam were apparent [360,362]. These rats also exhibited mitochondrial respiration deficits and α-syn aggregation, although different from Lewy body pathology [361,362].

16. Parkin (PRKN)

Parkin is a RING domain-containing E3 ubiquitin ligase important for mitochondria quality control through mitophagy [363]. Autosomal recessive juvenile parkinsonism was first identified in Japanese patients with early onset PD [364]. Currently, there are more than 100 known mutations identified in Parkin [365,366,367,368]. Parkin mutations account for 50% of familial and 15% of sporadic cases of European PD patients with onset before 45 years of age [369,370]. Parkin mutations are also the most common form of juvenile PD or early onset PD [365,366,369,371,372,373]. Carriers of Parkin mutations exhibit earlier and more symmetrical onset, slower disease progression, and greater response to L-dopa, but seem to develop dyskinesia earlier as well [369,374,375]. Furthermore, cognitive impairment is rare, and dementia or depression are not present in Parkin mutation patients [369,376,377]. Most PD patients with Parkin mutations do not develop Lewy body pathology [378]. Furthermore, it appears that missense mutations are equally as detrimental to truncation and deletion mutations [379].

17. PRKN-Induced Pluripotent Stem Cell Models

Parkin patient iPSC-derived neurons present abnormal or enlarged mitochondria accompanied by increased oxidative stress and enhanced activity of nuclear factor erythroid 2-related factor 2 (Nrf2) pathway [380]. Proteomics analysis of Parkin knockout iPSC-derived neurons revealed disturbances in oxidative stress defense, mitochondrial respiration and morphology, cell cycle control, and cell viability [381]. Structural and functional analysis verified an increase in mitochondrial area and the presence of elongated mitochondria as well as impaired glycolysis and lactate-supported respiration. This abnormal mitochondria phenotype, such as elongated shape and larger volume, has been observed by other groups as well [350,380,382]. However, two different studies using qPCR showed that there was no significant difference in mitochondria DNA copy number [380,383]. Additionally, it seems Parkin mutations in iPSC-derived neurons may alter mitochondrial morphology in a portion of cells, particularly TH+ dopaminergic neurons [382]. In fact, Parkin patient iPSC-derived dopaminergic neurons exhibited smaller and less functional mitochondria than those in non-dopaminergic neurons [384] and exhibited increased cytotoxicity with known PD environmental risk factors, such as exposure to heavy metals such as copper and cadmium [385]. They also showed a significant increase in mitochondrial fragmentation, initial ROS generation, and loss of mitochondrial membrane potential following copper exposure. Transfected Parkin cell lines have shown that Parkin is recruited to the mitochondria to ubiquitinate a variety of substrates for the induction of mitophagy [363,386]. However, the recruitment of endogenous Parkin to mitochondria has not been robustly seen in cell lines [387,388], mice [389,390], or iPSC-derived human neurons [346,383].
The precision of dopaminergic transmission is significantly disrupted by increased spontaneous dopamine release and decreased reuptake [383]. Dopamine induces oscillatory neuronal activities in Parkin iPSC-derived neurons but not in normal iPSC-derived neurons [391]. Interestingly, this phenotype mirrors the widespread rhythmic bursting of neuronal activities in the basal ganglia of PD patients [392]. Parkin iPSC-derived midbrain dopaminergic neurons exhibited reduced length and complexity of neuronal processes understood to be caused by a marked decrease in microtubule stability, as the phenotype was rescued by taxol, a microtubule-stabilizing drug, or the overexpression of Parkin [393]. This decreased neurite length and complexity was also mimicked in normal neurons treated with colchicine, a microtubule-depolymerizing agent. Reduced microtubule stability in iPSC-derived neurons was observed by other groups as well [394].
In iPSC-derived neurons from two Parkin patients, increased accumulation of α-syn was observed in the patient with Lewy body pathology, but not in the other patient who did not have Lewy body pathology [380]. There was no significant difference in α-syn protein levels in iPSC-derived midbrain dopaminergic neurons from two Parkin PD patients and two normal subjects [383]. Although, an increased level of α-syn protein was observed in two other studies using iPSC-derived neurons from PD patients with Parkin mutations [350,382]. The inconsistency of α-syn accumulation in Parkin PD patient-derived cells may suggest that α-syn accumulation is independent of Parkin as Parkin PD patients do not necessarily develop Lewy body pathology.

18. PRKN Animal Models

Transgenic rodent models with Parkin overexpression exhibited protection from 6-OHDA or α-syn overexpression in the SNpc [395,396,397,398]. Parkin knockout mice have nigrostriatal mitochondrial respiration deficits and increased markers for oxidative stress but did not exhibit any significant nigrostriatal dopaminergic neurodegeneration nor PD-like locomotor deficits [360,399,400,401,402,403,404,405]. Although there were instances of Parkin knockout mice exhibiting slight impairment of dopamine release [406,407], they did not show significant behavioral deficits or age-dependent nigral dopaminergic neurodegeneration [360]. They also lacked Lewy body pathology similarly to human parkin mutant carriers that rarely showed LB pathology [408]. Successful transgenic rodent models that recapitulated dopaminergic neurodegeneration were eventually developed. Indeed, Parkin-Q311X-DAT-BAC mice expressing a C-terminal truncated human mutant Parkin in dopaminergic neurons exhibited late onset and progressive hypokinetic motor deficits, age-dependent nigrostriatal dopaminergic neurodegeneration, α-syn accumulation, and a significant reduction in striatal dopamine neuron terminals and dopamine levels [409]. Overexpression of both T240R Parkin and human WT Parkin in rats with AAV2/8 induced progressive and dose-dependent dopaminergic neurodegeneration [410].

19. Parkinsonism Associated Deglycase (DJ-1 Also Known as PARK7)

The DJ-1 protein is 189 amino acid residues with three cysteines and normally forms a homodimer that exhibits antioxidant activities [411]. DJ-1 is highly expressed in astrocytes in the frontal cortex and SNpc of idiopathic PD brains and is not an essential component of Lewy bodies [412,413]. It is also involved in the regulation of apoptosis and pro-survival signaling, autophagy, inflammatory responses, and protection against oxidative stress [414,415]. In fact, DJ-1 overexpression protects against oxidative stress while DJ-1 knockout increases oxidative stress-induced cell death [416,417,418,419,420,421]. It seems that the cysteine residue at position 106 is required for DJ-1 mediated protection from oxidative stress [411,422,423,424]. DJ-1 levels increase in response to oxidative stress caused by dopamine to suppress ROS accumulation [425]. DJ-1 also functions as a redox-sensitive chaperone that is activated in an oxidative cytoplasmic environment and can inhibit the generation of α-syn aggregates [426,427]. It can directly interact with α-syn monomers and oligomers, where mutant DJ-1 exhibits α-syn dimerization [428]. Furthermore, DJ-1 deficiency decreases LAMP2A expression, a receptor required for CMA-mediated α-syn degradation, thus leading to α-syn accumulation [429]. It can also act as a glyoxalase III to detoxify reactive dicarbonyl species, such as glyoxal and methylglyoxal into glycolic or lactic acid in the absence of glutathione [430].
A deletion mutation in a Dutch family and a homozygous point mutation L166P in an Italian family were identified to cause parkinsonism [431]. DJ-1 mutations only account for less than 1% of all early onset PD cases [432]. The median age of onset for DJ-1 PD is 27 years [433]. PD patients carrying DJ-1 mutations exhibit early onset of dyskinesia, rigidity, and tremors and respond well to L-DOPA treatment [431,434,435]. Cognitive deficits and psychotic disturbances typically arise later in disease progression [436]. The L166P mutation abolishes DJ-1 dimerization, abrogating its neuroprotective activity [437]. Post-mortem brain tissue of a patient with L172Q DJ-1 mutation exhibited Lewy body pathology [438]. The clinical manifestations of DJ-1 patients are like those with Parkin and PINK1 mutations. However, compared to Parkin and PINK1, DJ-1 mutation carriers exhibit a higher percentage of non-motor symptoms such as anxiety, cognitive decline, depression, and psychopathic symptoms [433,439,440]. It has been reported that PD brains exhibit decreased levels of DJ-1 mRNA and protein, but also show the presence of extra-oxidized DJ-1 isoforms [441] and that acidic isoforms of DJ-1 monomer were accumulated in sporadic PD brains [442]. It remains unclear how DJ-1 contributes to PD pathogenesis.

20. DJ-1 Induced Pluripotent Stem Cell Models

Homozygous DJ-1 mutant iPSC-derived midbrain dopaminergic neurons exhibited increased dopamine oxidation and neuromelanin-like pigmented aggregates compared to an isogenic gene-corrected control [309]. Oxidative stress from dopamine metabolism triggered mitochondrial oxidative stress, which was significantly attenuated by blocking dopamine synthesis [309]. Dopamine-induced oxidative stress inactivated GCase inhibited lysosomal function and led to increased expression of α-syn [309]. DJ-1 knockout iPSC-derived dopaminergic neurons also exhibited enhanced α-syn fibril-induced aggregation and neuronal death [443]. Other DJ-1 patient-specific iPSCs have been generated very recently, but phenotypic analysis has yet to be performed [444,445,446].

21. DJ-1 Animal Models

DJ-1 knockout mice show mild motor deficits and altered nigrostriatal synaptic physiology, but without dopaminergic neuron loss or significant change in dopamine levels [417,447,448]. However, one report showed that the DJ-1 knockout mouse line exhibited a loss of dopaminergic neurons in the ventral tegmental area and slight behavioral changes, such as diminished rearing behavior and impaired object recognition [449]. It was also observed that DJ-1 deficient mice exhibit alteration in dopamine metabolism, specifically an increase in dopamine reuptake causing an accumulation of striatal dopamine [450]. The higher overall level of oxidized dopamine did not result in neurodegeneration as seen in PD patients and this may be due to the fact that human SNpc dopaminergic neurons have higher overall dopamine levels than those in mice. A DJ-1 nullizygous mouse was fully backcrossed with a C57BL/6 background and displayed dramatic early onset unilateral loss of nigrostriatal dopaminergic neurons that progressed bilaterally with aging [451]. These mice also exhibited age-dependent bilateral degeneration in the locus ceruleus nucleus and displayed mild motor deficits. DJ-1 knockout mice also exhibited more microglial activation, especially in response to lipopolysaccharide insult [452]. On the other hand, DJ-1 knockout rats showed significant age-dependent nigrostriatal dopaminergic neuron loss (~50%) between 6 and 8 months of age, accompanied by motor deficits [360]. Furthermore, mitochondria from DJ-1 knockout rats showed altered respiration compared to that of WT rats [453].

22. Overall Discussion

As PD is primarily an idiopathic disease, the major challenge is in understanding the interplay between genetic and environmental factors that influence susceptibility to PD. In the last two decades, research into PD genetics has deepened our understanding of PD risk, onset, progression, and therapeutic approaches. These studies have revealed pathogenic pathways of neurodegeneration shared between inherited and sporadic PD. No animal PD model is perfect as it is challenging to recapitulate the age of onset, the timing of disease progression, and the spectrum of pathologies present in PD patients. Single model modalities, such as neurotoxin models recapitulate neurodegeneration and PD-like symptoms and are appropriate for addressing specific questions tailored to this model, as is the case for the genetic models. Current research has been utilizing iPSCs for a disease-in-a-dish approach. Isogenic gene-corrected cell lines offer unequivocal perturbations caused by genetic mutations. Since iPSC-based in vitro models are patient-derived they present many advantages including the ability to generate a variety of neural lineages in 3D complex systems that serve to model the in vivo brain tissue cytoarchitecture for studying pathogenesis. PD is not a single disease entity, but rather made up of subtypes based on differences in the spectrum of symptoms and the nature and distribution of Lewy body pathologies.Thus, it is necessary to combine iPSC-based in vitro models, ex vivo post-mortem brain specimens, and in vivo models to further our understanding of the different cellular and molecular mechanisms underlying PD pathogenesis.

Author Contributions

J.K., E.W.D., T.O., E.S.D. and M.M.D. provided summaries of literature searches and presentations, performed research related to the topic; provided comments on the manuscript, J.K. and M.M.D. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

National Institute on Aging R56 AG059284 and the NIH Primate Center Base grant (Office of Research Infrastructure Programs/OD P51 OD011133).

Acknowledgments

This work was also supported by the Worth Family Fund, The Perry and Ruby Stevens Charitable Foundation, The Robert J. Jr. and Helen C. Kleberg Foundation, The Marmion Family Fund, The William and Ella Owens Medical Research Foundation, the National Institute on Aging R56 AG059284 and the NIH Primate Center Base grant (Office of Research Infrastructure Programs/OD P51 OD011133).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Braak, H.; Braak, E.; Yilmazer, D.; Schultz, C.; de Vos, R.A.; Jansen, E.N. Nigral and extranigral pathology in Parkinson’s disease. J. Neural Transm. Suppl. 1995, 46, 15–31. [Google Scholar] [PubMed]
  2. Jellinger, K.A. Pathology of Parkinson’s disease. Changes other than the nigrostriatal pathway. Mol. Chem. Neuropathol. 1991, 14, 153–197. [Google Scholar] [PubMed]
  3. Parkinson, T. Outlines of Zoonosological Tables. Lond. Med. Phys. J. 1817, 38, 449–453. [Google Scholar] [PubMed]
  4. Hirsch, L.; Jette, N.; Frolkis, A.; Steeves, T.; Pringsheim, T. The Incidence of Parkinson’s Disease: A Systematic Review and Meta-Analysis. Neuroepidemiology 2016, 46, 292–300. [Google Scholar] [CrossRef] [PubMed]
  5. Shalash, A.S.; Hamid, E.; Elrassas, H.H.; Bedair, A.S.; Abushouk, A.I.; Khamis, M.; Hashim, M.; Ahmed, N.S.; Ashour, S.; Elbalkimy, M. Non-Motor Symptoms as Predictors of Quality of Life in Egyptian Patients With Parkinson’s Disease: A Cross-Sectional Study Using a Culturally Adapted 39-Item Parkinson’s Disease Questionnaire. Front. Neurol. 2018, 9, 357. [Google Scholar] [CrossRef]
  6. Dorsey, E.R.; Sherer, T.; Okun, M.S.; Bloem, B.R. The Emerging Evidence of the Parkinson Pandemic. J. Park. Dis. 2018, 8, S3–S8. [Google Scholar] [CrossRef] [Green Version]
  7. Riederer, P.; Wuketich, S. Time course of nigrostriatal degeneration in parkinson’s disease. A detailed study of influential factors in human brain amine analysis. J. Neural Transm. 1976, 38, 277–301. [Google Scholar] [CrossRef]
  8. Schaeffer, E.; Postuma, R.B.; Berg, D. Prodromal PD: A new nosological entity. Prog. Brain Res. 2020, 252, 331–356. [Google Scholar] [CrossRef]
  9. Poewe, W.; Seppi, K.; Tanner, C.M.; Halliday, G.M.; Brundin, P.; Volkmann, J.; Schrag, A.E.; Lang, A.E. Parkinson disease. Nat. Rev. Dis. Primers. 2017, 3, 17013. [Google Scholar] [CrossRef]
  10. Hughes, A.J.; Daniel, S.E.; Kilford, L.; Lees, A.J. Accuracy of clinical diagnosis of idiopathic Parkinson’s disease: A clinico-pathological study of 100 cases. J. Neurol. Neurosurg. Psychiatry 1992, 55, 181–184. [Google Scholar] [CrossRef]
  11. Jankovic, J. Parkinson’s disease: Clinical features and diagnosis. J. Neurol. Neurosurg. Psychiatry 2008, 79, 368–376. [Google Scholar] [CrossRef] [Green Version]
  12. Gelb, D.J.; Oliver, E.; Gilman, S. Diagnostic criteria for Parkinson Disease. Arch. Neurol. 1999, 56, 7. [Google Scholar] [CrossRef]
  13. Reichmann, H. Clinical criteria for the diagnosis of Parkinson’s disease. Neurodegener Dis. 2010, 7, 284–290. [Google Scholar] [CrossRef] [Green Version]
  14. Virmani, T.; Moskowitz, C.B.; Vonsattel, J.P.; Fahn, S. Clinicopathological characteristics of freezing of gait in autopsy-confirmed Parkinson’s disease. Mov. Disord. 2015, 30, 1874–1884. [Google Scholar] [CrossRef]
  15. Tolosa, E.; Compta, Y. Dystonia in Parkinson’s disease. J. Neurol. 2006, 253 (Suppl. S7), VII7–VII13. [Google Scholar] [CrossRef]
  16. Doherty, K.M.; van de Warrenburg, B.P.; Peralta, M.C.; Silveira-Moriyama, L.; Azulay, J.P.; Gershanik, O.S.; Bloem, B.R. Postural deformities in Parkinson’s disease. Lancet Neurol. 2011, 10, 538–549. [Google Scholar] [CrossRef]
  17. Emre, M.; Aarsland, D.; Brown, R.; Burn, D.J.; Duyckaerts, C.; Mizuno, Y.; Broe, G.A.; Cummings, J.; Dickson, D.W.; Gauthier, S.; et al. Clinical diagnostic criteria for dementia associated with Parkinson’s disease. Mov. Disord. 2007, 22, 1689–1707. [Google Scholar] [CrossRef]
  18. Remy, P.; Doder, M.; Lees, A.; Turjanski, N.; Brooks, D. Depression in Parkinson’s disease: Loss of dopamine and noradrenaline innervation in the limbic system. Brain 2005, 128, 1314–1322. [Google Scholar] [CrossRef] [Green Version]
  19. Onofrj, M.; Thomas, A.; Bonanni, L. New approaches to understanding hallucinations in Parkinson’s disease: Phenomenology and possible origins. Expert. Rev. Neurother. 2007, 7, 1731–1750. [Google Scholar] [CrossRef]
  20. Williams-Gray, C.H.; Foltynie, T.; Lewis, S.J.; Barker, R.A. Cognitive deficits and psychosis in Parkinson’s disease: A review of pathophysiology and therapeutic options. CNS Drugs 2006, 20, 477–505. [Google Scholar] [CrossRef]
  21. Williams-Gray, C.H.; Foltynie, T.; Brayne, C.E.; Robbins, T.W.; Barker, R.A. Evolution of cognitive dysfunction in an incident Parkinson’s disease cohort. Brain 2007, 130, 1787–1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. McAuley, J.H.; Gregory, S. Prevalence and clinical course of olfactory hallucinations in idiopathic Parkinson’s disease. J. Park. Dis. 2012, 2, 199–205. [Google Scholar] [CrossRef] [PubMed]
  23. Inzelberg, R.; Kipervasser, S.; Korczyn, A.D. Auditory hallucinations in Parkinson’s disease. J. Neurol. Neurosurg Psychiatry 1998, 64, 533–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Fenelon, G.; Thobois, S.; Bonnet, A.M.; Broussolle, E.; Tison, F. Tactile hallucinations in Parkinson’s disease. J. Neurol. 2002, 249, 1699–1703. [Google Scholar] [CrossRef] [PubMed]
  25. Reijnders, J.S.; Ehrt, U.; Weber, W.E.; Aarsland, D.; Leentjens, A.F. A systematic review of prevalence studies of depression in Parkinson’s disease. Mov. Disord. 2008, 23, 183–189. [Google Scholar] [CrossRef] [Green Version]
  26. Marsden, C.D.; Parkes, J.D. Success and Problems of Long-Term Levodopa Therapy in Parkinson’S Disease. Lancet 1977, 309, 345–349. [Google Scholar] [CrossRef]
  27. Thanvi, B.R.; Lo, T.C. Long term motor complications of levodopa: Clinical features, mechanisms, and management strategies. Postgrad. Med. J. 2004, 80, 452–458. [Google Scholar] [CrossRef] [Green Version]
  28. Deuschl, G.; Schade-Brittinger, C.; Krack, P.; Volkmann, J.; Schäfer, H.; Bötzel, K.; Daniels, C.; Deutschländer, A.; Gruber, D.; Hamel, W.; et al. A Randomized Trial of Deep-Brain Stimulation for Parkinson’s Disease. N. Engl. J. Med. 2006, 355, 896–908. [Google Scholar] [CrossRef] [Green Version]
  29. Krack, P.; Pollak, P.; Limousin, P.; Hoffmann, D.; Benazzouz, A.; Benabid, A.L. Inhibition of levodopa effects by internal pallidal stimulation. Mov. Disord. 1998, 13, 648–652. [Google Scholar] [CrossRef]
  30. Kumar, R.; Lozano, A.M.; Kim, Y.J.; Hutchison, W.D.; Sime, E.; Halket, E.; Lang, A.E. Double-blind evaluation of subthalamic nucleus deep brain stimulation in advanced Parkinson’s disease. Neurology 1998, 51, 850–855. [Google Scholar] [CrossRef]
  31. Limousin, P.; Krack, P.; Pollak, P.; Benazzouz, A.; Ardouin, C.; Hoffmann, D.; Benabid, A.L. Electrical stimulation of the subthalamic nucleus in advanced Parkinson’s disease. N. Engl. J. Med. 1998, 339, 1105–1111. [Google Scholar] [CrossRef]
  32. Wichmann, T.; DeLong, M.R. Deep Brain Stimulation for Movement Disorders of Basal Ganglia Origin: Restoring Function or Functionality? Neurotherapeutics 2016, 13, 264–283. [Google Scholar] [CrossRef] [Green Version]
  33. Linazasoro, G.; Van Blercom, N.; Lasa, A. Unilateral subthalamic deep brain stimulation in advanced Parkinson’s disease. Mov. Disord. 2003, 18, 713–716. [Google Scholar] [CrossRef]
  34. Diamond, A.; Jankovic, J. The effect of deep brain stimulation on quality of life in movement disorders. J. Neurol. Neurosurg Psychiatry 2005, 76, 1188–1193. [Google Scholar] [CrossRef] [Green Version]
  35. Goldman, S.M. Environmental toxins and Parkinson’s disease. Annu. Rev. Pharmacol. Toxicol. 2014, 54, 141–164. [Google Scholar] [CrossRef]
  36. Callaghan, R.C.; Cunningham, J.K.; Sykes, J.; Kish, S.J. Increased risk of Parkinson’s disease in individuals hospitalized with conditions related to the use of methamphetamine or other amphetamine-type drugs. Drug Alcohol Depend. 2012, 120, 35–40. [Google Scholar] [CrossRef]
  37. Antony, P.M.; Diederich, N.J.; Kruger, R.; Balling, R. The hallmarks of Parkinson’s disease. FEBS J. 2013, 280, 5981–5993. [Google Scholar] [CrossRef] [Green Version]
  38. Verstraeten, A.; Theuns, J.; Van Broeckhoven, C. Progress in unraveling the genetic etiology of Parkinson disease in a genomic era. Trends Genet. 2015, 31, 140–149. [Google Scholar] [CrossRef]
  39. Kalinderi, K.; Bostantjopoulou, S.; Fidani, L. The genetic background of Parkinson’s disease: Current progress and future prospects. Acta Neurol. Scand. 2016, 134, 314–326. [Google Scholar] [CrossRef]
  40. Cenci, M.A.; Bjorklund, A. Animal models for preclinical Parkinson’s research: An update and critical appraisal. Prog. Brain Res. 2020, 252, 27–59. [Google Scholar] [CrossRef]
  41. Bose, A.; Petsko, G.A.; Studer, L. Induced pluripotent stem cells: A tool for modeling Parkinson’s disease. Trends Neurosci. 2022, 45, 608–620. [Google Scholar] [CrossRef]
  42. Antonov, S.A.; Novosadova, E.V. Current State-of-the-Art and Unresolved Problems in Using Human Induced Pluripotent Stem Cell-Derived Dopamine Neurons for Parkinson’s Disease Drug Development. Int. J. Mol. Sci. 2021, 22, 3381. [Google Scholar] [CrossRef] [PubMed]
  43. Creed, R.B.; Goldberg, M.S. New Developments in Genetic rat models of Parkinson’s Disease. Mov. Disord. 2018, 33, 717–729. [Google Scholar] [CrossRef] [PubMed]
  44. Mills, R.D.; Mulhern, T.D.; Liu, F.; Culvenor, J.G.; Cheng, H.C. Prediction of the repeat domain structures and impact of parkinsonism-associated variations on structure and function of all functional domains of leucine-rich repeat kinase 2 (LRRK2). Hum. Mutat. 2014, 35, 395–412. [Google Scholar] [CrossRef] [PubMed]
  45. Rudenko, I.N.; Kaganovich, A.; Langston, R.G.; Beilina, A.; Ndukwe, K.; Kumaran, R.; Dillman, A.A.; Chia, R.; Cookson, M.R. The G2385R risk factor for Parkinson’s disease enhances CHIP-dependent intracellular degradation of LRRK2. Biochem. J. 2017, 474, 1547–1558. [Google Scholar] [CrossRef] [PubMed]
  46. Wallings, R.L.; Tansey, M.G. LRRK2 regulation of immune-pathways and inflammatory disease. Biochem. Soc. Trans. 2019, 47, 1581–1595. [Google Scholar] [CrossRef]
  47. Webber, P.J.; Smith, A.D.; Sen, S.; Renfrow, M.B.; Mobley, J.A.; West, A.B. Autophosphorylation in the Leucine-Rich Repeat Kinase 2 (LRRK2) GTPase Domain Modifies Kinase and GTP-Binding Activities. J. Mol. Biol. 2011, 412, 94–110. [Google Scholar] [CrossRef] [Green Version]
  48. Smith, W.W.; Pei, Z.; Jiang, H.; Moore, D.J.; Liang, Y.; West, A.B.; Dawson, V.L.; Dawson, T.M.; Ross, C.A. Leucine-rich repeat kinase 2 (LRRK2) interacts with parkin, and mutant LRRK2 induces neuronal degeneration. Proc. Natl. Acad. Sci. USA 2005, 102, 18676–18681. [Google Scholar] [CrossRef] [Green Version]
  49. Ho, C.C.-Y.; Rideout, H.J.; Ribe, E.; Troy, C.M.; Dauer, W.T. The Parkinson Disease Protein LRRK2 Transduces Death Signals via FADD and Caspase-8 in a Cellular Model of Neurodegeneration. J. Neurosci. 2009, 29, 1011–1016. [Google Scholar] [CrossRef] [Green Version]
  50. Imai, Y.; Gehrke, S.; Wang, H.Q.; Takahashi, R.; Hasegawa, K.; Oota, E.; Lu, B. Phosphorylation of 4E-BP by LRRK2 affects the maintenance of dopaminergic neurons in Drosophila. EMBO J. 2008, 27, 2432–2443. [Google Scholar] [CrossRef]
  51. Kanao, T.; Venderova, K.; Park, D.S.; Unterman, T.; Lu, B.; Imai, Y. Activation of FoxO by LRRK2 induces expression of proapoptotic proteins and alters survival of postmitotic dopaminergic neuron in Drosophila. Hum. Mol. Genet. 2010, 19, 3747–3758. [Google Scholar] [CrossRef] [Green Version]
  52. Cookson, M.R. The role of leucine-rich repeat kinase 2 (LRRK2) in Parkinson’s disease. Nat. Rev. Neurosci. 2010, 11, 791–797. [Google Scholar] [CrossRef]
  53. Migheli, R.; Del Giudice, M.G.; Spissu, Y.; Sanna, G.; Xiong, Y.; Dawson, T.M.; Dawson, V.L.; Galioto, M.; Rocchitta, G.; Biosa, A.; et al. LRRK2 Affects Vesicle Trafficking, Neurotransmitter Extracellular Level and Membrane Receptor Localization. PLoS ONE 2013, 8, e77198. [Google Scholar] [CrossRef] [Green Version]
  54. Miklossy, J.; Arai, T.; Guo, J.-P.; Klegeris, A.; Yu, S.; McGeer, E.G.; McGeer, P.L. LRRK2 Expression in Normal and Pathologic Human Brain and in Human Cell Lines. J. Neuropathol. Exp. Neurol. 2006, 65, 953–963. [Google Scholar] [CrossRef]
  55. Dzamko, N.; Deak, M.; Hentati, F.; Reith, A.D.; Prescott, A.R.; Alessi, D.R.; Nichols, R.J. Inhibition of LRRK2 kinase activity leads to dephosphorylation of Ser(910)/Ser(935), disruption of 14-3-3 binding and altered cytoplasmic localization. Biochem. J. 2010, 430, 405–413. [Google Scholar] [CrossRef] [Green Version]
  56. Nichols, R.J.; Dzamko, N.; Morrice, N.A.; Campbell, D.G.; Deak, M.; Ordureau, A.; Macartney, T.; Tong, Y.; Shen, J.; Prescott, A.R.; et al. 14-3-3 binding to LRRK2 is disrupted by multiple Parkinson’s disease-associated mutations and regulates cytoplasmic localization. Biochem. J. 2010, 430, 393–404. [Google Scholar] [CrossRef] [Green Version]
  57. Tong, Y.; Yamaguchi, H.; Giaime, E.; Boyle, S.; Kopan, R.; Kelleher, R.J., 3rd; Shen, J. Loss of leucine-rich repeat kinase 2 causes impairment of protein degradation pathways, accumulation of α-synuclein, and apoptotic cell death in aged mice. Proc. Natl. Acad. Sci. USA 2010, 107, 9879–9884. [Google Scholar] [CrossRef] [Green Version]
  58. Cook, D.A.; Kannarkat, G.T.; Cintron, A.F.; Butkovich, L.M.; Fraser, K.B.; Chang, J.; Grigoryan, N.; Factor, S.A.; West, A.B.; Boss, J.M.; et al. LRRK2 levels in immune cells are increased in Parkinson’s disease. NPJ Park. Dis. 2017, 3, 11. [Google Scholar] [CrossRef]
  59. Dzamko, N.; Gysbers, A.M.; Bandopadhyay, R.; Bolliger, M.F.; Uchino, A.; Zhao, Y.; Takao, M.; Wauters, S.; van de Berg, W.D.; Takahashi-Fujigasaki, J.; et al. LRRK2 levels and phosphorylation in Parkinson’s disease brain and cases with restricted Lewy bodies. Mov. Disord. 2017, 32, 423–432. [Google Scholar] [CrossRef]
  60. Atashrazm, F.; Hammond, D.; Perera, G.; Bolliger, M.F.; Matar, E.; Halliday, G.M.; Schule, B.; Lewis, S.J.G.; Nichols, R.J.; Dzamko, N. LRRK2-mediated Rab10 phosphorylation in immune cells from Parkinson’s disease patients. Mov. Disord. 2019, 34, 406–415. [Google Scholar] [CrossRef]
  61. Nichols, R.J. LRRK2 Phosphorylation. Adv. Neurobiol. 2017, 14, 51–70. [Google Scholar] [CrossRef] [PubMed]
  62. Sheng, Z.; Zhang, S.; Bustos, D.; Kleinheinz, T.; Le Pichon, C.E.; Dominguez, S.L.; Solanoy, H.O.; Drummond, J.; Zhang, X.; Ding, X.; et al. Ser1292 autophosphorylation is an indicator of LRRK2 kinase activity and contributes to the cellular effects of PD mutations. Sci. Transl. Med. 2012, 4, 164ra161. [Google Scholar] [CrossRef] [PubMed]
  63. West, A.B.; Moore, D.J.; Biskup, S.; Bugayenko, A.; Smith, W.W.; Ross, C.A.; Dawson, V.L.; Dawson, T.M. Parkinson’s disease-associated mutations in leucine-rich repeat kinase 2 augment kinase activity. Proc. Natl. Acad. Sci. USA 2005, 102, 16842–16847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Greggio, E.; Jain, S.; Kingsbury, A.; Bandopadhyay, R.; Lewis, P.; Kaganovich, A.; van der Brug, M.P.; Beilina, A.; Blackinton, J.; Thomas, K.J.; et al. Kinase activity is required for the toxic effects of mutant LRRK2/dardarin. Neurobiol. Dis. 2006, 23, 329–341. [Google Scholar] [CrossRef] [PubMed]
  65. Smith, W.W.; Pei, Z.; Jiang, H.; Dawson, V.L.; Dawson, T.M.; Ross, C.A. Kinase activity of mutant LRRK2 mediates neuronal toxicity. Nat. Neurosci. 2006, 9, 1231–1233. [Google Scholar] [CrossRef]
  66. Lee, S.; Liu, H.P.; Lin, W.Y.; Guo, H.; Lu, B. LRRK2 kinase regulates synaptic morphology through distinct substrates at the presynaptic and postsynaptic compartments of the Drosophila neuromuscular junction. J. Neurosci. 2010, 30, 16959–16969. [Google Scholar] [CrossRef] [Green Version]
  67. Fuji, R.N.; Flagella, M.; Baca, M.; Baptista, M.A.; Brodbeck, J.; Chan, B.K.; Fiske, B.K.; Honigberg, L.; Jubb, A.M.; Katavolos, P.; et al. Effect of selective LRRK2 kinase inhibition on nonhuman primate lung. Sci. Transl. Med. 2015, 7, 273ra215. [Google Scholar] [CrossRef]
  68. Baptista, M.A.; Dave, K.D.; Frasier, M.A.; Sherer, T.B.; Greeley, M.; Beck, M.J.; Varsho, J.S.; Parker, G.A.; Moore, C.; Churchill, M.J.; et al. Loss of leucine-rich repeat kinase 2 (LRRK2) in rats leads to progressive abnormal phenotypes in peripheral organs. PLoS ONE 2013, 8, e80705. [Google Scholar] [CrossRef] [Green Version]
  69. Herzig, M.C.; Kolly, C.; Persohn, E.; Theil, D.; Schweizer, T.; Hafner, T.; Stemmelen, C.; Troxler, T.J.; Schmid, P.; Danner, S.; et al. LRRK2 protein levels are determined by kinase function and are crucial for kidney and lung homeostasis in mice. Hum. Mol. Genet. 2011, 20, 4209–4223. [Google Scholar] [CrossRef] [Green Version]
  70. Kawakami, F.; Yabata, T.; Ohta, E.; Maekawa, T.; Shimada, N.; Suzuki, M.; Maruyama, H.; Ichikawa, T.; Obata, F. LRRK2 phosphorylates tubulin-associated tau but not the free molecule: LRRK2-mediated regulation of the tau-tubulin association and neurite outgrowth. PLoS ONE 2012, 7, e30834. [Google Scholar] [CrossRef] [Green Version]
  71. Bailey, R.M.; Covy, J.P.; Melrose, H.L.; Rousseau, L.; Watkinson, R.; Knight, J.; Miles, S.; Farrer, M.J.; Dickson, D.W.; Giasson, B.I.; et al. LRRK2 phosphorylates novel tau epitopes and promotes tauopathy. Acta Neuropathol. 2013, 126, 809–827. [Google Scholar] [CrossRef] [Green Version]
  72. Krumova, P.; Reyniers, L.; Meyer, M.; Lobbestael, E.; Stauffer, D.; Gerrits, B.; Muller, L.; Hoving, S.; Kaupmann, K.; Voshol, J.; et al. Chemical genetic approach identifies microtubule affinity-regulating kinase 1 as a leucine-rich repeat kinase 2 substrate. FASEB J. 2015, 29, 2980–2992. [Google Scholar] [CrossRef] [Green Version]
  73. Gillardon, F. Leucine-rich repeat kinase 2 phosphorylates brain tubulin-beta isoforms and modulates microtubule stability—A point of convergence in Parkinsonian neurodegeneration? J. Neurochem. 2009, 110, 1514–1522. [Google Scholar] [CrossRef]
  74. Haebig, K.; Gloeckner, C.J.; Miralles, M.G.; Gillardon, F.; Schulte, C.; Riess, O.; Ueffing, M.; Biskup, S.; Bonin, M. ARHGEF7 (Beta-PIX) acts as guanine nucleotide exchange factor for leucine-rich repeat kinase 2. PLoS ONE 2010, 5, e13762. [Google Scholar] [CrossRef]
  75. Jaleel, M.; Nichols, R.J.; Deak, M.; Campbell, D.G.; Gillardon, F.; Knebel, A.; Alessi, D.R. LRRK2 phosphorylates moesin at threonine-558: Characterization of how Parkinson’s disease mutants affect kinase activity. Biochem. J. 2007, 405, 307–317. [Google Scholar] [CrossRef] [Green Version]
  76. Parisiadou, L.; Xie, C.; Cho, H.J.; Lin, X.; Gu, X.L.; Long, C.X.; Lobbestael, E.; Baekelandt, V.; Taymans, J.M.; Sun, L.; et al. Phosphorylation of ezrin/radixin/moesin proteins by LRRK2 promotes the rearrangement of actin cytoskeleton in neuronal morphogenesis. J. Neurosci. 2009, 29, 13971–13980. [Google Scholar] [CrossRef]
  77. Gandhi, P.N.; Wang, X.; Zhu, X.; Chen, S.G.; Wilson-Delfosse, A.L. The Roc domain of leucine-rich repeat kinase 2 is sufficient for interaction with microtubules. J. Neurosci. Res. 2008, 86, 1711–1720. [Google Scholar] [CrossRef] [Green Version]
  78. MacLeod, D.; Dowman, J.; Hammond, R.; Leete, T.; Inoue, K.; Abeliovich, A. The familial Parkinsonism gene LRRK2 regulates neurite process morphology. Neuron 2006, 52, 587–593. [Google Scholar] [CrossRef] [Green Version]
  79. Godena, V.K.; Brookes-Hocking, N.; Moller, A.; Shaw, G.; Oswald, M.; Sancho, R.M.; Miller, C.C.; Whitworth, A.J.; de Vos, K.J. Increasing microtubule acetylation rescues axonal transport and locomotor deficits caused by LRRK2 Roc-COR domain mutations. Nat. Commun. 2014, 5, 5245. [Google Scholar] [CrossRef] [Green Version]
  80. Steger, M.; Tonelli, F.; Ito, G.; Davies, P.; Trost, M.; Vetter, M.; Wachter, S.; Lorentzen, E.; Duddy, G.; Wilson, S.; et al. Phosphoproteomics reveals that Parkinson’s disease kinase LRRK2 regulates a subset of Rab GTPases. Elife 2016, 5, e12813. [Google Scholar] [CrossRef]
  81. Rivero-Rios, P.; Romo-Lozano, M.; Fernandez, B.; Fdez, E.; Hilfiker, S. Distinct Roles for RAB10 and RAB29 in Pathogenic LRRK2-Mediated Endolysosomal Trafficking Alterations. Cells 2020, 9, 1719. [Google Scholar] [CrossRef] [PubMed]
  82. Di Maio, R.; Hoffman, E.K.; Rocha, E.M.; Keeney, M.T.; Sanders, L.H.; De Miranda, B.R.; Zharikov, A.; Van Laar, A.; Stepan, A.F.; Lanz, T.A.; et al. LRRK2 activation in idiopathic Parkinson’s disease. Sci. Transl. Med. 2018, 10, eaar5429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Piccoli, G.; Condliffe, S.B.; Bauer, M.; Giesert, F.; Boldt, K.; de Astis, S.; Meixner, A.; Sarioglu, H.; Vogt-Weisenhorn, D.M.; Wurst, W.; et al. LRRK2 controls synaptic vesicle storage and mobilization within the recycling pool. J. Neurosci. 2011, 31, 2225–2237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Cirnaru, M.D.; Marte, A.; Belluzzi, E.; Russo, I.; Gabrielli, M.; Longo, F.; Arcuri, L.; Murru, L.; Bubacco, L.; Matteoli, M.; et al. LRRK2 kinase activity regulates synaptic vesicle trafficking and neurotransmitter release through modulation of LRRK2 macro-molecular complex. Front. Mol. Neurosci. 2014, 7, 49. [Google Scholar] [CrossRef] [Green Version]
  85. Yun, H.J.; Park, J.; Ho, D.H.; Kim, H.; Kim, C.H.; Oh, H.; Ga, I.; Seo, H.; Chang, S.; Son, I.; et al. LRRK2 phosphorylates Snapin and inhibits interaction of Snapin with SNAP-25. Exp. Mol. Med. 2013, 45, e36. [Google Scholar] [CrossRef] [Green Version]
  86. Matta, S.; van Kolen, K.; da Cunha, R.; van den Bogaart, G.; Mandemakers, W.; Miskiewicz, K.; de Bock, P.J.; Morais, V.A.; Vilain, S.; Haddad, D.; et al. LRRK2 controls an EndoA phosphorylation cycle in synaptic endocytosis. Neuron 2012, 75, 1008–1021. [Google Scholar] [CrossRef] [Green Version]
  87. Arranz, A.M.; Delbroek, L.; van Kolen, K.; Guimaraes, M.R.; Mandemakers, W.; Daneels, G.; Matta, S.; Calafate, S.; Shaban, H.; Baatsen, P.; et al. LRRK2 functions in synaptic vesicle endocytosis through a kinase-dependent mechanism. J. Cell. Sci. 2015, 128, 541–552. [Google Scholar] [CrossRef] [Green Version]
  88. Yun, H.J.; Kim, H.; Ga, I.; Oh, H.; Ho, D.H.; Kim, J.; Seo, H.; Son, I.; Seol, W. An early endosome regulator, Rab5b, is an LRRK2 kinase substrate. J. Biochem. 2015, 157, 485–495. [Google Scholar] [CrossRef]
  89. Schapansky, J.; Nardozzi, J.D.; Felizia, F.; LaVoie, M.J. Membrane recruitment of endogenous LRRK2 precedes its potent regulation of autophagy. Hum. Mol. Genet. 2014, 23, 4201–4214. [Google Scholar] [CrossRef] [Green Version]
  90. Gomez-Suaga, P.; Fdez, E.; Blanca Ramirez, M.; Hilfiker, S. A Link between Autophagy and the Pathophysiology of LRRK2 in Parkinson’s Disease. Park. Dis. 2012, 2012, 324521. [Google Scholar] [CrossRef]
  91. Manzoni, C.; Mamais, A.; Dihanich, S.; McGoldrick, P.; Devine, M.J.; Zerle, J.; Kara, E.; Taanman, J.W.; Healy, D.G.; Marti-Masso, J.F.; et al. Pathogenic Parkinson’s disease mutations across the functional domains of LRRK2 alter the autophagic/lysosomal response to starvation. Biochem. Biophys. Res. Commun. 2013, 441, 862–866. [Google Scholar] [CrossRef] [Green Version]
  92. Trancikova, A.; Mamais, A.; Webber, P.J.; Stafa, K.; Tsika, E.; Glauser, L.; West, A.B.; Bandopadhyay, R.; Moore, D.J. Phosphorylation of 4E-BP1 in the mammalian brain is not altered by LRRK2 expression or pathogenic mutations. PLoS ONE 2012, 7, e47784. [Google Scholar] [CrossRef] [Green Version]
  93. Martin, I.; Abalde-Atristain, L.; Kim, J.W.; Dawson, T.M.; Dawson, V.L. Abberant protein synthesis in G2019S LRRK2 Drosophila Parkinson disease-related phenotypes. Fly 2014, 8, 165–169. [Google Scholar] [CrossRef] [Green Version]
  94. Paisan-Ruiz, C.; Jain, S.; Evans, E.W.; Gilks, W.P.; Simon, J.; van der Brug, M.; Lopez de Munain, A.; Aparicio, S.; Gil, A.M.; Khan, N.; et al. Cloning of the gene containing mutations that cause PARK8-linked Parkinson’s disease. Neuron 2004, 44, 595–600. [Google Scholar] [CrossRef] [Green Version]
  95. 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]
  96. Pankratz, N.; Foroud, T. Genetics of Parkinson disease. Genet. Med. 2007, 9, 801–811. [Google Scholar] [CrossRef] [Green Version]
  97. Klein, C.; Westenberger, A. Genetics of Parkinson’s disease. Cold Spring Harb. Perspect Med. 2012, 2, a008888. [Google Scholar] [CrossRef] [Green Version]
  98. Li, X.X.; Liao, Q.; Xia, H.; Yang, X.L. Association between Parkinson’s disease and G2019S and R1441C mutations of the LRRK2 gene. Exp. Ther. Med. 2015, 10, 1450–1454. [Google Scholar] [CrossRef]
  99. Fu, X.; Zheng, Y.; Hong, H.; He, Y.; Zhou, S.; Guo, C.; Liu, Y.; Xian, W.; Zeng, J.; Li, J.; et al. LRRK2 G2385R and LRRK2 R1628P increase risk of Parkinson’s disease in a Han Chinese population from Southern Mainland China. Park. Relat. Disord. 2013, 19, 397–398. [Google Scholar] [CrossRef]
  100. Funayama, M.; Li, Y.; Tomiyama, H.; Yoshino, H.; Imamichi, Y.; Yamamoto, M.; Murata, M.; Toda, T.; Mizuno, Y.; Hattori, N. Leucine-rich repeat kinase 2 G2385R variant is a risk factor for Parkinson disease in Asian population. Neuroreport 2007, 18, 273–275. [Google Scholar] [CrossRef]
  101. Di Fonzo, A.; Wu-Chou, Y.H.; Lu, C.S.; van Doeselaar, M.; Simons, E.J.; Rohe, C.F.; Chang, H.C.; Chen, R.S.; Weng, Y.H.; Vanacore, N.; et al. A common missense variant in the LRRK2 gene, Gly2385Arg, associated with Parkinson’s disease risk in Taiwan. Neurogenetics 2006, 7, 133–138. [Google Scholar] [CrossRef] [PubMed]
  102. Tan, E.K.; Peng, R.; Teo, Y.Y.; Tan, L.C.; Angeles, D.; Ho, P.; Chen, M.L.; Lin, C.H.; Mao, X.Y.; Chang, X.L.; et al. Multiple LRRK2 variants modulate risk of Parkinson disease: A Chinese multicenter study. Hum. Mutat. 2010, 31, 561–568. [Google Scholar] [CrossRef] [PubMed]
  103. Shu, L.; Zhang, Y.; Sun, Q.; Pan, H.; Tang, B. A Comprehensive Analysis of Population Differences in LRRK2 Variant Distribution in Parkinson’s Disease. Front. Aging Neurosci. 2019, 11, 13. [Google Scholar] [CrossRef] [PubMed]
  104. Mata, I.F.; Kachergus, J.M.; Taylor, J.P.; Lincoln, S.; Aasly, J.; Lynch, T.; Hulihan, M.M.; Cobb, S.A.; Wu, R.M.; Lu, C.S.; et al. Lrrk2 pathogenic substitutions in Parkinson’s disease. Neurogenetics 2005, 6, 171–177. [Google Scholar] [CrossRef] [PubMed]
  105. Khan, N.L.; Jain, S.; Lynch, J.M.; Pavese, N.; Abou-Sleiman, P.; Holton, J.L.; Healy, D.G.; Gilks, W.P.; Sweeney, M.G.; Ganguly, M.; et al. Mutations in the gene LRRK2 encoding dardarin (PARK8) cause familial Parkinson’s disease: Clinical, pathological, olfactory and functional imaging and genetic data. Brain 2005, 128, 2786–2796. [Google Scholar] [CrossRef]
  106. Kachergus, J.; Mata, I.F.; Hulihan, M.; Taylor, J.P.; Lincoln, S.; Aasly, J.; Gibson, J.M.; Ross, O.A.; Lynch, T.; Wiley, J.; et al. Identification of a novel LRRK2 mutation linked to autosomal dominant parkinsonism: Evidence of a common founder across European populations. Am. J. Hum. Genet. 2005, 76, 672–680. [Google Scholar] [CrossRef] [Green Version]
  107. Lu, C.S.; Simons, E.J.; Wu-Chou, Y.H.; Fonzo, A.D.; Chang, H.C.; Chen, R.S.; Weng, Y.H.; Rohe, C.F.; Breedveld, G.J.; Hattori, N.; et al. The LRRK2 I2012T, G2019S, and I2020T mutations are rare in Taiwanese patients with sporadic Parkinson’s disease. Park. Relat. Disord. 2005, 11, 521–522. [Google Scholar] [CrossRef]
  108. Lewis, P.A.; Greggio, E.; Beilina, A.; Jain, S.; Baker, A.; Cookson, M.R. The R1441C mutation of LRRK2 disrupts GTP hydrolysis. Biochem. Biophys. Res. Commun. 2007, 357, 668–671. [Google Scholar] [CrossRef] [Green Version]
  109. Liao, J.; Wu, C.X.; Burlak, C.; Zhang, S.; Sahm, H.; Wang, M.; Zhang, Z.Y.; Vogel, K.W.; Federici, M.; Riddle, S.M.; et al. Parkinson disease-associated mutation R1441H in LRRK2 prolongs the “active state” of its GTPase domain. Proc. Natl. Acad. Sci. USA 2014, 111, 4055–4060. [Google Scholar] [CrossRef] [Green Version]
  110. Puschmann, A.; Englund, E.; Ross, O.A.; Vilarino-Guell, C.; Lincoln, S.J.; Kachergus, J.M.; Cobb, S.A.; Tornqvist, A.L.; Rehncrona, S.; Widner, H.; et al. First neuropathological description of a patient with Parkinson’s disease and LRRK2 p.N1437H mutation. Park. Relat. Disord. 2012, 18, 332–338. [Google Scholar] [CrossRef]
  111. Funayama, M.; Hasegawa, K.; Kowa, H.; Saito, M.; Tsuji, S.; Obata, F. A new locus for Parkinson’s Disease (PARK8) maps to chromosome 12p11.2-q13.1. Ann. Neurol. 2002, 51, 296–301. [Google Scholar] [CrossRef]
  112. Gloeckner, C.J.; Kinkl, N.; Schumacher, A.; Braun, R.J.; O’Neill, E.; Meitinger, T.; Kolch, W.; Prokisch, H.; Ueffing, M. The Parkinson disease causing LRRK2 mutation I2020T is associated with increased kinase activity. Hum. Mol. Genet. 2006, 15, 223–232. [Google Scholar] [CrossRef]
  113. Greggio, E.; Cookson, M.R. Leucine-rich repeat kinase 2 mutations and Parkinson’s disease: Three questions. ASN Neuro. 2009, 1, e00002. [Google Scholar] [CrossRef] [Green Version]
  114. Gilks, W.P.; Abou-Sleiman, P.M.; Gandhi, S.; Jain, S.; Singleton, A.; Lees, A.J.; Shaw, K.; Bhatia, K.P.; Bonifati, V.; Quinn, N.P.; et al. A common LRRK2 mutation in idiopathic Parkinson’s disease. Lancet 2005, 365, 415–416. [Google Scholar] [CrossRef]
  115. Lesage, S.; Durr, A.; Tazir, M.; Lohmann, E.; Leutenegger, A.L.; Janin, S.; Pollak, P.; Brice, A.; French Parkinson’s Disease Genetics Study Group. LRRK2 G2019S as a cause of Parkinson’s disease in North African Arabs. N. Engl. J. Med. 2006, 354, 422–423. [Google Scholar] [CrossRef]
  116. Ozelius, L.J.; Senthil, G.; Saunders-Pullman, R.; Ohmann, E.; Deligtisch, A.; Tagliati, M.; Hunt, A.L.; Klein, C.; Henick, B.; Hailpern, S.M.; et al. LRRK2 G2019S as a cause of Parkinson’s disease in Ashkenazi Jews. N. Engl. J. Med. 2006, 354, 424–425. [Google Scholar] [CrossRef]
  117. Dachsel, J.C.; Farrer, M.J. LRRK2 and Parkinson disease. Arch. Neurol. 2010, 67, 542–547. [Google Scholar] [CrossRef]
  118. Tan, E.K.; Shen, H.; Tan, L.C.; Farrer, M.; Yew, K.; Chua, E.; Jamora, R.D.; Puvan, K.; Puong, K.Y.; Zhao, Y.; et al. The G2019S LRRK2 mutation is uncommon in an Asian cohort of Parkinson’s disease patients. Neurosci. Lett. 2005, 384, 327–329. [Google Scholar] [CrossRef]
  119. Fung, H.C.; Chen, C.M.; Hardy, J.; Hernandez, D.; Singleton, A.; Wu, Y.R. Lack of G2019S LRRK2 mutation in a cohort of Taiwanese with sporadic Parkinson’s disease. Mov. Disord. 2006, 21, 880–881. [Google Scholar] [CrossRef]
  120. Tomiyama, H.; Li, Y.; Funayama, M.; Hasegawa, K.; Yoshino, H.; Kubo, S.; Sato, K.; Hattori, T.; Lu, C.S.; Inzelberg, R.; et al. Clinicogenetic study of mutations in LRRK2 exon 41 in Parkinson’s disease patients from 18 countries. Mov. Disord. 2006, 21, 1102–1108. [Google Scholar] [CrossRef]
  121. Kalinderi, K.; Fidani, L.; Bostantjopoulou, S.; Katsarou, Z.; Kotsis, A. The G2019S LRRK2 mutation is uncommon amongst Greek patients with sporadic Parkinson’s disease. Eur. J. Neurol. 2007, 14, 1088–1090. [Google Scholar] [CrossRef] [PubMed]
  122. Tan, E.K.; Zhao, Y.; Skipper, L.; Tan, M.G.; Di Fonzo, A.; Sun, L.; Fook-Chong, S.; Tang, S.; Chua, E.; Yuen, Y.; et al. The LRRK2 Gly2385Arg variant is associated with Parkinson’s disease: Genetic and functional evidence. Hum. Genet. 2007, 120, 857–863. [Google Scholar] [CrossRef] [PubMed]
  123. Gandhi, P.N.; Chen, S.G.; Wilson-Delfosse, A.L. Leucine-rich repeat kinase 2 (LRRK2): A key player in the pathogenesis of Parkinson’s disease. J. Neurosci. Res. 2009, 87, 1283–1295. [Google Scholar] [CrossRef] [Green Version]
  124. Luzón-Toro, B.; de la Torre, E.R.; Delgado, A.; Pérez-Tur, J.; Hilfiker, S. Mechanistic insight into the dominant mode of the Parkinson’s disease-associated G2019S LRRK2 mutation. Hum. Mol. Genet. 2007, 16, 2031–2039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Marras, C.; Alcalay, R.N.; Caspell-Garcia, C.; Coffey, C.; Chan, P.; Duda, J.E.; Facheris, M.F.; Fernandez-Santiago, R.; Ruiz-Martinez, J.; Mestre, T.; et al. Motor and nonmotor heterogeneity of LRRK2-related and idiopathic Parkinson’s disease. Mov. Disord. 2016, 31, 1192–1202. [Google Scholar] [CrossRef]
  126. Beilina, A.; Rudenko, I.N.; Kaganovich, A.; Civiero, L.; Chau, H.; Kalia, S.K.; Kalia, L.V.; Lobbestael, E.; Chia, R.; Ndukwe, K.; et al. Unbiased screen for interactors of leucine-rich repeat kinase 2 supports a common pathway for sporadic and familial Parkinson disease. Proc. Natl. Acad. Sci. USA 2014, 111, 2626–2631. [Google Scholar] [CrossRef] [Green Version]
  127. Ishihara, L.; Warren, L.; Gibson, R.; Amouri, R.; Lesage, S.; Durr, A.; Tazir, M.; Wszolek, Z.K.; Uitti, R.J.; Nichols, W.C.; et al. Clinical features of Parkinson disease patients with homozygous leucine-rich repeat kinase 2 G2019S mutations. Arch. Neurol. 2006, 63, 1250–1254. [Google Scholar] [CrossRef] [Green Version]
  128. Reinhardt, P.; Schmid, B.; Burbulla, L.F.; Schondorf, D.C.; Wagner, L.; Glatza, M.; Hoing, S.; Hargus, G.; Heck, S.A.; Dhingra, A.; et al. Genetic correction of a LRRK2 mutation in human iPSCs links parkinsonian neurodegeneration to ERK-dependent changes in gene expression. Cell. Stem. Cell. 2013, 12, 354–367. [Google Scholar] [CrossRef] [Green Version]
  129. Ho, D.H.; Seol, W.; Son, I. Upregulation of the p53–p21 pathway by G2019S LRRK2 contributes to the cellular senescence and accumulation of α-synuclein. Cell. Cycle 2019, 18, 467–475. [Google Scholar] [CrossRef] [Green Version]
  130. Ross, C.A.; Akimov, S.S. Human-induced pluripotent stem cells: Potential for neurodegenerative diseases. Hum. Mol. Genet. 2014, 23, 17–26. [Google Scholar] [CrossRef]
  131. Pu, J.; Jiang, H.; Zhang, B.; Feng, J. Redefining Parkinson’s disease research using induced pluripotent stem cells. Curr. Neurol. Neurosci. Rep. 2012, 12, 392–398. [Google Scholar] [CrossRef]
  132. Soldner, F.; Jaenisch, R. Medicine iPSC disease modeling. Science 2012, 338, 1155–1156. [Google Scholar] [CrossRef]
  133. Shi, Y.; Inoue, H.; Wu, J.C.; Yamanaka, S. Induced pluripotent stem cell technology: A decade of progress. Nat. Rev. Drug Discov. 2017, 16, 115–130. [Google Scholar] [CrossRef]
  134. Gunaseeli, I.; Doss, M.X.; Antzelevitch, C.; Hescheler, J.; Sachinidis, A. Induced pluripotent stem cells as a model for accelerated patient- and disease-specific drug discovery. Curr. Med. Chem. 2010, 17, 759–766. [Google Scholar] [CrossRef] [Green Version]
  135. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [Green Version]
  136. Tiscornia, G.; Vivas, E.L.; Belmonte, J.C.I. Diseases in a dish: Modeling human genetic disorders using induced pluripotent cells. Nat. Med. 2011, 17, 1570–1576. [Google Scholar] [CrossRef]
  137. Grobarczyk, B.; Franco, B.; Hanon, K.; Malgrange, B. Generation of Isogenic Human iPS Cell Line Precisely Corrected by Genome Editing Using the CRISPR/Cas9 System. Stem. Cell. Rev. Rep. 2015, 11, 774–787. [Google Scholar] [CrossRef]
  138. Bibikova, M.; Carroll, D.; Segal, D.J.; Trautman, J.K.; Smith, J.; Kim, Y.G.; Chandrasegaran, S. Stimulation of homologous recombination through targeted cleavage by chimeric nucleases. Mol. Cell. Biol. 2001, 21, 289–297. [Google Scholar] [CrossRef] [Green Version]
  139. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [Green Version]
  140. Soldner, F.; Laganiere, J.; Cheng, A.W.; Hockemeyer, D.; Gao, Q.; Alagappan, R.; Khurana, V.; Golbe, L.I.; Myers, R.H.; Lindquist, S.; et al. Generation of isogenic pluripotent stem cells differing exclusively at two early onset Parkinson point mutations. Cell 2011, 146, 318–331. [Google Scholar] [CrossRef]
  141. Gasiunas, G.; Barrangou, R.; Horvath, P.; Siksnys, V. Cas9-crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proc. Natl. Acad. Sci. USA 2012, 109, E2579–E2586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  143. Liu, G.H.; Qu, J.; Suzuki, K.; Nivet, E.; Li, M.; Montserrat, N.; Yi, F.; Xu, X.; Ruiz, S.; Zhang, W.; et al. Progressive degeneration of human neural stem cells caused by pathogenic LRRK2. Nature 2012, 491, 603–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Sanchez-Danes, A.; Richaud-Patin, Y.; Carballo-Carbajal, I.; Jimenez-Delgado, S.; Caig, C.; Mora, S.; Di Guglielmo, C.; Ezquerra, M.; Patel, B.; Giralt, A.; et al. Disease-specific phenotypes in dopamine neurons from human iPS-based models of genetic and sporadic Parkinson’s disease. EMBO Mol. Med. 2012, 4, 380–395. [Google Scholar] [CrossRef] [PubMed]
  145. Nguyen, H.N.; Byers, B.; Cord, B.; Shcheglovitov, A.; Byrne, J.; Gujar, P.; Kee, K.; Schüle, B.; Dolmetsch, R.E.; Langston, W.; et al. LRRK2 Mutant iPSC-Derived DA Neurons Demonstrate Increased Susceptibility to Oxidative Stress. Cell. Stem. Cell. 2011, 8, 267–280. [Google Scholar] [CrossRef] [Green Version]
  146. Bieri, G.; Brahic, M.; Bousset, L.; Couthouis, J.; Kramer, N.J.; Ma, R.; Nakayama, L.; Monbureau, M.; Defensor, E.; Schule, B.; et al. LRRK2 modifies α-syn pathology and spread in mouse models and human neurons. Acta Neuropathol. 2019, 137, 961–980. [Google Scholar] [CrossRef] [Green Version]
  147. Borgs, L.; Peyre, E.; Alix, P.; Hanon, K.; Grobarczyk, B.; Godin, J.D.; Purnelle, A.; Krusy, N.; Maquet, P.; Lefebvre, P.; et al. Dopaminergic neurons differentiating from LRRK2 G2019S induced pluripotent stem cells show early neuritic branching defects. Sci. Rep. 2016, 6, 33377. [Google Scholar] [CrossRef] [Green Version]
  148. Schwab, A.J.; Ebert, A.D. Neurite Aggregation and Calcium Dysfunction in iPSC-Derived Sensory Neurons with Parkinson’s Disease-Related LRRK2 G2019S Mutation. Stem. Cell. Rep. 2015, 5, 1039–1052. [Google Scholar] [CrossRef] [Green Version]
  149. Korecka, J.A.; Talbot, S.; Osborn, T.M.; de Leeuw, S.M.; Levy, S.A.; Ferrari, E.J.; Moskites, A.; Atkinson, E.; Jodelka, F.M.; Hinrich, A.J.; et al. Neurite Collapse and Altered ER Ca2+ Control in Human Parkinson Disease Patient iPSC-Derived Neurons with LRRK2 G2019S Mutation. Stem. Cell. Rep. 2019, 12, 29–41. [Google Scholar] [CrossRef] [Green Version]
  150. Kim, J.; Daadi, M.M. Non-cell autonomous mechanism of Parkinson’s disease pathology caused by G2019S LRRK2 mutation in Ashkenazi Jewish patient: Single cell analysis. Brain Res. 2019, 1722, 146342. [Google Scholar] [CrossRef]
  151. Walter, J.; Bolognin, S.; Antony, P.M.A.; Nickels, S.L.; Poovathingal, S.K.; Salamanca, L.; Magni, S.; Perfeito, R.; Hoel, F.; Qing, X.; et al. Neural Stem Cells of Parkinson’s Disease Patients Exhibit Aberrant Mitochondrial Morphology and Functionality. Stem. Cell. Rep. 2019, 12, 878–889. [Google Scholar] [CrossRef] [Green Version]
  152. Walter, J.; Bolognin, S.; Poovathingal, S.K.; Magni, S.; Gerard, D.; Antony, P.M.A.; Nickels, S.L.; Salamanca, L.; Berger, E.; Smits, L.M.; et al. The Parkinson’s-disease-associated mutation LRRK2-G2019S alters dopaminergic differentiation dynamics via NR2F1. Cell. Rep. 2021, 37, 109864. [Google Scholar] [CrossRef]
  153. Skibinski, G.; Hwang, V.; Ando, D.M.; Daub, A.; Lee, A.K.; Ravisankar, A.; Modan, S.; Finucane, M.M.; Shaby, B.A.; Finkbeiner, S. Nrf2 mitigates LRRK2- and α-synuclein-induced neurodegeneration by modulating proteostasis. Proc. Natl. Acad. Sci. USA 2017, 114, 1165–1170. [Google Scholar] [CrossRef] [Green Version]
  154. Sanders, L.H.; Laganiere, J.; Cooper, O.; Mak, S.K.; Vu, B.J.; Huang, Y.A.; Paschon, D.E.; Vangipuram, M.; Sundararajan, R.; Urnov, F.D.; et al. LRRK2 mutations cause mitochondrial DNA damage in iPSC-derived neural cells from Parkinson’s disease patients: Reversal by gene correction. Neurobiol. Dis. 2014, 62, 381–386. [Google Scholar] [CrossRef] [Green Version]
  155. Howlett, E.H.; Jensen, N.; Belmonte, F.; Zafar, F.; Hu, X.; Kluss, J.; Schule, B.; Kaufman, B.A.; Greenamyre, J.T.; Sanders, L.H. LRRK2 G2019S-induced mitochondrial DNA damage is LRRK2 kinase dependent and inhibition restores mtDNA integrity in Parkinson’s disease. Hum. Mol. Genet. 2017, 26, 4340–4351. [Google Scholar] [CrossRef] [Green Version]
  156. Hsieh, C.H.; Shaltouki, A.; Gonzalez, A.E.; Bettencourt da Cruz, A.; Burbulla, L.F.; St Lawrence, E.; Schule, B.; Krainc, D.; Palmer, T.D.; Wang, X. Functional Impairment in Miro Degradation and Mitophagy Is a Shared Feature in Familial and Sporadic Parkinson’s Disease. Cell. Stem. Cell. 2016, 19, 709–724. [Google Scholar] [CrossRef] [Green Version]
  157. Schwab, A.J.; Sison, S.L.; Meade, M.R.; Broniowska, K.A.; Corbett, J.A.; Ebert, A.D. Decreased Sirtuin Deacetylase Activity in LRRK2 G2019S iPSC-Derived Dopaminergic Neurons. Stem. Cell. Rep. 2017, 9, 1839–1852. [Google Scholar] [CrossRef] [Green Version]
  158. Pan, P.Y.; Li, X.; Wang, J.; Powell, J.; Wang, Q.; Zhang, Y.; Chen, Z.; Wicinski, B.; Hof, P.; Ryan, T.A.; et al. Parkinson’s Disease-Associated LRRK2 Hyperactive Kinase Mutant Disrupts Synaptic Vesicle Trafficking in Ventral Midbrain Neurons. J. Neurosci. 2017, 37, 11366–11376. [Google Scholar] [CrossRef] [Green Version]
  159. Connor-Robson, N.; Booth, H.; Martin, J.G.; Gao, B.; Li, K.; Doig, N.; Vowles, J.; Browne, C.; Klinger, L.; Juhasz, P.; et al. An integrated transcriptomics and proteomics analysis reveals functional endocytic dysregulation caused by mutations in LRRK2. Neurobiol. Dis. 2019, 127, 512–526. [Google Scholar] [CrossRef]
  160. Nguyen, M.; Krainc, D. LRRK2 phosphorylation of auxilin mediates synaptic defects in dopaminergic neurons from patients with Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2018, 115, 5576–5581. [Google Scholar] [CrossRef]
  161. Schapansky, J.; Khasnavis, S.; DeAndrade, M.P.; Nardozzi, J.D.; Falkson, S.R.; Boyd, J.D.; Sanderson, J.B.; Bartels, T.; Melrose, H.L.; LaVoie, M.J. Familial knockin mutation of LRRK2 causes lysosomal dysfunction and accumulation of endogenous insoluble α-synuclein in neurons. Neurobiol. Dis. 2018, 111, 26–35. [Google Scholar] [CrossRef] [PubMed]
  162. Su, Y.C.; Qi, X. Inhibition of excessive mitochondrialfissionreduced aberrant autophagy and neuronal damage caused by LRRK2 G2019S mutation. Hum. Mol. Genet. 2013, 22, 4545–4561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ho, D.H.; Kim, H.; Nam, D.; Sim, H.; Kim, J.; Kim, H.G.; Son, I.; Seol, W. LRRK2 impairs autophagy by mediating phosphorylation of leucyl-tRNA synthetase. Cell. Biochem. Funct. 2018, 36, 431–442. [Google Scholar] [CrossRef] [PubMed]
  164. Lee, H.; James, W.S.; Cowley, S.A. LRRK2 in peripheral and central nervous system innate immunity: Its link to Parkinson’s disease. Biochem. Soc. Trans. 2017, 45, 131–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Filippini, A.; Gennarelli, M.; Russo, I. Leucine-rich repeat kinase 2-related functions in GLIA: An update of the last years. Biochem. Soc. Trans. 2021, 49, 1375–1384. [Google Scholar] [CrossRef]
  166. Russo, I.; Bubacco, L.; Greggio, E. LRRK2 and neuroinflammation: Partners in crime in Parkinson’s disease? J. Neuroinflamm. 2014, 11, 52. [Google Scholar] [CrossRef] [Green Version]
  167. Li, T.; Yang, D.; Zhong, S.; Thomas, J.M.; Xue, F.; Liu, J.; Kong, L.; Voulalas, P.; Hassan, H.E.; Park, J.S.; et al. Novel LRRK2 GTP-binding inhibitors reduced degeneration in Parkinson’s disease cell and mouse models. Hum. Mol. Genet. 2014, 23, 6212–6222. [Google Scholar] [CrossRef] [Green Version]
  168. Li, T.; He, X.; Thomas, J.M.; Yang, D.; Zhong, S.; Xue, F.; Smith, W.W. A novel GTP-binding inhibitor, FX2149, attenuates LRRK2 toxicity in Parkinson’s disease models. PLoS ONE 2015, 10, e0122461. [Google Scholar] [CrossRef] [Green Version]
  169. Li, T.; Ning, B.; Kong, L.; Dai, B.; He, X.; Thomas, J.M.; Sawa, A.; Ross, C.A.; Smith, W.W. A LRRK2 GTP Binding Inhibitor, 68, Reduces LPS-Induced Signaling Events and TNF-alpha Release in Human Lymphoblasts. Cells 2021, 10, 480. [Google Scholar] [CrossRef]
  170. Caesar, M.; Felk, S.; Zach, S.; Bronstad, G.; Aasly, J.O.; Gasser, T.; Gillardon, F. Changes in matrix metalloprotease activity and progranulin levels may contribute to the pathophysiological function of mutant leucine-rich repeat kinase 2. Glia 2014, 62, 1075–1092. [Google Scholar] [CrossRef]
  171. Gillardon, F.; Schmid, R.; Draheim, H. Parkinson’s disease-linked leucine-rich repeat kinase 2(R1441G) mutation increases proinflammatory cytokine release from activated primary microglial cells and resultant neurotoxicity. Neuroscience 2012, 208, 41–48. [Google Scholar] [CrossRef]
  172. Ho, D.H.; Seol, W.; Eun, J.H.; Son, I.H. Phosphorylation of p53 by LRRK2 induces microglial tumor necrosis factor α-mediated neurotoxicity. Biochem. Biophys. Res. Commun. 2017, 482, 1088–1094. [Google Scholar] [CrossRef]
  173. Sonninen, T.M.; Hamalainen, R.H.; Koskuvi, M.; Oksanen, M.; Shakirzyanova, A.; Wojciechowski, S.; Puttonen, K.; Naumenko, N.; Goldsteins, G.; Laham-Karam, N.; et al. Metabolic alterations in Parkinson’s disease astrocytes. Sci. Rep. 2020, 10, 14474. [Google Scholar] [CrossRef]
  174. Russo, I.; Di Benedetto, G.; Kaganovich, A.; Ding, J.; Mercatelli, D.; Morari, M.; Cookson, M.R.; Bubacco, L.; Greggio, E. Leucine-rich repeat kinase 2 controls protein kinase A activation state through phosphodiesterase 4. J. Neuroinflamm. 2018, 15, 297. [Google Scholar] [CrossRef]
  175. Kim, B.; Yang, M.S.; Choi, D.; Kim, J.H.; Kim, H.S.; Seol, W.; Choi, S.; Jou, I.; Kim, E.Y.; Joe, E.H. Impaired inflammatory responses in murine Lrrk2-knockdown brain microglia. PLoS ONE 2012, 7, e34693. [Google Scholar] [CrossRef] [Green Version]
  176. Kim, J.; Pajarillo, E.; Rizor, A.; Son, D.S.; Lee, J.; Aschner, M.; Lee, E. LRRK2 kinase plays a critical role in manganese-induced inflammation and apoptosis in microglia. PLoS ONE 2019, 14, e0210248. [Google Scholar] [CrossRef] [Green Version]
  177. Russo, I.; Kaganovich, A.; Ding, J.; Landeck, N.; Mamais, A.; Varanita, T.; Biosa, A.; Tessari, I.; Bubacco, L.; Greggio, E.; et al. Transcriptome analysis of LRRK2 knock-out microglia cells reveals alterations of inflammatory- and oxidative stress-related pathways upon treatment with α-synuclein fibrils. Neurobiol. Dis. 2019, 129, 67–78. [Google Scholar] [CrossRef]
  178. Moehle, M.S.; Webber, P.J.; Tse, T.; Sukar, N.; Standaert, D.G.; DeSilva, T.M.; Cowell, R.M.; West, A.B. LRRK2 Inhibition Attenuates Microglial Inflammatory Responses. J. Neurosci. 2012, 32, 1602–1611. [Google Scholar] [CrossRef] [Green Version]
  179. Munoz, L.; Kavanagh, M.E.; Phoa, A.F.; Heng, B.; Dzamko, N.; Chen, E.J.; Doddareddy, M.R.; Guillemin, G.J.; Kassiou, M. Optimisation of LRRK2 inhibitors and assessment of functional efficacy in cell-based models of neuroinflammation. Eur. J. Med. Chem. 2015, 95, 29–34. [Google Scholar] [CrossRef] [PubMed]
  180. De Maturana, R.L.; Lang, V.; Zubiarrain, A.; Sousa, A.; Vazquez, N.; Gorostidi, A.; Aguila, J.; Lopez de Munain, A.; Rodriguez, M.; Sanchez-Pernaute, R. Mutations in LRRK2 impair NF-kappaB pathway in iPSC-derived neurons. J. Neuroinflamm. 2016, 13, 295. [Google Scholar] [CrossRef]
  181. Booth, H.D.E.; Wessely, F.; Connor-Robson, N.; Rinaldi, F.; Vowles, J.; Browne, C.; Evetts, S.G.; Hu, M.T.; Cowley, S.A.; Webber, C.; et al. RNA sequencing reveals MMP2 and TGFB1 downregulation in LRRK2 G2019S Parkinson’s iPSC-derived astrocytes. Neurobiol. Dis. 2019, 129, 56–66. [Google Scholar] [CrossRef] [PubMed]
  182. Speidel, A.; Felk, S.; Reinhardt, P.; Sterneckert, J.; Gillardon, F. Leucine-Rich Repeat Kinase 2 Influences Fate Decision of Human Monocytes Differentiated from Induced Pluripotent Stem Cells. PLoS ONE 2016, 11, e0165949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Cookson, M.R. LRRK2 Pathways Leading to Neurodegeneration. Curr. Neurol. Neurosci. Rep. 2015, 15, 42. [Google Scholar] [CrossRef] [PubMed]
  184. Nikonova, E.V.; Xiong, Y.; Tanis, K.Q.; Dawson, V.L.; Vogel, R.L.; Finney, E.M.; Stone, D.J.; Reynolds, I.J.; Kern, J.T.; Dawson, T.M. Transcriptional responses to loss or gain of function of the leucine-rich repeat kinase 2 (LRRK2) gene uncover biological processes modulated by LRRK2 activity. Hum. Mol. Genet. 2012, 21, 163–174. [Google Scholar] [CrossRef] [PubMed]
  185. Liu, Z.; Wang, X.; Yu, Y.; Li, X.; Wang, T.; Jiang, H.; Ren, Q.; Jiao, Y.; Sawa, A.; Moran, T.; et al. A Drosophila model for LRRK2-linked parkinsonism. Proc. Natl. Acad. Sci. USA 2008, 105, 2693–2698. [Google Scholar] [CrossRef] [Green Version]
  186. Venderova, K.; Kabbach, G.; Abdel-Messih, E.; Zhang, Y.; Parks, R.J.; Imai, Y.; Gehrke, S.; Ngsee, J.; Lavoie, M.J.; Slack, R.S.; et al. Leucine-Rich Repeat Kinase 2 interacts with Parkin, DJ-1 and PINK-1 in a Drosophila melanogaster model of Parkinson’s disease. Hum. Mol. Genet. 2009, 18, 4390–4404. [Google Scholar] [CrossRef]
  187. Ng, C.H.; Mok, S.Z.; Koh, C.; Ouyang, X.; Fivaz, M.L.; Tan, E.K.; Dawson, V.L.; Dawson, T.M.; Yu, F.; Lim, K.L. Parkin protects against LRRK2 G2019S mutant-induced dopaminergic neurodegeneration in Drosophila. J. Neurosci. 2009, 29, 11257–11262. [Google Scholar] [CrossRef] [Green Version]
  188. Hindle, S.; Afsari, F.; Stark, M.; Middleton, C.A.; Evans, G.J.; Sweeney, S.T.; Elliott, C.J. Dopaminergic expression of the Parkinsonian gene LRRK2-G2019S leads to non-autonomous visual neurodegeneration, accelerated by increased neural demands for energy. Hum. Mol. Genet. 2013, 22, 2129–2140. [Google Scholar] [CrossRef] [Green Version]
  189. Lin, C.H.; Tsai, P.I.; Wu, R.M.; Chien, C.T. LRRK2 G2019S Mutation Induces Dendrite Degeneration through Mislocalization and Phosphorylation of Tau by Recruiting Autoactivated GSK3. J. Neurosci. 2010, 30, 13138–13149. [Google Scholar] [CrossRef] [Green Version]
  190. Saha, S.; Guillily, M.D.; Ferree, A.; Lanceta, J.; Chan, D.; Ghosh, J.; Hsu, C.H.; Segal, L.; Raghavan, K.; Matsumoto, K.; et al. LRRK2 modulates vulnerability to mitochondrial dysfunction in Caenorhabditis elegans. J. Neurosci. 2009, 29, 9210–9218. [Google Scholar] [CrossRef] [Green Version]
  191. Saha, S.; Liu-Yesucevitz, L.; Wolozin, B. Regulation of autophagy by LRRK2 in Caenorhabditis elegans. Neurodegener. Dis. 2014, 13, 110–113. [Google Scholar] [CrossRef] [Green Version]
  192. Li, X.; Patel, J.C.; Wang, J.; Avshalumov, M.V.; Nicholson, C.; Buxbaum, J.D.; Elder, G.A.; Rice, M.E.; Yue, Z. Enhanced striatal dopamine transmission and motor performance with LRRK2 overexpression in mice is eliminated by familial Parkinson’s disease mutation G2019S. J. Neurosci. Off. J. Soc. Neurosci. 2010, 30, 1788–1797. [Google Scholar] [CrossRef] [Green Version]
  193. Chen, H.; Chan, B.K.; Drummond, J.; Estrada, A.A.; Gunzner-Toste, J.; Liu, X.; Liu, Y.; Moffat, J.; Shore, D.; Sweeney, Z.K.; et al. Discovery of selective LRRK2 inhibitors guided by computational analysis and molecular modeling. J. Med. Chem. 2012, 55, 5536–5545. [Google Scholar] [CrossRef]
  194. 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] [Green Version]
  195. Lee, J.W.; Cannon, J.R. LRRK2 mutations and neurotoxicant susceptibility. Exp. Biol. Med. 2015, 240, 752–759. [Google Scholar] [CrossRef]
  196. Walker, M.D.; Volta, M.; Cataldi, S.; Dinelle, K.; Beccano-Kelly, D.; Munsie, L.; Kornelsen, R.; Mah, C.; Chou, P.; Co, K.; et al. Behavioral deficits and striatal DA signaling in LRRK2 p.G2019S transgenic rats: A multimodal investigation including PET neuroimaging. J. Park. Dis. 2014, 4, 483–498. [Google Scholar] [CrossRef]
  197. Pingale, T.; Gupta, G.L. Classic and evolving animal models in Parkinson’s disease. Pharmacol. Biochem. Behav. 2020, 199, 173060. [Google Scholar] [CrossRef]
  198. Totterdell, S.; Meredith, G.E. Localization of alpha-synuclein to identified fibers and synapses in the normal mouse brain. Neuroscience 2005, 135, 907–913. [Google Scholar] [CrossRef]
  199. Alim, M.A.; Hossain, M.S.; Arima, K.; Takeda, K.; Izumiyama, Y.; Nakamura, M.; Kaji, H.; Shinoda, T.; Hisanaga, S.; Ueda, K. Tubulin seeds α-synuclein fibril formation. J. Biol. Chem. 2002, 277, 2112–2117. [Google Scholar] [CrossRef] [Green Version]
  200. Cole, N.B.; Murphy, D.D.; Grider, T.; Rueter, S.; Brasaemle, D.; Nussbaum, R.L. Lipid droplet binding and oligomerization properties of the Parkinson’s disease protein α-synuclein. J. Biol. Chem. 2002, 277, 6344–6352. [Google Scholar] [CrossRef]
  201. Mak, S.K.; McCormack, A.L.; Langston, J.W.; Kordower, J.H.; Di Monte, D.A. Decreased α-synuclein expression in the aging mouse substantia nigra. Exp. Neurol. 2009, 220, 359–365. [Google Scholar] [CrossRef] [PubMed]
  202. Michell, A.W.; Tofaris, G.K.; Gossage, H.; Tyers, P.; Spillantini, M.G.; Barker, R.A. The effect of truncated human α-synuclein (1–120) on dopaminergic cells in a transgenic mouse model of Parkinson’s disease. Cell. Transplant. 2007, 16, 461–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Schneider, B.L.; Seehus, C.R.; Capowski, E.E.; Aebischer, P.; Zhang, S.C.; Svendsen, C.N. Over-expression of alpha-synuclein in human neural progenitors leads to specific changes in fate and differentiation. Hum. Mol. Genet. 2007, 16, 651–666. [Google Scholar] [CrossRef] [PubMed]
  204. Purisai, M.G.; McCormack, A.L.; Langston, W.J.; Johnston, L.C.; di Monte, D.A. Alpha-synuclein expression in the substantia nigra of MPTP-lesioned non-human primates. Neurobiol. Dis. 2005, 20, 898–906. [Google Scholar] [CrossRef] [PubMed]
  205. Bodles, A.M.; Guthrie, D.J.; Greer, B.; Irvine, G.B. Identification of the region of non-Abeta component (NAC) of Alzheimer’s disease amyloid responsible for its aggregation and toxicity. J. Neurochem. 2001, 78, 384–395. [Google Scholar] [CrossRef]
  206. Halliday, G.M.; del Tredici, K.; Braak, H. Critical appraisal of brain pathology staging related to presymptomatic and symptomatic cases of sporadic Parkinson’s disease. J. Neural Transm. Suppl. 2006, 70, 99–103. [Google Scholar] [CrossRef]
  207. Burke, W.J.; Kumar, V.B.; Pandey, N.; Panneton, W.M.; Gan, Q.; Franko, M.W.; O’Dell, M.; Li, S.W.; Pan, Y.; Chung, H.D.; et al. Aggregation of α-synuclein by DOPAL, the monoamine oxidase metabolite of dopamine. Acta Neuropathol. 2008, 115, 193–203. [Google Scholar] [CrossRef]
  208. Sian-Hulsmann, J.; Monoranu, C.; Strobel, S.; Riederer, P. Lewy Bodies: A Spectator or Salient Killer? CNS Neurol. Disord. Drug Targets 2015, 14, 947–955. [Google Scholar] [CrossRef]
  209. Foley, P.; Riederer, P. Pathogenesis and preclinical course of Parkinson’s disease. J. Neural Transm. Suppl. 1999, 56, 31–74. [Google Scholar] [CrossRef]
  210. Conway, K.A.; Lee, S.J.; Rochet, J.C.; Ding, T.T.; Williamson, R.E.; Lansbury, P.T., Jr. Acceleration of oligomerization, not fibrillization, is a shared property of both α-synuclein mutations linked to early-onset Parkinson’s disease: Implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. USA 2000, 97, 571–576. [Google Scholar] [CrossRef]
  211. Ren, P.H.; Lauckner, J.E.; Kachirskaia, I.; Heuser, J.E.; Melki, R.; Kopito, R.R. Cytoplasmic penetration and persistent infection of mammalian cells by polyglutamine aggregates. Nat. Cell. Biol. 2009, 11, 219–225. [Google Scholar] [CrossRef] [Green Version]
  212. Sherer, T.B.; Betarbet, R.; Stout, A.K.; Lund, S.; Baptista, M.; Panov, A.V.; Cookson, M.R.; Greenamyre, J.T. An in vitro model of Parkinson’s disease: Linking mitochondrial impairment to altered α-synuclein metabolism and oxidative damage. J. Neurosci. 2002, 22, 7006–7015. [Google Scholar] [CrossRef] [Green Version]
  213. Lardenoije, R.; Iatrou, A.; Kenis, G.; Kompotis, K.; Steinbusch, H.W.; Mastroeni, D.; Coleman, P.; Lemere, C.A.; Hof, P.R.; van den Hove, D.L.; et al. The epigenetics of aging and neurodegeneration. Prog. Neurobiol. 2015, 131, 21–64. [Google Scholar] [CrossRef]
  214. Wong, Y.C.; Krainc, D. α-synuclein toxicity in neurodegeneration: Mechanism and therapeutic strategies. Nat. Med. 2017, 23, 1–13. [Google Scholar] [CrossRef]
  215. Surgucheva, I.; Newell, K.L.; Burns, J.; Surguchov, A. New α- and γ-synuclein immunopathological lesions in human brain. Acta Neuropathol. Commun. 2014, 2, 132. [Google Scholar] [CrossRef] [Green Version]
  216. Anderson, J.P.; Walker, D.E.; Goldstein, J.M.; de Laat, R.; Banducci, K.; Caccavello, R.J.; Barbour, R.; Huang, J.; Kling, K.; Lee, M.; et al. Phosphorylation of Ser-129 is the dominant pathological modification of α-synuclein in familial and sporadic Lewy body disease. J. Biol. Chem. 2006, 281, 29739–29752. [Google Scholar] [CrossRef] [Green Version]
  217. Sato, H.; Kato, T.; Arawaka, S. The role of Ser129 phosphorylation of α-synuclein in neurodegeneration of Parkinson’s disease: A review of in vivo models. Rev. Neurosci. 2013, 24, 115–123. [Google Scholar] [CrossRef]
  218. Waxman, E.A.; Giasson, B.I. Specificity and regulation of casein kinase-mediated phosphorylation of α-synuclein. J. Neuropathol. Exp. Neurol. 2008, 67, 402–416. [Google Scholar] [CrossRef] [Green Version]
  219. Waxman, E.A.; Giasson, B.I. Characterization of kinases involved in the phosphorylation of aggregated α-synuclein. J. Neurosci. Res. 2011, 89, 231–247. [Google Scholar] [CrossRef] [Green Version]
  220. Mbefo, M.K.; Paleologou, K.E.; Boucharaba, A.; Oueslati, A.; Schell, H.; Fournier, M.; Olschewski, D.; Yin, G.; Zweckstetter, M.; Masliah, E.; et al. Phosphorylation of synucleins by members of the Polo-like kinase family. J. Biol. Chem. 2010, 285, 2807–2822. [Google Scholar] [CrossRef]
  221. Inglis, K.J.; Chereau, D.; Brigham, E.F.; Chiou, S.S.; Schobel, S.; Frigon, N.L.; Yu, M.; Caccavello, R.J.; Nelson, S.; Motter, R.; et al. Polo-like kinase 2 (PLK2) phosphorylates α-synuclein at serine 129 in central nervous system. J. Biol. Chem. 2009, 284, 2598–2602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Pronin, A.N.; Morris, A.J.; Surguchov, A.; Benovic, J.L. Synucleins are a novel class of substrates for G protein-coupled receptor kinases. J. Biol. Chem. 2000, 275, 26515–26522. [Google Scholar] [CrossRef] [Green Version]
  223. Paleologou, K.E.; Oueslati, A.; Shakked, G.; Rospigliosi, C.C.; Kim, H.Y.; Lamberto, G.R.; Fernandez, C.O.; Schmid, A.; Chegini, F.; Gai, W.P.; et al. Phosphorylation at S87 is enhanced in synucleinopathies, inhibits α-synuclein oligomerization, and influences synuclein-membrane interactions. J. Neurosci. 2010, 30, 3184–3198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Sakamoto, M.; Arawaka, S.; Hara, S.; Sato, H.; Cui, C.; Machiya, Y.; Koyama, S.; Wada, M.; Kawanami, T.; Kurita, K.; et al. Contribution of endogenous G-protein-coupled receptor kinases to Ser129 phosphorylation of α-synuclein in HEK293 cells. Biochem. Biophys. Res. Commun. 2009, 384, 378–382. [Google Scholar] [CrossRef] [PubMed]
  225. Dzamko, N.; Zhou, J.; Huang, Y.; Halliday, G.M. Parkinson’s disease-implicated kinases in the brain; insights into disease pathogenesis. Front. Mol. Neurosci. 2014, 7, 57. [Google Scholar] [CrossRef] [Green Version]
  226. Herzig, M.C.; Bidinosti, M.; Schweizer, T.; Hafner, T.; Stemmelen, C.; Weiss, A.; Danner, S.; Vidotto, N.; Stauffer, D.; Barske, C.; et al. High LRRK2 levels fail to induce or exacerbate neuronal alpha-synucleinopathy in mouse brain. PLoS ONE 2012, 7, e36581. [Google Scholar] [CrossRef]
  227. Lin, X.; Parisiadou, L.; Gu, X.L.; Wang, L.; Shim, H.; Sun, L.; Xie, C.; Long, C.X.; Yang, W.J.; Ding, J.; et al. Leucine-rich repeat kinase 2 regulates the progression of neuropathology induced by Parkinson’s-disease-related mutant α-synuclein. Neuron 2009, 64, 807–827. [Google Scholar] [CrossRef] [Green Version]
  228. Buck, K.; Landeck, N.; Ulusoy, A.; Majbour, N.K.; El-Agnaf, O.M.; Kirik, D. Ser129 phosphorylation of endogenous α-synuclein induced by overexpression of polo-like kinases 2 and 3 in nigral dopamine neurons is not detrimental to their survival and function. Neurobiol. Dis. 2015, 78, 100–114. [Google Scholar] [CrossRef]
  229. 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]
  230. Braak, H.; Rub, U.; Gai, W.P.; del Tredici, K. Idiopathic Parkinson’s disease: Possible routes by which vulnerable neuronal types may be subject to neuroinvasion by an unknown pathogen. J. Neural Transm. 2003, 110, 517–536. [Google Scholar] [CrossRef]
  231. Pfeiffer, R.F. Gastrointestinal dysfunction in Parkinson’s disease. Park. Relat. Disord. 2011, 17, 10–15. [Google Scholar] [CrossRef]
  232. Cersosimo, M.G.; Benarroch, E.E. Pathological correlates of gastrointestinal dysfunction in Parkinson’s disease. Neurobiol. Dis. 2012, 46, 559–564. [Google Scholar] [CrossRef]
  233. Doty, R.L. Olfactory dysfunction in Parkinson disease. Nat. Rev. Neurol. 2012, 8, 329–339. [Google Scholar] [CrossRef]
  234. Beach, T.G.; White, C.L., 3rd; Hladik, C.L.; Sabbagh, M.N.; Connor, D.J.; Shill, H.A.; Sue, L.I.; Sasse, J.; Bachalakuri, J.; Henry-Watson, J.; et al. Olfactory bulb α-synucleinopathy has high specificity and sensitivity for Lewy body disorders. Acta Neuropathol. 2009, 117, 169–174. [Google Scholar] [CrossRef] [Green Version]
  235. Hubbard, P.S.; Esiri, M.M.; Reading, M.; McShane, R.; Nagy, Z. Alpha-synuclein pathology in the olfactory pathways of dementia patients. J. Anat. 2007, 211, 117–124. [Google Scholar] [CrossRef]
  236. Shannon, K.M.; Keshavarzian, A.; Mutlu, E.; Dodiya, H.B.; Daian, D.; Jaglin, J.A.; Kordower, J.H. Alpha-synuclein in colonic submucosa in early untreated Parkinson’s disease. Mov. Disord. 2012, 27, 709–715. [Google Scholar] [CrossRef]
  237. Braak, H.; de Vos, R.A.; Bohl, J.; Del Tredici, K. Gastric alpha-synuclein immunoreactive inclusions in Meissner’s and Auerbach’s plexuses in cases staged for Parkinson’s disease-related brain pathology. Neurosci. Lett. 2006, 396, 67–72. [Google Scholar] [CrossRef]
  238. Oueslati, A.; Ximerakis, M.; Vekrellis, K. Protein Transmission, Seeding and Degradation: Key Steps for α-Synuclein Prion-Like Propagation. Exp. Neurobiol. 2014, 23, 324–336. [Google Scholar] [CrossRef] [Green Version]
  239. Abdelmotilib, H.; Maltbie, T.; Delic, V.; Liu, Z.; Hu, X.; Fraser, K.B.; Moehle, M.S.; Stoyka, L.; Anabtawi, N.; Krendelchtchikova, V.; et al. α-Synuclein fibril-induced inclusion spread in rats and mice correlates with dopaminergic Neurodegeneration. Neurobiol. Dis. 2017, 105, 84–98. [Google Scholar] [CrossRef]
  240. Desplats, P.; Lee, H.J.; Bae, E.J.; Patrick, C.; Rockenstein, E.; Crews, L.; Spencer, B.; Masliah, E.; Lee, S.J. Inclusion formation and neuronal cell death through neuron-to-neuron transmission of α-synuclein. Proc. Natl. Acad. Sci. USA 2009, 106, 13010–13015. [Google Scholar] [CrossRef]
  241. Luk, K.C.; Kehm, V.; Carroll, J.; Zhang, B.; O’Brien, P.; Trojanowski, J.Q.; Lee, V.M. Pathological α-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science 2012, 338, 949–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Mougenot, A.L.; Nicot, S.; Bencsik, A.; Morignat, E.; Verchère, J.; Lakhdar, L.; Legastelois, S.; Baron, T. Prion-like acceleration of a synucleinopathy in a transgenic mouse model. Neurobiol. Aging 2012, 33, 2225–2228. [Google Scholar] [CrossRef] [PubMed]
  243. Paumier, K.L.; Luk, K.C.; Manfredsson, F.P.; Kanaan, N.M.; Lipton, J.W.; Collier, T.J.; Steece-Collier, K.; Kemp, C.J.; Celano, S.; Schulz, E.; et al. Intrastriatal injection of pre-formed mouse α-synuclein fibrils into rats triggers α-synuclein pathology and bilateral nigrostriatal degeneration. Neurobiol. Dis. 2015, 82, 185–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Peelaerts, W.; Bousset, L.; van der Perren, A.; Moskalyuk, A.; Pulizzi, R.; Giugliano, M.; van den Haute, C.; Melki, R.; Baekelandt, V. α-Synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature 2015, 522, 340–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. α-Synuclein locus triplication causes Parkinson’s disease. Science 2003, 302, 841. [Google Scholar] [CrossRef] [Green Version]
  246. Golbe, L.I.; di Iorio, G.; Bonavita, V.; Miller, D.C.; Duvoisin, R.C. A large kindred with autosomal dominant Parkinson’s disease. Ann. Neurol. 1990, 27, 276–282. [Google Scholar] [CrossRef]
  247. 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 α-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef]
  248. Kruger, R.; Kuhn, W.; Leenders, K.L.; Sprengelmeyer, R.; Muller, T.; Woitalla, D.; Portman, A.T.; Maguire, R.P.; Veenma, L.; Schroder, U.; et al. Familial parkinsonism with synuclein pathology: Clinical and PET studies of A30P mutation carriers. Neurology 2001, 56, 1355–1362. [Google Scholar] [CrossRef]
  249. Seidel, K.; Schols, L.; Nuber, S.; Petrasch-Parwez, E.; Gierga, K.; Wszolek, Z.; Dickson, D.; Gai, W.P.; Bornemann, A.; Riess, O.; et al. First appraisal of brain pathology owing to A30P mutant alpha-synuclein. Ann. Neurol. 2010, 67, 684–689. [Google Scholar] [CrossRef]
  250. 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 α-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef]
  251. Martikainen, M.H.; Paivarinta, M.; Hietala, M.; Kaasinen, V. Clinical and imaging findings in Parkinson disease associated with the A53E SNCA mutation. Neurol. Genet. 2015, 1, e27. [Google Scholar] [CrossRef]
  252. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Poyhonen, M.; Paetau, A. Novel α-synuclein mutation A53E associated with atypical multiple system atrophy and Parkinson’s disease-type pathology. Neurobiol. Aging 2014, 35, 2180 e2181–e2185. [Google Scholar] [CrossRef]
  253. 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, 248 248.e7–248.e12. [Google Scholar] [CrossRef]
  254. Kiely, A.P.; Asi, Y.T.; Kara, E.; Limousin, P.; Ling, H.; Lewis, P.; Proukakis, C.; Quinn, N.; Lees, A.J.; Hardy, J.; et al. α-Synucleinopathy associated with G51D SNCA mutation: A link between Parkinson’s disease and multiple system atrophy? Acta Neuropathol. 2013, 125, 753–769. [Google Scholar] [CrossRef] [Green Version]
  255. Kiely, A.P.; Ling, H.; Asi, Y.T.; Kara, E.; Proukakis, C.; Schapira, A.H.; Morris, H.R.; Roberts, H.C.; Lubbe, S.; Limousin, P.; et al. Distinct clinical and neuropathological features of G51D SNCA mutation cases compared with SNCA duplication and H50Q mutation. Mol. Neurodegener. 2015, 10, 41. [Google Scholar] [CrossRef] [Green Version]
  256. 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]
  257. Hoffman-Zacharska, D.; Koziorowski, D.; Ross, O.A.; Milewski, M.; Poznanski, J.A.; Jurek, M.; Wszolek, Z.K.; Soto-Ortolaza, A.; Awek, J.A.S.; Janik, P.; et al. Novel A18T and pA29S substitutions in α-synuclein may be associated with sporadic Parkinson’s disease. Park. Relat. Disord. 2013, 19, 1057–1060. [Google Scholar] [CrossRef] [Green Version]
  258. Lucking, C.B.; Brice, A. Alpha-synuclein and Parkinson’s disease. Cell. Mol. Life Sci. 2000, 57, 1894–1908. [Google Scholar] [CrossRef]
  259. Narhi, L.; Wood, S.J.; Steavenson, S.; Jiang, Y.; Wu, G.M.; Anafi, D.; Kaufman, S.A.; Martin, F.; Sitney, K.; Denis, P.; et al. Both familial Parkinson’s disease mutations accelerate α-synuclein aggregation. J. Biol. Chem. 1999, 274, 9843–9846. [Google Scholar] [CrossRef] [Green Version]
  260. Spira, P.J.; Sharpe, D.M.; Halliday, G.; Cavanagh, J.; Nicholson, G.A. Clinical and pathological features of a Parkinsonian syndrome in a family with an Ala53Thr α-synuclein mutation. Ann. Neurol. 2001, 49, 313–319. [Google Scholar] [CrossRef]
  261. Pandey, N.; Schmidt, R.E.; Galvin, J.E. The alpha-synuclein mutation E46K promotes aggregation in cultured cells. Exp. Neurol. 2006, 197, 515–520. [Google Scholar] [CrossRef] [PubMed]
  262. Devine, M.J.; Ryten, M.; Vodicka, P.; Thomson, A.J.; Burdon, T.; Houlden, H.; Cavaleri, F.; Nagano, M.; Drummond, N.J.; Taanman, J.W.; et al. Parkinson’s disease induced pluripotent stem cells with triplication of the α-synuclein locus. Nat. Commun. 2011, 2, 440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Flierl, A.; Oliveira, L.M.; Falomir-Lockhart, L.J.; Mak, S.K.; Hesley, J.; Soldner, F.; Arndt-Jovin, D.J.; Jaenisch, R.; Langston, J.W.; Jovin, T.M.; et al. Higher vulnerability and stress sensitivity of neuronal precursor cells carrying an alpha-synuclein gene triplication. PLoS ONE 2014, 9, e112413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Byers, B.; Cord, B.; Nguyen, H.N.; Schule, B.; Fenno, L.; Lee, P.C.; Deisseroth, K.; Langston, J.W.; Pera, R.R.; Palmer, T.D. SNCA triplication Parkinson’s patient’s iPSC-derived DA neurons accumulate α-synuclein and are susceptible to oxidative stress. PLoS ONE 2011, 6, e26159. [Google Scholar] [CrossRef] [PubMed]
  265. Oliveira, L.M.; Falomir-Lockhart, L.J.; Botelho, M.G.; Lin, K.H.; Wales, P.; Koch, J.C.; Gerhardt, E.; Taschenberger, H.; Outeiro, T.F.; Lingor, P.; et al. Elevated α-synuclein caused by SNCA gene triplication impairs neuronal differentiation and maturation in Parkinson’s patient-derived induced pluripotent stem cells. Cell Death Dis. 2015, 6, e1994. [Google Scholar] [CrossRef] [Green Version]
  266. Mazzulli, J.R.; Zunke, F.; Isacson, O.; Studer, L.; Krainc, D. α-Synuclein-induced lysosomal dysfunction occurs through disruptions in protein trafficking in human midbrain synucleinopathy models. Proc. Natl. Acad. Sci. USA 2016, 113, 1931–1936. [Google Scholar] [CrossRef] [Green Version]
  267. Ludtmann, M.H.R.; Angelova, P.R.; Horrocks, M.H.; Choi, M.L.; Rodrigues, M.; Baev, A.Y.; Berezhnov, A.V.; Yao, Z.; Little, D.; Banushi, B.; et al. α-synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease. Nat. Commun. 2018, 9, 2293. [Google Scholar] [CrossRef] [Green Version]
  268. Haenseler, W.; Zambon, F.; Lee, H.; Vowles, J.; Rinaldi, F.; Duggal, G.; Houlden, H.; Gwinn, K.; Wray, S.; Luk, K.C.; et al. Excess α-synuclein compromises phagocytosis in iPSC-derived macrophages. Sci. Rep. 2017, 7, 9003. [Google Scholar] [CrossRef] [Green Version]
  269. Chung, C.Y.; Khurana, V.; Auluck, P.K.; Tardiff, D.F.; Mazzulli, J.R.; Soldner, F.; Baru, V.; Lou, Y.; Freyzon, Y.; Cho, S.; et al. Identification and rescue of α-synuclein toxicity in Parkinson patient-derived neurons. Science 2013, 342, 983–987. [Google Scholar] [CrossRef] [Green Version]
  270. Ryan, S.D.; Dolatabadi, N.; Chan, S.F.; Zhang, X.; Akhtar, M.W.; Parker, J.; Soldner, F.; Sunico, C.R.; Nagar, S.; Talantova, M.; et al. Isogenic human iPSC Parkinson’s model shows nitrosative stress-induced dysfunction in MEF2-PGC1alpha transcription. Cell 2013, 155, 1351–1364. [Google Scholar] [CrossRef]
  271. Kouroupi, G.; Taoufik, E.; Vlachos, I.S.; Tsioras, K.; Antoniou, N.; Papastefanaki, F.; Chroni-Tzartou, D.; Wrasidlo, W.; Bohl, D.; Stellas, D.; et al. Defective synaptic connectivity and axonal neuropathology in a human iPSC-based model of familial Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2017, 114, E3679–E3688. [Google Scholar] [CrossRef] [Green Version]
  272. Dettmer, U.; Newman, A.J.; Soldner, F.; Luth, E.S.; Kim, N.C.; von Saucken, V.E.; Sanderson, J.B.; Jaenisch, R.; Bartels, T.; Selkoe, D. Parkinson-causing α-synuclein missense mutations shift native tetramers to monomers as a mechanism for disease initiation. Nat. Commun. 2015, 6, 7314. [Google Scholar] [CrossRef] [Green Version]
  273. Ono, K. The Oligomer Hypothesis in α-Synucleinopathy. Neurochem. Res. 2017, 42, 3362–3371. [Google Scholar] [CrossRef]
  274. Zambon, F.; Cherubini, M.; Fernandes, H.J.R.; Lang, C.; Ryan, B.J.; Volpato, V.; Bengoa-Vergniory, N.; Vingill, S.; Attar, M.; Booth, H.D.E.; et al. Cellular α-synuclein pathology is associated with bioenergetic dysfunction in Parkinson’s iPSC-derived dopamine neurons. Hum. Mol. Genet. 2019, 28, 2001–2013. [Google Scholar] [CrossRef]
  275. Stykel, M.G.; Humphries, K.; Kirby, M.P.; Czaniecki, C.; Wang, T.; Ryan, T.; Bamm, V.; Ryan, S.D. Nitration of microtubules blocks axonal mitochondrial transport in a human pluripotent stem cell model of Parkinson’s disease. FASEB J. 2018, 32, 5350–5364. [Google Scholar] [CrossRef] [Green Version]
  276. Snowden, S.G.; Fernandes, H.J.R.; Kent, J.; Foskolou, S.; Tate, P.; Field, S.F.; Metzakopian, E.; Koulman, A. Development and Application of High-Throughput Single Cell Lipid Profiling: A Study of SNCA-A53T Human Dopamine Neurons. iScience 2020, 23, 101703. [Google Scholar] [CrossRef]
  277. Azevedo, C.; Teku, G.; Pomeshchik, Y.; Reyes, J.F.; Chumarina, M.; Russ, K.; Savchenko, E.; Hammarberg, A.; Lamas, N.J.; Collin, A.; et al. Parkinson’s disease and multiple system atrophy patient iPSC-derived oligodendrocytes exhibit alpha-synuclein-induced changes in maturation and immune reactive properties. Proc. Natl. Acad. Sci. USA 2022, 119, e2111405119. [Google Scholar] [CrossRef]
  278. Fleming, S.M.; Salcedo, J.; Hutson, C.B.; Rockenstein, E.; Masliah, E.; Levine, M.S.; Chesselet, M.F. Behavioral effects of dopaminergic agonists in transgenic mice overexpressing human wildtype α-synuclein. Neuroscience 2006, 142, 1245–1253. [Google Scholar] [CrossRef] [Green Version]
  279. Wakamatsu, M.; Ishii, A.; Iwata, S.; Sakagami, J.; Ukai, Y.; Ono, M.; Kanbe, D.; Muramatsu, S.; Kobayashi, K.; Iwatsubo, T.; et al. Selective loss of nigral dopamine neurons induced by overexpression of truncated human α-synuclein in mice. Neurobiol. Aging 2008, 29, 574–585. [Google Scholar] [CrossRef]
  280. Konnova, E.A.; Swanberg, M. Animal Models of Parkinson’s Disease. In Parkinson’s Disease: Pathogenesis and Clinical Aspects; Stoker, T.B., Greenland, J.C., Eds.; Codon Publications: Brisbane, Australia, 2018. [Google Scholar]
  281. Chesselet, M.F.; Fleming, S.; Mortazavi, F.; Meurers, B. Strengths and limitations of genetic mouse models of Parkinson’s disease. Park. Relat. Disord. 2008, 14 (Suppl. S2), S84–S87. [Google Scholar] [CrossRef]
  282. Chesselet, M.F.; Richter, F.; Zhu, C.; Magen, I.; Watson, M.B.; Subramaniam, S.R. A progressive mouse model of Parkinson’s disease: The Thy1-aSyn (“Line 61”) mice. Neurotherapeutics 2012, 9, 297–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Cannon, J.R.; Geghman, K.D.; Tapias, V.; Sew, T.; Dail, M.K.; Li, C.; Greenamyre, J.T. Expression of human E46K-mutated α-synuclein in BAC-transgenic rats replicates early-stage Parkinson’s disease features and enhances vulnerability to mitochondrial impairment. Exp. Neurol. 2013, 240, 44–56. [Google Scholar] [CrossRef] [Green Version]
  284. Decressac, M.; Mattsson, B.; Lundblad, M.; Weikop, P.; Bjorklund, A. Progressive neurodegenerative and behavioural changes induced by AAV-mediated overexpression of α-synuclein in midbrain dopamine neurons. Neurobiol. Dis. 2012, 45, 939–953. [Google Scholar] [CrossRef] [PubMed]
  285. Van der Perren, A.; Toelen, J.; Casteels, C.; Macchi, F.; Van Rompuy, A.S.; Sarre, S.; Casadei, N.; Nuber, S.; Himmelreich, U.; Osorio Garcia, M.I.; et al. Longitudinal follow-up and characterization of a robust rat model for Parkinson’s disease based on overexpression of alpha-synuclein with adeno-associated viral vectors. Neurobiol. Aging 2015, 36, 1543–1558. [Google Scholar] [CrossRef] [Green Version]
  286. Oliveras-Salva, M.; van der Perren, A.; Casadei, N.; Stroobants, S.; Nuber, S.; D’Hooge, R.; van den Haute, C.; Baekelandt, V. rAAV2/7 vector-mediated overexpression of alpha-synuclein in mouse substantia nigra induces protein aggregation and progressive dose-dependent neurodegeneration. Mol. Neurodegener. 2013, 8, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Kirik, D.; Annett, L.E.; Burger, C.; Muzyczka, N.; Mandel, R.J.; Bjorklund, A. Nigrostriatal α-synucleinopathy induced by viral vector-mediated overexpression of human α-synuclein: A new primate model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2003, 100, 2884–2889. [Google Scholar] [CrossRef] [Green Version]
  288. Koprich, J.B.; Johnston, T.H.; Reyes, G.; Omana, V.; Brotchie, J.M. Towards a Non-Human Primate Model of Alpha-Synucleinopathy for Development of Therapeutics for Parkinson’s Disease: Optimization of AAV1/2 Delivery Parameters to Drive Sustained Expression of Alpha Synuclein and Dopaminergic Degeneration in Macaque. PLoS ONE 2016, 11, e0167235. [Google Scholar] [CrossRef] [Green Version]
  289. Aharon-Peretz, J.; Rosenbaum, H.; Gershoni-Baruch, R. Mutations in the glucocerebrosidase gene and Parkinson’s disease in Ashkenazi Jews. N. Engl. J. Med. 2004, 351, 1972–1977. [Google Scholar] [CrossRef]
  290. Alcalay, R.N.; Dinur, T.; Quinn, T.; Sakanaka, K.; Levy, O.; Waters, C.; Fahn, S.; Dorovski, T.; Chung, W.K.; Pauciulo, M.; et al. Comparison of Parkinson risk in Ashkenazi Jewish patients with Gaucher disease and GBA heterozygotes. JAMA Neurol. 2014, 71, 752–757. [Google Scholar] [CrossRef] [Green Version]
  291. Sidransky, E.; Samaddar, T.; Tayebi, N. Mutations in GBA are associated with familial Parkinson disease susceptibility and age at onset. Neurology 2009, 73, 1424–1425. [Google Scholar] [CrossRef]
  292. Gan-Or, Z.; Giladi, N.; Rozovski, U.; Shifrin, C.; Rosner, S.; Gurevich, T.; Bar-Shira, A.; Orr-Urtreger, A. Genotype-phenotype correlations between GBA mutations and Parkinson disease risk and onset. Neurology 2008, 70, 2277–2283. [Google Scholar] [CrossRef]
  293. Lesage, S.; Condroyer, C.; Hecham, N.; Anheim, M.; Belarbi, S.; Lohman, E.; Viallet, F.; Pollak, P.; Abada, M.; Durr, A.; et al. Mutations in the glucocerebrosidase gene confer a risk for Parkinson disease in North Africa. Neurology 2011, 76, 301–303. [Google Scholar] [CrossRef] [Green Version]
  294. Rosenbloom, B.; Balwani, M.; Bronstein, J.M.; Kolodny, E.; Sathe, S.; Gwosdow, A.R.; Taylor, J.S.; Cole, J.A.; Zimran, A.; Weinreb, N.J. The incidence of Parkinsonism in patients with type 1 Gaucher disease: Data from the ICGG Gaucher Registry. Blood Cells Mol. Dis. 2011, 46, 95–102. [Google Scholar] [CrossRef] [Green Version]
  295. Hruska, K.S.; LaMarca, M.E.; Scott, C.R.; Sidransky, E. Gaucher disease: Mutation and polymorphism spectrum in the glucocerebrosidase gene (GBA). Hum. Mutat. 2008, 29, 567–583. [Google Scholar] [CrossRef]
  296. Gan-Or, Z.; Amshalom, I.; Kilarski, L.L.; Bar-Shira, A.; Gana-Weisz, M.; Mirelman, A.; Marder, K.; Bressman, S.; Giladi, N.; Orr-Urtreger, A. Differential effects of severe vs. mild GBA mutations on Parkinson disease. Neurology 2015, 84, 880–887. [Google Scholar] [CrossRef] [Green Version]
  297. Cilia, R.; Tunesi, S.; Marotta, G.; Cereda, E.; Siri, C.; Tesei, S.; Zecchinelli, A.L.; Canesi, M.; Mariani, C.B.; Meucci, N.; et al. Survival and dementia in GBA-associated Parkinson’s disease: The mutation matters. Ann. Neurol. 2016, 80, 662–673. [Google Scholar] [CrossRef]
  298. Blandini, F.; Cilia, R.; Cerri, S.; Pezzoli, G.; Schapira, A.H.V.; Mullin, S.; Lanciego, J.L. Glucocerebrosidase mutations and synucleinopathies: Toward a model of precision medicine. Mov. Disord. 2019, 34, 9–21. [Google Scholar] [CrossRef]
  299. Thaler, A.; Gurevich, T.; Bar Shira, A.; Gana Weisz, M.; Ash, E.; Shiner, T.; Orr-Urtreger, A.; Giladi, N.; Mirelman, A. A “dose” effect of mutations in the GBA gene on Parkinson’s disease phenotype. Park. Relat. Disord. 2017, 36, 47–51. [Google Scholar] [CrossRef]
  300. Corti, O.; Lesage, S.; Brice, A. What genetics tells us about the causes and mechanisms of Parkinson’s disease. Physiol. Rev. 2011, 91, 1161–1218. [Google Scholar] [CrossRef]
  301. Yap, T.L.; Gruschus, J.M.; Velayati, A.; Westbroek, W.; Goldin, E.; Moaven, N.; Sidransky, E.; Lee, J.C. α-synuclein interacts with Glucocerebrosidase providing a molecular link between Parkinson and Gaucher diseases. J. Biol. Chem. 2011, 286, 28080–28088. [Google Scholar] [CrossRef]
  302. Mazzulli, J.R.; Xu, Y.H.; Sun, Y.; Knight, A.L.; McLean, P.J.; Caldwell, G.A.; Sidransky, E.; Grabowski, G.A.; Krainc, D. Gaucher disease glucocerebrosidase and α-synuclein form a bidirectional pathogenic loop in synucleinopathies. Cell 2011, 146, 37–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Murphy, K.E.; Gysbers, A.M.; Abbott, S.K.; Tayebi, N.; Kim, W.S.; Sidransky, E.; Cooper, A.; Garner, B.; Halliday, G.M. Reduced glucocerebrosidase is associated with increased α-synuclein in sporadic Parkinson’s disease. Brain 2014, 137, 834–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Chiasserini, D.; Paciotti, S.; Eusebi, P.; Persichetti, E.; Tasegian, A.; Kurzawa-Akanbi, M.; Chinnery, P.F.; Morris, C.M.; Calabresi, P.; Parnetti, L.; et al. Selective loss of glucocerebrosidase activity in sporadic Parkinson’s disease and dementia with Lewy bodies. Mol. Neurodegener. 2015, 10, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Zunke, F.; Moise, A.C.; Belur, N.R.; Gelyana, E.; Stojkovska, I.; Dzaferbegovic, H.; Toker, N.J.; Jeon, S.; Fredriksen, K.; Mazzulli, J.R. Reversible Conformational Conversion of α-Synuclein into Toxic Assemblies by Glucosylceramide. Neuron 2018, 97, 92–107.e110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Suzuki, M.; Sango, K.; Wada, K.; Nagai, Y. Pathological role of lipid interaction with α-synuclein in Parkinson’s disease. Neurochem. Int. 2018, 119, 97–106. [Google Scholar] [CrossRef]
  307. Burbulla, L.F.; Jeon, S.; Zheng, J.; Song, P.; Silverman, R.B.; Krainc, D. A modulator of wild-type glucocerebrosidase improves pathogenic phenotypes in dopaminergic neuronal models of Parkinson’s disease. Sci. Transl. Med. 2019, 11, eaau6870. [Google Scholar] [CrossRef]
  308. Mazzulli, J.R.; Zunke, F.; Tsunemi, T.; Toker, N.J.; Jeon, S.; Burbulla, L.F.; Patnaik, S.; Sidransky, E.; Marugan, J.J.; Sue, C.M.; et al. Activation of β-Glucocerebrosidase Reduces Pathological α-Synuclein and Restores Lysosomal Function in Parkinson’s Patient Midbrain Neurons. J. Neurosci. 2016, 36, 7693–7706. [Google Scholar] [CrossRef]
  309. Burbulla, L.F.; Song, P.; Mazzulli, J.R.; Zampese, E.; Wong, Y.C.; Jeon, S.; Santos, D.P.; Blanz, J.; Obermaier, C.D.; Strojny, C.; et al. Dopamine oxidation mediates mitochondrial and lysosomal dysfunction in Parkinson’s disease. Science 2017, 357, 1255–1261. [Google Scholar] [CrossRef] [Green Version]
  310. Schondorf, D.C.; Aureli, M.; McAllister, F.E.; Hindley, C.J.; Mayer, F.; Schmid, B.; Sardi, S.P.; Valsecchi, M.; Hoffmann, S.; Schwarz, L.K.; et al. iPSC-derived neurons from GBA1-associated Parkinson’s disease patients show autophagic defects and impaired calcium homeostasis. Nat. Commun. 2014, 5, 4028. [Google Scholar] [CrossRef] [Green Version]
  311. Fernandes, H.J.; Hartfield, E.M.; Christian, H.C.; Emmanoulidou, E.; Zheng, Y.; Booth, H.; Bogetofte, H.; Lang, C.; Ryan, B.J.; Sardi, S.P.; et al. ER Stress and Autophagic Perturbations Lead to Elevated Extracellular α-Synuclein in GBA-N370S Parkinson’s iPSC-Derived Dopamine Neurons. Stem. Cell. Rep. 2016, 6, 342–356. [Google Scholar] [CrossRef]
  312. Schondorf, D.C.; Ivanyuk, D.; Baden, P.; Sanchez-Martinez, A.; de Cicco, S.; Yu, C.; Giunta, I.; Schwarz, L.K.; Di Napoli, G.; Panagiotakopoulou, V.; et al. The NAD+ Precursor Nicotinamide Riboside Rescues Mitochondrial Defects and Neuronal Loss in iPSC and Fly Models of Parkinson’s Disease. Cell. Rep. 2018, 23, 2976–2988. [Google Scholar] [CrossRef]
  313. Sanyal, A.; DeAndrade, M.P.; Novis, H.S.; Lin, S.; Chang, J.; Lengacher, N.; Tomlinson, J.J.; Tansey, M.G.; LaVoie, M.J. Lysosome and Inflammatory Defects in GBA1-Mutant Astrocytes Are Normalized by LRRK2 Inhibition. Mov. Disord. 2020, 35, 760–773. [Google Scholar] [CrossRef]
  314. Sanyal, A.; Novis, H.S.; Gasser, E.; Lin, S.; LaVoie, M.J. LRRK2 Kinase Inhibition Rescues Deficits in Lysosome Function Due to Heterozygous GBA1 Expression in Human iPSC-Derived Neurons. Front. Neurosci. 2020, 14, 442. [Google Scholar] [CrossRef]
  315. Kim, S.; Wong, Y.C.; Gao, F.; Krainc, D. Dysregulation of mitochondria-lysosome contacts by GBA1 dysfunction in dopaminergic neuronal models of Parkinson’s disease. Nat. Commun. 2021, 12, 1807. [Google Scholar] [CrossRef]
  316. Ysselstein, D.; Nguyen, M.; Young, T.J.; Severino, A.; Schwake, M.; Merchant, K.; Krainc, D. LRRK2 kinase activity regulates lysosomal glucocerebrosidase in neurons derived from Parkinson’s disease patients. Nat. Commun. 2019, 10, 5570. [Google Scholar] [CrossRef] [Green Version]
  317. Liu, Y.; Suzuki, K.; Reed, J.D.; Grinberg, A.; Westphal, H.; Hoffmann, A.; Doring, T.; Sandhoff, K.; Proia, R.L. Mice with type 2 and 3 Gaucher disease point mutations generated by a single insertion mutagenesis procedure. Proc. Natl. Acad. Sci. USA 1998, 95, 2503–2508. [Google Scholar] [CrossRef] [Green Version]
  318. Mizukami, H.; Mi, Y.; Wada, R.; Kono, M.; Yamashita, T.; Liu, Y.; Werth, N.; Sandhoff, R.; Sandhoff, K.; Proia, R.L. Systemic inflammation in glucocerebrosidase-deficient mice with minimal glucosylceramide storage. J. Clin. Invest. 2002, 109, 1215–1221. [Google Scholar] [CrossRef]
  319. Ginns, E.I.; Mak, S.K.; Ko, N.; Karlgren, J.; Akbarian, S.; Chou, V.P.; Guo, Y.; Lim, A.; Samuelsson, S.; LaMarca, M.L.; et al. Neuroinflammation and α-synuclein accumulation in response to glucocerebrosidase deficiency are accompanied by synaptic dysfunction. Mol. Genet. Metab. 2014, 111, 152–162. [Google Scholar] [CrossRef]
  320. Li, H.; Ham, A.; Ma, T.C.; Kuo, S.H.; Kanter, E.; Kim, D.; Ko, H.S.; Quan, Y.; Sardi, S.P.; Li, A.; et al. Mitochondrial dysfunction and mitophagy defect triggered by heterozygous GBA mutations. Autophagy 2019, 15, 113–130. [Google Scholar] [CrossRef] [Green Version]
  321. Yun, S.P.; Kim, D.; Kim, S.; Kim, S.; Karuppagounder, S.S.; Kwon, S.H.; Lee, S.; Kam, T.I.; Lee, S.; Ham, S.; et al. α-Synuclein accumulation and GBA deficiency due to L444P GBA mutation contributes to MPTP-induced parkinsonism. Mol. Neurodegener. 2018, 13, 1. [Google Scholar] [CrossRef]
  322. Migdalska-Richards, A.; Wegrzynowicz, M.; Rusconi, R.; Deangeli, G.; Di Monte, D.A.; Spillantini, M.G.; Schapira, A.H.V. The L444P Gba1 mutation enhances alpha-synuclein induced loss of nigral dopaminergic neurons in mice. Brain 2017, 140, 2706–2721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Taguchi, Y.V.; Liu, J.; Ruan, J.; Pacheco, J.; Zhang, X.; Abbasi, J.; Keutzer, J.; Mistry, P.K.; Chandra, S.S. Glucosylsphingosine Promotes α-Synuclein Pathology in Mutant GBA-Associated Parkinson’s Disease. J. Neurosci. 2017, 37, 9617–9631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Zhang, Y.; Shu, L.; Sun, Q.; Zhou, X.; Pan, H.; Guo, J.; Tang, B. Integrated Genetic Analysis of Racial Differences of Common GBA Variants in Parkinson’s Disease: A Meta-Analysis. Front. Mol. Neurosci. 2018, 11, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Xu, Y.H.; Sun, Y.; Ran, H.; Quinn, B.; Witte, D.; Grabowski, G.A. Accumulation and distribution of α-synuclein and ubiquitin in the CNS of Gaucher disease mouse models. Mol. Genet. Metab. 2011, 102, 436–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Kim, D.; Hwang, H.; Choi, S.; Kwon, S.H.; Lee, S.; Park, J.H.; Kim, S.; Ko, H.S. D409H GBA1 mutation accelerates the progression of pathology in A53T α-synuclein transgenic mouse model. Acta Neuropathol. Commun. 2018, 6, 32. [Google Scholar] [CrossRef] [PubMed]
  327. Farfel-Becker, T.; Vitner, E.B.; Pressey, S.N.; Eilam, R.; Cooper, J.D.; Futerman, A.H. Spatial and temporal correlation between neuron loss and neuroinflammation in a mouse model of neuronopathic Gaucher disease. Hum. Mol. Genet. 2011, 20, 1375–1386. [Google Scholar] [CrossRef]
  328. Fishbein, I.; Kuo, Y.M.; Giasson, B.I.; Nussbaum, R.L. Augmentation of phenotype in a transgenic Parkinson mouse heterozygous for a Gaucher mutation. Brain 2014, 137, 3235–3247. [Google Scholar] [CrossRef] [Green Version]
  329. Silvestri, L.; Caputo, V.; Bellacchio, E.; Atorino, L.; Dallapiccola, B.; Valente, E.M.; Casari, G. Mitochondrial import and enzymatic activity of PINK1 mutants associated to recessive parkinsonism. Hum. Mol. Genet. 2005, 14, 3477–3492. [Google Scholar] [CrossRef] [Green Version]
  330. Yamano, K.; Youle, R.J. PINK1 is degraded through the N-end rule pathway. Autophagy 2013, 9, 1758–1769. [Google Scholar] [CrossRef] [Green Version]
  331. Kondapalli, C.; Kazlauskaite, A.; Zhang, N.; Woodroof, H.I.; Campbell, D.G.; Gourlay, R.; Burchell, L.; Walden, H.; Macartney, T.J.; Deak, M.; et al. PINK1 is activated by mitochondrial membrane potential depolarization and stimulates Parkin E3 ligase activity by phosphorylating Serine 65. Open Biol. 2012, 2, 120080. [Google Scholar] [CrossRef]
  332. Shiba-Fukushima, K.; Imai, Y.; Yoshida, S.; Ishihama, Y.; Kanao, T.; Sato, S.; Hattori, N. PINK1-mediated phosphorylation of the Parkin ubiquitin-like domain primes mitochondrial translocation of Parkin and regulates mitophagy. Sci. Rep. 2012, 2, 1002. [Google Scholar] [CrossRef] [Green Version]
  333. Sarraf, S.A.; Raman, M.; Guarani-Pereira, V.; Sowa, M.E.; Huttlin, E.L.; Gygi, S.P.; Harper, J.W. Landscape of the PARKIN-dependent ubiquitylome in response to mitochondrial depolarization. Nature 2013, 496, 372–376. [Google Scholar] [CrossRef] [Green Version]
  334. Wauer, T.; Simicek, M.; Schubert, A.; Komander, D. Mechanism of phospho-ubiquitin-induced PARKIN activation. Nature 2015, 524, 370–374. [Google Scholar] [CrossRef] [Green Version]
  335. Tanaka, A.; Cleland, M.M.; Xu, S.; Narendra, D.P.; Suen, D.F.; Karbowski, M.; Youle, R.J. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J. Cell. Biol. 2010, 191, 1367–1380. [Google Scholar] [CrossRef] [Green Version]
  336. 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] [Green Version]
  337. Plun-Favreau, H.; Klupsch, K.; Moisoi, N.; Gandhi, S.; Kjaer, S.; Frith, D.; Harvey, K.; Deas, E.; Harvey, R.J.; McDonald, N.; et al. The mitochondrial protease HtrA2 is regulated by Parkinson’s disease-associated kinase PINK1. Nat. Cell. Biol. 2007, 9, 1243–1252. [Google Scholar] [CrossRef]
  338. Pridgeon, J.W.; Olzmann, J.A.; Chin, L.S.; Li, L. PINK1 protects against oxidative stress by phosphorylating mitochondrial chaperone TRAP1. PLoS Biol. 2007, 5, e172. [Google Scholar] [CrossRef]
  339. Kim, Y.; Park, J.; Kim, S.; Song, S.; Kwon, S.K.; Lee, S.H.; Kitada, T.; Kim, J.M.; Chung, J. PINK1 controls mitochondrial localization of Parkin through direct phosphorylation. Biochem. Biophys. Res. Commun. 2008, 377, 975–980. [Google Scholar] [CrossRef]
  340. Shimura, H.; Schlossmacher, M.G.; Hattori, N.; Frosch, M.P.; Trockenbacher, A.; Schneider, R.; Mizuno, Y.; Kosik, K.S.; Selkoe, D.J. Ubiquitination of a new form of α-synuclein by parkin from human brain: Implications for Parkinson’s disease. Science 2001, 293, 263–269. [Google Scholar] [CrossRef]
  341. Beilina, A.; van der Brug, M.; Ahmad, R.; Kesavapany, S.; Miller, D.W.; Petsko, G.A.; Cookson, M.R. Mutations in PTEN-induced putative kinase 1 associated with recessive parkinsonism have differential effects on protein stability. Proc. Natl. Acad. Sci. USA 2005, 102, 5703–5708. [Google Scholar] [CrossRef]
  342. Steinlechner, S.; Stahlberg, J.; Volkel, B.; Djarmati, A.; Hagenah, J.; Hiller, A.; Hedrich, K.; Konig, I.; Klein, C.; Lencer, R. Co-occurrence of affective and schizophrenia spectrum disorders with PINK1 mutations. J. Neurol. Neurosurg. Psychiatry 2007, 78, 532–535. [Google Scholar] [CrossRef] [PubMed]
  343. Samaranch, L.; Lorenzo-Betancor, O.; Arbelo, J.M.; Ferrer, I.; Lorenzo, E.; Irigoyen, J.; Pastor, M.A.; Marrero, C.; Isla, C.; Herrera-Henriquez, J.; et al. PINK1-linked parkinsonism is associated with Lewy body pathology. Brain 2010, 133, 1128–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Puschmann, A.; Fiesel, F.C.; Caulfield, T.R.; Hudec, R.; Ando, M.; Truban, D.; Hou, X.; Ogaki, K.; Heckman, M.G.; James, E.D.; et al. Heterozygous PINK1 p.G411S increases risk of Parkinson’s disease via a dominant-negative mechanism. Brain 2017, 140, 98–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Morais, V.A.; Haddad, D.; Craessaerts, K.; De Bock, P.J.; Swerts, J.; Vilain, S.; Aerts, L.; Overbergh, L.; Grunewald, A.; Seibler, P.; et al. PINK1 loss-of-function mutations affect mitochondrial complex I activity via NdufA10 ubiquinone uncoupling. Science 2014, 344, 203–207. [Google Scholar] [CrossRef] [PubMed]
  346. Rakovic, A.; Shurkewitsch, K.; Seibler, P.; Grunewald, A.; Zanon, A.; Hagenah, J.; Krainc, D.; Klein, C. Phosphatase and tensin homolog (PTEN)-induced putative kinase 1 (PINK1)-dependent ubiquitination of endogenous Parkin attenuates mitophagy: Study in human primary fibroblasts and induced pluripotent stem cell-derived neurons. J. Biol. Chem. 2013, 288, 2223–2237. [Google Scholar] [CrossRef] [Green Version]
  347. Seibler, P.; Graziotto, J.; Jeong, H.; Simunovic, F.; Klein, C.; Krainc, D. Mitochondrial Parkin recruitment is impaired in neurons derived from mutant PINK1 induced pluripotent stem cells. J. Neurosci. 2011, 31, 5970–5976. [Google Scholar] [CrossRef] [Green Version]
  348. Vos, M.; Geens, A.; Bohm, C.; Deaulmerie, L.; Swerts, J.; Rossi, M.; Craessaerts, K.; Leites, E.P.; Seibler, P.; Rakovic, A.; et al. Cardiolipin promotes electron transport between ubiquinone and complex I to rescue PINK1 deficiency. J. Cell. Biol. 2017, 216, 695–708. [Google Scholar] [CrossRef] [Green Version]
  349. Cooper, O.; Seo, H.; Andrabi, S.; Guardia-Laguarta, C.; Graziotto, J.; Sundberg, M.; McLean, J.R.; Carrillo-Reid, L.; Xie, Z.; Osborn, T.; et al. Pharmacological rescue of mitochondrial deficits in iPSC-derived neural cells from patients with familial Parkinson’s disease. Sci. Transl. Med. 2012, 4, 1–13. [Google Scholar] [CrossRef] [Green Version]
  350. Chung, S.Y.; Kishinevsky, S.; Mazzulli, J.R.; Graziotto, J.; Mrejeru, A.; Mosharov, E.V.; Puspita, L.; Valiulahi, P.; Sulzer, D.; Milner, T.A.; et al. Parkin and PINK1 Patient iPSC-Derived Midbrain Dopamine Neurons Exhibit Mitochondrial Dysfunction and α-Synuclein Accumulation. Stem. Cell. Rep. 2016, 7, 664–677. [Google Scholar] [CrossRef] [Green Version]
  351. Azkona, G.; Lopez de Maturana, R.; del Rio, P.; Sousa, A.; Vazquez, N.; Zubiarrain, A.; Jimenez-Blasco, D.; Bolanos, J.P.; Morales, B.; Auburger, G.; et al. LRRK2 Expression Is Deregulated in Fibroblasts and Neurons from Parkinson Patients with Mutations in PINK1. Mol. Neurobiol. 2018, 55, 506–516. [Google Scholar] [CrossRef]
  352. Gautier, C.A.; Kitada, T.; Shen, J. Loss of PINK1 causes mitochondrial functional defects and increased sensitivity to oxidative stress. Proc. Natl. Acad. Sci. USA 2008, 105, 11364–11369. [Google Scholar] [CrossRef] [Green Version]
  353. Gispert, S.; Ricciardi, F.; Kurz, A.; Azizov, M.; Hoepken, H.H.; Becker, D.; Voos, W.; Leuner, K.; Muller, W.E.; Kudin, A.P.; et al. Parkinson phenotype in aged PINK1-deficient mice is accompanied by progressive mitochondrial dysfunction in absence of neurodegeneration. PLoS ONE 2009, 4, e5777. [Google Scholar] [CrossRef] [Green Version]
  354. Glasl, L.; Kloos, K.; Giesert, F.; Roethig, A.; di Benedetto, B.; Kuhn, R.; Zhang, J.; Hafen, U.; Zerle, J.; Hofmann, A.; et al. Pink1-deficiency in mice impairs gait, olfaction and serotonergic innervation of the olfactory bulb. Exp. Neurol. 2012, 235, 214–227. [Google Scholar] [CrossRef]
  355. Kitada, T.; Pisani, A.; Porter, D.R.; Yamaguchi, H.; Tscherter, A.; Martella, G.; Bonsi, P.; Zhang, C.; Pothos, E.N.; Shen, J. Impaired dopamine release and synaptic plasticity in the striatum of PINK1-deficient mice. Proc. Natl. Acad. Sci. USA 2007, 104, 11441–11446. [Google Scholar] [CrossRef] [Green Version]
  356. Morais, V.A.; Verstreken, P.; Roethig, A.; Smet, J.; Snellinx, A.; Vanbrabant, M.; Haddad, D.; Frezza, C.; Mandemakers, W.; Vogt-Weisenhorn, D.; et al. Parkinson’s disease mutations in PINK1 result in decreased Complex I activity and deficient synaptic function. EMBO Mol. Med. 2009, 1, 99–111. [Google Scholar] [CrossRef]
  357. Smith, G.A.; Isacson, O.; Dunnett, S.B. The search for genetic mouse models of prodromal Parkinson’s disease. Exp. Neurol. 2012, 237, 267–273. [Google Scholar] [CrossRef]
  358. Jiang, P.; Dickson, D.W. Parkinson’s disease: Experimental models and reality. Acta Neuropathol. 2018, 135, 13–32. [Google Scholar] [CrossRef]
  359. Akundi, R.S.; Huang, Z.; Eason, J.; Pandya, J.D.; Zhi, L.; Cass, W.A.; Sullivan, P.G.; Bueler, H. Increased mitochondrial calcium sensitivity and abnormal expression of innate immunity genes precede dopaminergic defects in Pink1-deficient mice. PLoS ONE 2011, 6, e16038. [Google Scholar] [CrossRef] [Green Version]
  360. Dave, K.D.; De Silva, S.; Sheth, N.P.; Ramboz, S.; Beck, M.J.; Quang, C.; Switzer, R.C., 3rd; Ahmad, S.O.; Sunkin, S.M.; Walker, D.; et al. Phenotypic characterization of recessive gene knockout rat models of Parkinson’s disease. Neurobiol. Dis. 2014, 70, 190–203. [Google Scholar] [CrossRef] [Green Version]
  361. Villeneuve, L.M.; Purnell, P.R.; Boska, M.D.; Fox, H.S. Early Expression of Parkinson’s Disease-Related Mitochondrial Abnormalities in PINK1 Knockout Rats. Mol. Neurobiol. 2016, 53, 171–186. [Google Scholar] [CrossRef]
  362. Grant, L.M.; Kelm-Nelson, C.A.; Hilby, B.L.; Blue, K.V.; Paul Rajamanickam, E.S.; Pultorak, J.D.; Fleming, S.M.; Ciucci, M.R. Evidence for early and progressive ultrasonic vocalization and oromotor deficits in a PINK1 gene knockout rat model of Parkinson’s disease. J. Neurosci. Res. 2015, 93, 1713–1727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. 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]
  364. Matsumine, H.; Saito, M.; Shimoda-Matsubayashi, S.; Tanaka, H.; Ishikawa, A.; Nakagawa-Hattori, Y.; Yokochi, M.; Kobayashi, T.; Igarashi, S.; Takano, H.; et al. Localization of a gene for an autosomal recessive form of juvenile Parkinsonism to chromosome 6q25.2-27. Am. J. Hum. Genet. 1997, 60, 588–596. [Google Scholar] [PubMed]
  365. Mata, I.F.; Lockhart, P.J.; Farrer, M.J. Parkin genetics: One model for Parkinson’s disease. Hum. Mol. Genet. 2004, 13, R127–R133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  366. 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]
  367. Charan, R.A.; LaVoie, M.J. Pathologic and therapeutic implications for the cell biology of parkin. Mol. Cell. Neurosci. 2015, 66, 62–71. [Google Scholar] [CrossRef] [Green Version]
  368. Conceicao, I.C.; Rama, M.M.; Oliveira, B.; Cafe, C.; Almeida, J.; Mouga, S.; Duque, F.; Oliveira, G.; Vicente, A.M. Definition of a putative pathological region in PARK2 associated with autism spectrum disorder through in silico analysis of its functional structure. Psychiatr. Genet. 2017, 27, 54–61. [Google Scholar] [CrossRef]
  369. Lucking, C.B.; Durr, A.; Bonifati, V.; Vaughan, J.; de Michele, G.; Gasser, T.; Harhangi, B.S.; Meco, G.; Denefle, P.; Wood, N.W.; et al. Association between early-onset Parkinson’s disease and mutations in the parkin gene. N. Engl. J. Med. 2000, 342, 1560–1567. [Google Scholar] [CrossRef]
  370. Kann, M.; Jacobs, H.; Mohrmann, K.; Schumacher, K.; Hedrich, K.; Garrels, J.; Wiegers, K.; Schwinger, E.; Pramstaller, P.P.; Breakefield, X.O.; et al. Role of parkin mutations in 111 community-based patients with early-onset parkinsonism. Ann. Neurol. 2002, 51, 621–625. [Google Scholar] [CrossRef]
  371. Foroud, T.; Uniacke, S.K.; Liu, L.; Pankratz, N.; Rudolph, A.; Halter, C.; Shults, C.; Marder, K.; Conneally, P.M.; Nichols, W.C.; et al. Heterozygosity for a mutation in the parkin gene leads to later onset Parkinson disease. Neurology 2003, 60, 796–801. [Google Scholar] [CrossRef]
  372. Periquet, M.; Latouche, M.; Lohmann, E.; Rawal, N.; de Michele, G.; Ricard, S.; Teive, H.; Fraix, V.; Vidailhet, M.; Nicholl, D.; et al. Parkin mutations are frequent in patients with isolated early-onset parkinsonism. Brain 2003, 126, 1271–1278. [Google Scholar] [CrossRef] [Green Version]
  373. Kachidian, P.; Gubellini, P. Genetic Models of Parkinson’s Disease. In Clinical Trials In Parkinson’s Disease; Perez-Lloret, S., Ed.; Springer: New York, NY, USA, 2021; pp. 37–84. [Google Scholar]
  374. Abbas, N.; Lucking, C.B.; Ricard, S.; Durr, A.; Bonifati, V.; de Michele, G.; Bouley, S.; Vaughan, J.R.; Gasser, T.; Marconi, R.; et al. A wide variety of mutations in the parkin gene are responsible for autosomal recessive parkinsonism in Europe. French Parkinson’s Disease Genetics Study Group and the European Consortium on Genetic Susceptibility in Parkinson’s Disease. Hum. Mol. Genet. 1999, 8, 567–574. [Google Scholar] [CrossRef]
  375. Lohmann, E.; Periquet, M.; Bonifati, V.; Wood, N.W.; de Michele, G.; Bonnet, A.M.; Fraix, V.; Broussolle, E.; Horstink, M.W.; Vidailhet, M.; et al. How much phenotypic variation can be attributed to parkin genotype? Ann. Neurol. 2003, 54, 176–185. [Google Scholar] [CrossRef]
  376. Benbunan, B.R.; Korczyn, A.D.; Giladi, N. Parkin mutation associated parkinsonism and cognitive decline, comparison to early onset Parkinson’s disease. J. Neural Transm. 2004, 111, 47–57. [Google Scholar] [CrossRef]
  377. Srivastava, A.; Tang, M.X.; Mejia-Santana, H.; Rosado, L.; Louis, E.D.; Caccappolo, E.; Comella, C.; Colcher, A.; Siderowf, A.; Jennings, D.; et al. The relation between depression and parkin genotype: The CORE-PD study. Park. Relat. Disord. 2011, 17, 740–744. [Google Scholar] [CrossRef] [Green Version]
  378. Farrer, M.; Chan, P.; Chen, R.; Tan, L.; Lincoln, S.; Hernandez, D.; Forno, L.; Gwinn-Hardy, K.; Petrucelli, L.; Hussey, J.; et al. Lewy bodies and parkinsonism in families with parkin mutations. Ann. Neurol. 2001, 50, 293–300. [Google Scholar] [CrossRef]
  379. Oczkowska, A.; Kozubski, W.; Lianeri, M.; Dorszewska, J. Mutations in PRKN and SNCA Genes Important for the Progress of Parkinson’s Disease. Curr. Genom. 2013, 14, 502–517. [Google Scholar] [CrossRef]
  380. Imaizumi, Y.; Okada, Y.; Akamatsu, W.; Koike, M.; Kuzumaki, N.; Hayakawa, H.; Nihira, T.; Kobayashi, T.; Ohyama, M.; Sato, S.; et al. Mitochondrial dysfunction associated with increased oxidative stress and α-synuclein accumulation in PARK2 iPSC-derived neurons and postmortem brain tissue. Mol. Brain 2012, 5, 35. [Google Scholar] [CrossRef] [Green Version]
  381. Bogetofte, H.; Jensen, P.; Ryding, M.; Schmidt, S.I.; Okarmus, J.; Ritter, L.; Worm, C.S.; Hohnholt, M.C.; Azevedo, C.; Roybon, L.; et al. PARK2 Mutation Causes Metabolic Disturbances and Impaired Survival of Human iPSC-Derived Neurons. Front. Cell. Neurosci. 2019, 13, 297. [Google Scholar] [CrossRef] [Green Version]
  382. Shaltouki, A.; Sivapatham, R.; Pei, Y.; Gerencser, A.A.; Momcilovic, O.; Rao, M.S.; Zeng, X. Mitochondrial alterations by PARKIN in dopaminergic neurons using PARK2 patient-specific and PARK2 knockout isogenic iPSC lines. Stem. Cell. Rep. 2015, 4, 847–859. [Google Scholar] [CrossRef]
  383. Jiang, H.; Ren, Y.; Yuen, E.Y.; Zhong, P.; Ghaedi, M.; Hu, Z.; Azabdaftari, G.; Nakaso, K.; Yan, Z.; Feng, J. Parkin controls dopamine utilization in human midbrain dopaminergic neurons derived from induced pluripotent stem cells. Nat. Commun. 2012, 3, 668. [Google Scholar] [CrossRef] [Green Version]
  384. Yokota, M.; Kakuta, S.; Shiga, T.; Ishikawa, K.I.; Okano, H.; Hattori, N.; Akamatsu, W.; Koike, M. Establishment of an in vitro model for analyzing mitochondrial ultrastructure in PRKN-mutated patient iPSC-derived dopaminergic neurons. Mol. Brain 2021, 14, 58. [Google Scholar] [CrossRef] [PubMed]
  385. Aboud, A.A.; Tidball, A.M.; Kumar, K.K.; Neely, M.D.; Han, B.; Ess, K.C.; Hong, C.C.; Erikson, K.M.; Hedera, P.; Bowman, A.B. PARK2 patient neuroprogenitors show increased mitochondrial sensitivity to copper. Neurobiol. Dis. 2015, 73, 204–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  386. Pickrell, A.M.; Youle, R.J. The roles of PINK1, parkin, and mitochondrial fidelity in Parkinson’s disease. Neuron 2015, 85, 257–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  387. Scarffe, L.A.; Stevens, D.A.; Dawson, V.L.; Dawson, T.M. Parkin and PINK1: Much more than mitophagy. Trends. Neurosci. 2014, 37, 315–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  388. Grenier, K.; McLelland, G.L.; Fon, E.A. Parkin- and PINK1-Dependent Mitophagy in Neurons: Will the Real Pathway Please Stand Up? Front. Neurol. 2013, 4, 100. [Google Scholar] [CrossRef] [Green Version]
  389. Sterky, F.H.; Lee, S.; Wibom, R.; Olson, L.; Larsson, N.G. Impaired mitochondrial transport and Parkin-independent degeneration of respiratory chain-deficient dopamine neurons in vivo. Proc. Natl. Acad. Sci. USA 2011, 108, 12937–12942. [Google Scholar] [CrossRef] [Green Version]
  390. Kageyama, Y.; Hoshijima, M.; Seo, K.; Bedja, D.; Sysa-Shah, P.; Andrabi, S.A.; Chen, W.; Hoke, A.; Dawson, V.L.; Dawson, T.M.; et al. Parkin-independent mitophagy requires Drp1 and maintains the integrity of mammalian heart and brain. EMBO J. 2014, 33, 2798–2813. [Google Scholar] [CrossRef] [Green Version]
  391. Zhong, P.; Hu, Z.; Jiang, H.; Yan, Z.; Feng, J. Dopamine Induces Oscillatory Activities in Human Midbrain Neurons with Parkin Mutations. Cell. Rep. 2017, 19, 1033–1044. [Google Scholar] [CrossRef] [Green Version]
  392. Brittain, J.S.; Brown, P. Oscillations and the basal ganglia: Motor control and beyond. Neuroimage 2014, 85, 637–647. [Google Scholar] [CrossRef]
  393. Ren, Y.; Jiang, H.; Hu, Z.; Fan, K.; Wang, J.; Janoschka, S.; Wang, X.; Ge, S.; Feng, J. Parkin mutations reduce the complexity of neuronal processes in iPSC-derived human neurons. Stem. Cells 2015, 33, 68–78. [Google Scholar] [CrossRef] [Green Version]
  394. Cartelli, D.; Amadeo, A.; Calogero, A.M.; Casagrande, F.V.M.; de Gregorio, C.; Gioria, M.; Kuzumaki, N.; Costa, I.; Sassone, J.; Ciammola, A.; et al. Parkin absence accelerates microtubule aging in dopaminergic neurons. Neurobiol. Aging 2018, 61, 66–74. [Google Scholar] [CrossRef]
  395. Vercammen, L.; van der Perren, A.; Vaudano, E.; Gijsbers, R.; Debyser, Z.; van den Haute, C.; Baekelandt, V. Parkin protects against neurotoxicity in the 6-hydroxydopamine rat model for Parkinson’s disease. Mol. Ther. 2006, 14, 716–723. [Google Scholar] [CrossRef]
  396. Yamada, M.; Mizuno, Y.; Mochizuki, H. Parkin gene therapy for α-synucleinopathy: A rat model of Parkinson’s disease. Hum. Gene Ther. 2005, 16, 262–270. [Google Scholar] [CrossRef]
  397. Lo Bianco, C.; Schneider, B.L.; Bauer, M.; Sajadi, A.; Brice, A.; Iwatsubo, T.; Aebischer, P. Lentiviral vector delivery of parkin prevents dopaminergic degeneration in an α-synuclein rat model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2004, 101, 17510–17515. [Google Scholar] [CrossRef] [Green Version]
  398. Manfredsson, F.P.; Burger, C.; Sullivan, L.F.; Muzyczka, N.; Lewin, A.S.; Mandel, R.J. rAAV-mediated nigral human parkin over-expression partially ameliorates motor deficits via enhanced dopamine neurotransmission in a rat model of Parkinson’s disease. Exp. Neurol. 2007, 207, 289–301. [Google Scholar] [CrossRef]
  399. Goldberg, M.S.; Fleming, S.M.; Palacino, J.J.; Cepeda, C.; Lam, H.A.; Bhatnagar, A.; Meloni, E.G.; Wu, N.; Ackerson, L.C.; Klapstein, G.J.; et al. Parkin-deficient mice exhibit nigrostriatal deficits but not loss of dopaminergic neurons. J. Biol. Chem. 2003, 278, 43628–43635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Hennis, M.R.; Seamans, K.W.; Marvin, M.A.; Casey, B.H.; Goldberg, M.S. Behavioral and neurotransmitter abnormalities in mice deficient for Parkin, DJ-1 and superoxide dismutase. PLoS ONE 2013, 8, e84894. [Google Scholar] [CrossRef] [Green Version]
  401. Hennis, M.R.; Marvin, M.A.; Taylor, C.M., 2nd; Goldberg, M.S. Surprising behavioral and neurochemical enhancements in mice with combined mutations linked to Parkinson’s disease. Neurobiol. Dis. 2014, 62, 113–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Palacino, J.J.; Sagi, D.; Goldberg, M.S.; Krauss, S.; Motz, C.; Wacker, M.; Klose, J.; Shen, J. Mitochondrial dysfunction and oxidative damage in parkin-deficient mice. J. Biol. Chem. 2004, 279, 18614–18622. [Google Scholar] [CrossRef] [PubMed]
  403. Perez, F.A.; Palmiter, R.D. Parkin-deficient mice are not a robust model of parkinsonism. Proc. Natl. Acad. Sci. USA 2005, 102, 2174–2179. [Google Scholar] [CrossRef] [Green Version]
  404. Von Coelln, R.; Thomas, B.; Savitt, J.M.; Lim, K.L.; Sasaki, M.; Hess, E.J.; Dawson, V.L.; Dawson, T.M. Loss of locus coeruleus neurons and reduced startle in parkin null mice. Proc. Natl. Acad. Sci. USA 2004, 101, 10744–10749. [Google Scholar] [CrossRef] [Green Version]
  405. Zhu, X.R.; Maskri, L.; Herold, C.; Bader, V.; Stichel, C.C.; Gunturkun, O.; Lubbert, H. Non-motor behavioural impairments in parkin-deficient mice. Eur. J. Neurosci. 2007, 26, 1902–1911. [Google Scholar] [CrossRef]
  406. Itier, J.M.; Ibanez, P.; Mena, M.A.; Abbas, N.; Cohen-Salmon, C.; Bohme, G.A.; Laville, M.; Pratt, J.; Corti, O.; Pradier, L.; et al. Parkin gene inactivation alters behaviour and dopamine neurotransmission in the mouse. Hum. Mol. Genet. 2003, 12, 2277–2291. [Google Scholar] [CrossRef] [Green Version]
  407. Kitada, T.; Tong, Y.; Gautier, C.A.; Shen, J. Absence of nigral degeneration in aged parkin/DJ-1/PINK1 triple knockout mice. J. Neurochem. 2009, 111, 696–702. [Google Scholar] [CrossRef] [Green Version]
  408. Doherty, K.M.; Hardy, J. Parkin disease and the Lewy body conundrum. Mov. Disord. 2013, 28, 702–704. [Google Scholar] [CrossRef] [Green Version]
  409. Lu, X.H.; Fleming, S.M.; Meurers, B.; Ackerson, L.C.; Mortazavi, F.; Lo, V.; Hernandez, D.; Sulzer, D.; Jackson, G.R.; Maidment, N.T.; et al. Bacterial artificial chromosome transgenic mice expressing a truncated mutant parkin exhibit age-dependent hypokinetic motor deficits, dopaminergic neuron degeneration, and accumulation of proteinase K-resistant α-synuclein. J. Neurosci. 2009, 29, 1962–1976. [Google Scholar] [CrossRef] [Green Version]
  410. Van Rompuy, A.S.; Lobbestael, E.; van der Perren, A.; van den Haute, C.; Baekelandt, V. Long-term overexpression of human wild-type and T240R mutant Parkin in rat substantia nigra induces progressive dopaminergic neurodegeneration. J. Neuropathol. Exp. Neurol. 2014, 73, 159–174. [Google Scholar] [CrossRef] [Green Version]
  411. Wilson, M.A. The role of cysteine oxidation in DJ-1 function and dysfunction. Antioxid. Redox. Signal 2011, 15, 111–122. [Google Scholar] [CrossRef] [Green Version]
  412. Bandopadhyay, R.; Kingsbury, A.E.; Cookson, M.R.; Reid, A.R.; Evans, I.M.; Hope, A.D.; Pittman, A.M.; Lashley, T.; Canet-Aviles, R.; Miller, D.W.; et al. The expression of DJ-1 (PARK7) in normal human CNS and idiopathic Parkinson’s disease. Brain 2004, 127, 420–430. [Google Scholar] [CrossRef]
  413. Rizzu, P.; Hinkle, D.A.; Zhukareva, V.; Bonifati, V.; Severijnen, L.A.; Martinez, D.; Ravid, R.; Kamphorst, W.; Eberwine, J.H.; Lee, V.M.; et al. DJ-1 colocalizes with tau inclusions: A link between parkinsonism and dementia. Ann. Neurol. 2004, 55, 113–118. [Google Scholar] [CrossRef] [PubMed]
  414. Oh, S.E.; Mouradian, M.M. Regulation of Signal Transduction by DJ-1. Adv. Exp. Med. Biol. 2017, 1037, 97–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Kahle, P.J.; Waak, J.; Gasser, T. DJ-1 and prevention of oxidative stress in Parkinson’s disease and other age-related disorders. Free Radic. Biol. Med. 2009, 47, 1354–1361. [Google Scholar] [CrossRef] [PubMed]
  416. Batelli, S.; Invernizzi, R.W.; Negro, A.; Calcagno, E.; Rodilossi, S.; Forloni, G.; Albani, D. The Parkinson’s disease-related protein DJ-1 protects dopaminergic neurons in vivo and cultured cells from ahpha-synuclein and 6-hydroxydopamine toxicity. Neurodegener. Dis. 2015, 15, 13–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  417. Kim, R.H.; Smith, P.D.; Aleyasin, H.; Hayley, S.; Mount, M.P.; Pownall, S.; Wakeham, A.; You-Ten, A.J.; Kalia, S.K.; Horne, P.; et al. Hypersensitivity of DJ-1-deficient mice to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyrindine (MPTP) and oxidative stress. Proc. Natl. Acad. Sci. USA 2005, 102, 5215–5220. [Google Scholar] [CrossRef] [Green Version]
  418. Meulener, M.; Whitworth, A.J.; Armstrong-Gold, C.E.; Rizzu, P.; Heutink, P.; Wes, P.D.; Pallanck, L.J.; Bonini, N.M. Drosophila DJ-1 mutants are selectively sensitive to environmental toxins associated with Parkinson’s disease. Curr. Biol. 2005, 15, 1572–1577. [Google Scholar] [CrossRef] [Green Version]
  419. Ottolini, D.; Cali, T.; Negro, A.; Brini, M. The Parkinson disease-related protein DJ-1 counteracts mitochondrial impairment induced by the tumour suppressor protein p53 by enhancing endoplasmic reticulum-mitochondria tethering. Hum. Mol. Genet. 2013, 22, 2152–2168. [Google Scholar] [CrossRef]
  420. Taira, T.; Saito, Y.; Niki, T.; Iguchi-Ariga, S.M.; Takahashi, K.; Ariga, H. DJ-1 has a role in antioxidative stress to prevent cell death. EMBO Rep. 2004, 5, 213–218. [Google Scholar] [CrossRef] [Green Version]
  421. Thomas, K.J.; McCoy, M.K.; Blackinton, J.; Beilina, A.; van der Brug, M.; Sandebring, A.; Miller, D.; Maric, D.; Cedazo-Minguez, A.; Cookson, M.R. DJ-1 acts in parallel to the PINK1/parkin pathway to control mitochondrial function and autophagy. Hum. Mol. Genet. 2011, 20, 40–50. [Google Scholar] [CrossRef] [Green Version]
  422. Kinumi, T.; Kimata, J.; Taira, T.; Ariga, H.; Niki, E. Cysteine-106 of DJ-1 is the most sensitive cysteine residue to hydrogen peroxide-mediated oxidation in vivo in human umbilical vein endothelial cells. Biochem. Biophys. Res. Commun. 2004, 317, 722–728. [Google Scholar] [CrossRef]
  423. Andres-Mateos, E.; Perier, C.; Zhang, L.; Blanchard-Fillion, B.; Greco, T.M.; Thomas, B.; Ko, H.S.; Sasaki, M.; Ischiropoulos, H.; Przedborski, S.; et al. DJ-1 gene deletion reveals that DJ-1 is an atypical peroxiredoxin-like peroxidase. Proc. Natl. Acad. Sci. USA 2007, 104, 14807–14812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  424. Smith, N.; Wilson, M.A. Structural Biology of the DJ-1 Superfamily. Adv. Exp. Med. Biol. 2017, 1037, 5–24. [Google Scholar] [CrossRef] [PubMed]
  425. Lev, N.; Barhum, Y.; Pilosof, N.S.; Ickowicz, D.; Cohen, H.Y.; Melamed, E.; Offen, D. DJ-1 protects against dopamine toxicity: Implications for Parkinson’s disease and aging. J. Gerontol. Biol. Sci. Med. Sci. 2013, 68, 215–225. [Google Scholar] [CrossRef] [PubMed]
  426. Shendelman, S.; Jonason, A.; Martinat, C.; Leete, T.; Abeliovich, A. DJ-1 is a redox-dependent molecular chaperone that inhibits α-synuclein aggregate formation. PLoS Biol. 2004, 2, e362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  427. Atieh, T.B.; Roth, J.; Yang, X.; Hoop, C.L.; Baum, J. DJ-1 Acts as a Scavenger of α-Synuclein Oligomers and Restores Monomeric Glycated α-Synuclein. Biomolecules 2021, 11, 1466. [Google Scholar] [CrossRef] [PubMed]
  428. Zondler, L.; Miller-Fleming, L.; Repici, M.; Goncalves, S.; Tenreiro, S.; Rosado-Ramos, R.; Betzer, C.; Straatman, K.R.; Jensen, P.H.; Giorgini, F.; et al. DJ-1 interactions with α-synuclein attenuate aggregation and cellular toxicity in models of Parkinson’s disease. Cell. Death Dis. 2014, 5, e1350. [Google Scholar] [CrossRef] [Green Version]
  429. Xu, C.Y.; Kang, W.Y.; Chen, Y.M.; Jiang, T.F.; Zhang, J.; Zhang, L.N.; Ding, J.Q.; Liu, J.; Chen, S.D. DJ-1 Inhibits α-Synuclein Aggregation by Regulating Chaperone-Mediated Autophagy. Front. Aging Neurosci. 2017, 9, 308. [Google Scholar] [CrossRef] [Green Version]
  430. Lee, J.Y.; Song, J.; Kwon, K.; Jang, S.; Kim, C.; Baek, K.; Kim, J.; Park, C. Human DJ-1 and its homologs are novel glyoxalases. Hum. Mol. Genet. 2012, 21, 3215–3225. [Google Scholar] [CrossRef] [Green Version]
  431. Bonifati, V.; Rizzu, P.; van Baren, M.J.; Schaap, O.; Breedveld, G.J.; Krieger, E.; Dekker, M.C.; Squitieri, F.; Ibanez, P.; Joosse, M.; et al. Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism. Science 2003, 299, 256–259. [Google Scholar] [CrossRef] [Green Version]
  432. Sironi, F.; Primignani, P.; Ricca, S.; Tunesi, S.; Zini, M.; Tesei, S.; Cilia, R.; Pezzoli, G.; Seia, M.; Goldwurm, S. DJ1 analysis in a large cohort of Italian early onset Parkinson Disease patients. Neurosci. Lett. 2013, 557, 165–170. [Google Scholar] [CrossRef]
  433. Weissbach, A.; Wittke, C.; Kasten, M.; Klein, C. ‘Atypical’ Parkinson’s disease—Genetic. Int. Rev. Neurobiol. 2019, 149, 207–235. [Google Scholar] [CrossRef]
  434. Abou-Sleiman, P.M.; Healy, D.G.; Quinn, N.; Lees, A.J.; Wood, N.W. The role of pathogenic DJ-1 mutations in Parkinson’s disease. Ann. Neurol. 2003, 54, 283–286. [Google Scholar] [CrossRef]
  435. Hague, S.; Rogaeva, E.; Hernandez, D.; Gulick, C.; Singleton, A.; Hanson, M.; Johnson, J.; Weiser, R.; Gallardo, M.; Ravina, B.; et al. Early-onset Parkinson’s disease caused by a compound heterozygous DJ-1 mutation. Ann. Neurol. 2003, 54, 271–274. [Google Scholar] [CrossRef]
  436. Annesi, G.; Savettieri, G.; Pugliese, P.; D’Amelio, M.; Tarantino, P.; Ragonese, P.; La Bella, V.; Piccoli, T.; Civitelli, D.; Annesi, F.; et al. DJ-1 mutations and parkinsonism-dementia-amyotrophic lateral sclerosis complex. Ann. Neurol. 2005, 58, 803–807. [Google Scholar] [CrossRef]
  437. Olzmann, J.A.; Brown, K.; Wilkinson, K.D.; Rees, H.D.; Huai, Q.; Ke, H.; Levey, A.I.; Li, L.; Chin, L.S. Familial Parkinson’s disease-associated L166P mutation disrupts DJ-1 protein folding and function. J. Biol. Chem. 2004, 279, 8506–8515. [Google Scholar] [CrossRef] [Green Version]
  438. 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]
  439. Kilarski, L.L.; Pearson, J.P.; Newsway, V.; Majounie, E.; Knipe, M.D.; Misbahuddin, A.; Chinnery, P.F.; Burn, D.J.; Clarke, C.E.; Marion, M.H.; et al. Systematic review and UK-based study of PARK2 (parkin), PINK1, PARK7 (DJ-1) and LRRK2 in early-onset Parkinson’s disease. Mov. Disord. 2012, 27, 1522–1529. [Google Scholar] [CrossRef]
  440. Kasten, M.; Hartmann, C.; Hampf, J.; Schaake, S.; Westenberger, A.; Vollstedt, E.J.; Balck, A.; Domingo, A.; Vulinovic, F.; Dulovic, M.; et al. Genotype-Phenotype Relations for the Parkinson’s Disease Genes Parkin, PINK1, DJ1: MDSGene Systematic Review. Mov. Disord. 2018, 33, 730–741. [Google Scholar] [CrossRef]
  441. Kumaran, R.; Vandrovcova, J.; Luk, C.; Sharma, S.; Renton, A.; Wood, N.W.; Hardy, J.A.; Lees, A.J.; Bandopadhyay, R. Differential DJ-1 gene expression in Parkinson’s disease. Neurobiol. Dis. 2009, 36, 393–400. [Google Scholar] [CrossRef]
  442. Choi, J.; Sullards, M.C.; Olzmann, J.A.; Rees, H.D.; Weintraub, S.T.; Bostwick, D.E.; Gearing, M.; Levey, A.I.; Chin, L.S.; Li, L. Oxidative damage of DJ-1 is linked to sporadic Parkinson and Alzheimer diseases. J. Biol. Chem. 2006, 281, 10816–10824. [Google Scholar] [CrossRef]
  443. Tanudjojo, B.; Shaikh, S.S.; Fenyi, A.; Bousset, L.; Agarwal, D.; Marsh, J.; Zois, C.; Heman-Ackah, S.; Fischer, R.; Sims, D.; et al. Phenotypic manifestation of α-synuclein strains derived from Parkinson’s disease and multiple system atrophy in human dopaminergic neurons. Nat. Commun. 2021, 12, 3817. [Google Scholar] [CrossRef] [PubMed]
  444. Mencke, P.; Boussaad, I.; Onal, G.; Kievit, A.J.A.; Boon, A.J.W.; Mandemakers, W.; Bonifati, V.; Kruger, R. Generation and characterization of a genetic Parkinson’s disease-patient derived iPSC line DJ-1-delP (LCSBi008-A). Stem. Cell. Res. 2022, 62, 102792. [Google Scholar] [CrossRef] [PubMed]
  445. Mazza, M.C.; Beilina, A.; Roosen, D.A.; Hauser, D.; Cookson, M.R. Generation of iPSC line from a Parkinson patient with PARK7 mutation and CRISPR-edited Gibco human episomal iPSC line to mimic PARK7 mutation. Stem. Cell. Res. 2021, 55, 102506. [Google Scholar] [CrossRef] [PubMed]
  446. Li, Y.; Ibanez, D.P.; Fan, W.; Zhao, P.; Chen, S.; Md Abdul, M.; Jiang, Y.; Fu, L.; Luo, Z.; Liu, Z.; et al. Generation of an induced pluripotent stem cell line (GIBHi004-A) from a Parkinson’s disease patient with mutant DJ-1/PARK7 (p.L10P). Stem. Cell. Res. 2020, 46, 101845. [Google Scholar] [CrossRef] [PubMed]
  447. Chen, L.; Cagniard, B.; Mathews, T.; Jones, S.; Koh, H.C.; Ding, Y.; Carvey, P.M.; Ling, Z.; Kang, U.J.; Zhuang, X. Age-dependent motor deficits and dopaminergic dysfunction in DJ-1 null mice. J. Biol. Chem. 2005, 280, 21418–21426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  448. Goldberg, M.S.; Pisani, A.; Haburcak, M.; Vortherms, T.A.; Kitada, T.; Costa, C.; Tong, Y.; Martella, G.; Tscherter, A.; Martins, A.; et al. Nigrostriatal dopaminergic deficits and hypokinesia caused by inactivation of the familial Parkinsonism-linked gene DJ-1. Neuron 2005, 45, 489–496. [Google Scholar] [CrossRef] [Green Version]
  449. Pham, T.T.; Giesert, F.; Rothig, A.; Floss, T.; Kallnik, M.; Weindl, K.; Holter, S.M.; Ahting, U.; Prokisch, H.; Becker, L.; et al. DJ-1-deficient mice show less TH-positive neurons in the ventral tegmental area and exhibit non-motoric behavioural impairments. Genes Brain Behav. 2010, 9, 305–317. [Google Scholar] [CrossRef]
  450. Raman, A.V.; Chou, V.P.; Atienza-Duyanen, J.; Di Monte, D.A.; Bellinger, F.P.; Manning-Bog, A.B. Evidence of oxidative stress in young and aged DJ-1-deficient mice. FEBS Lett. 2013, 587, 1562–1570. [Google Scholar] [CrossRef] [Green Version]
  451. Rousseaux, M.W.; Marcogliese, P.C.; Qu, D.; Hewitt, S.J.; Seang, S.; Kim, R.H.; Slack, R.S.; Schlossmacher, M.G.; Lagace, D.C.; Mak, T.W.; et al. Progressive dopaminergic cell loss with unilateral-to-bilateral progression in a genetic model of Parkinson disease. Proc. Natl. Acad. Sci. USA 2012, 109, 15918–15923. [Google Scholar] [CrossRef] [Green Version]
  452. Lin, Z.; Chen, C.; Yang, D.; Ding, J.; Wang, G.; Ren, H. DJ-1 inhibits microglial activation and protects dopaminergic neurons in vitro and in vivo through interacting with microglial p65. Cell. Death Dis. 2021, 12, 715. [Google Scholar] [CrossRef]
  453. Stauch, K.L.; Villeneuve, L.M.; Purnell, P.R.; Pandey, S.; Guda, C.; Fox, H.S. SWATH-MS proteome profiling data comparison of DJ-1, Parkin, and PINK1 knockout rat striatal mitochondria. Data Brief 2016, 9, 589–593. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, J.; Daadi, E.W.; Oh, T.; Daadi, E.S.; Daadi, M.M. Human Induced Pluripotent Stem Cell Phenotyping and Preclinical Modeling of Familial Parkinson’s Disease. Genes 2022, 13, 1937. https://doi.org/10.3390/genes13111937

AMA Style

Kim J, Daadi EW, Oh T, Daadi ES, Daadi MM. Human Induced Pluripotent Stem Cell Phenotyping and Preclinical Modeling of Familial Parkinson’s Disease. Genes. 2022; 13(11):1937. https://doi.org/10.3390/genes13111937

Chicago/Turabian Style

Kim, Jeffrey, Etienne W. Daadi, Thomas Oh, Elyas S. Daadi, and Marcel M. Daadi. 2022. "Human Induced Pluripotent Stem Cell Phenotyping and Preclinical Modeling of Familial Parkinson’s Disease" Genes 13, no. 11: 1937. https://doi.org/10.3390/genes13111937

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