Next Article in Journal
Age-Related Changes of the Synucleins Profile in the Mouse Retina
Next Article in Special Issue
Overview of the Impact of Pathogenic LRRK2 Mutations in Parkinson’s Disease
Previous Article in Journal
Multifunctional Modified Tumor Cell Membranes-Coated Adjuvant PTX against Melanoma
Previous Article in Special Issue
Age- and Sex-Dependent Behavioral and Neurochemical Alterations in hLRRK2-G2019S BAC Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Is Glial Dysfunction the Key Pathogenesis of LRRK2-Linked Parkinson’s Disease?

1
Department of Neurology, School of Medicine, Juntendo University, 2-1-1 Hongo, Bunkyo-ku, Tokyo 113-8421, Japan
2
Department of Research for Parkinson’s Disease, Graduate School of Medicine, Juntendo University, 2-1-1 Hongo, Bunkyo-ku, Tokyo 113-8421, Japan
3
Research Institute for Diseases of Old Age, Graduate School of Medicine, Juntendo University, 2-1-1 Hongo, Bunkyo-ku, Tokyo 113-8421, Japan
4
Center for Genomic and Regenerative Medicine, Graduate School of Medicine, Juntendo University, 2-1-1 Hongo, Bunkyo-ku, Tokyo 113-8421, Japan
5
Neurodegenerative Disorders Collaborative Laboratory, RIKEN Center for Brain Science, 2-1-Hirosawa, Wako-shi, Saitama 351-0198, Japan
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(1), 178; https://doi.org/10.3390/biom13010178
Submission received: 30 November 2022 / Revised: 12 January 2023 / Accepted: 13 January 2023 / Published: 15 January 2023
(This article belongs to the Special Issue Pathological Roles of LRRK2)

Abstract

:
Leucine rich-repeat kinase 2 (LRRK2) is the most well-known etiologic gene for familial Parkinson’s disease (PD). Its gene product is a large kinase with multiple functional domains that phosphorylates a subset of Rab small GTPases. However, studies of autopsy cases with LRRK2 mutations indicate a varied pathology, and the molecular functions of LRRK2 and its relationship to PD pathogenesis are largely unknown. Recently, non-autonomous neurodegeneration associated with glial cell dysfunction has attracted attention as a possible mechanism of dopaminergic neurodegeneration. Molecular studies of LRRK2 in astrocytes and microglia have also suggested that LRRK2 is involved in the regulation of lysosomal and other organelle dynamics and inflammation. In this review, we describe the proposed functions of LRRK2 in glial cells and discuss its involvement in the pathomechanisms of PD.

1. Introduction

Leucine rich-repeat kinase 2 (LRRK2) is the causative gene for PARK8-linked Parkinson’s disease (PD) and the most common etiologic gene for familial PD [1,2,3]. Among the various mutations of LRRK2, G2019S mutations are found in 4% of familial and 1% of sporadic PD cases [4]. Other missense mutations linked to PD pathogenesis have also been identified throughout the LRRK2 protein (Figure 1). Although LRRK2 is the most important causative gene of familial PD, its pathogenic mechanism remains unresolved. Autopsy brain studies of families with PD associated with LRRK2 mutations have shown the accumulation of TAR DNA binding protein of 43 kDa and tau inclusions with or without Lewy pathology [5], suggesting that the pathology may differ even within the same family [3]. Glial cell inclusions and progressive supranuclear palsy-like tau inclusions in the glia have also been reported [5,6]. Considering the multiple pathological findings including non-α-synuclein pathology and glial pathology in LRRK2-linked PD, a broad approach taking into account the whole brain environment may be required in addition to research focused on neurons and α-synuclein aggregation/propagation.
In the brain, LRRK2 is widely expressed in the neurons of the cortex, striatum, hippocampus, and substantia nigra [8,9], as well as in the microglia and astrocytes [10,11]. Moreover, LRRK2 shows a higher expression in inflammatory microglia than in resting-state microglia [10]. Several studies have proposed promising models of non-autonomous neuronal cell death caused by altered LRRK2 function in glial cells [12,13,14,15,16]. In this review, we summarize the glial phenotypes and altered organelle dynamics in glia caused by LRRK2 mutations and discuss the pathological mechanism of LRRK2-mediated non-autonomous neuronal cell death.

2. Structure and Function of LRRK2

LRRK2 is a large protein with a molecular weight of approximately 290 kDa (2527 amino acids) [3] and contains multiple domains, including an armadillo domain, an ankyrin domain, a Ras of complex protein (ROC) domain, a carboxyl-terminal of ROC (COR) domain, a kinase domain, and a WD40 domain (Figure 1). Due to the presence of both ROC and COR domains, LRRK2 and its paralog, LRRK1, belong to the ROCO family of proteins [17]. However, LRRK1 does not have an armadillo domain, which makes this domain unique to LRRK2. Rab29, a possible activator of LRRK2, and Rab10, a known LRRK2 substrate, are assumed to bind to the armadillo domain [18,19]. Depending on the cell context, LRRK2 exists as a monomer or as a dimer, wherein the dimer formation is mediated by binding via the ROC-COR and WD40 domains [20,21]. LRRK2 is mostly distributed in the cytoplasm as well as in vesicular membranes, such as endolysosomes [8,22,23] and synaptic vesicles [24,25,26], and regulates vesicular trafficking [27,28], mitochondrial homeostasis [29,30,31], and lysosomal function [11,22]. The loss of LRRK2 has been shown to alter the lysosomal structures in alveolar epithelial cells and urothelial cells in mice [32,33]. In primates, LRRK2 kinase inhibitors induced reversible histological changes, including vacuolation of the lung [34]. Additionally, the lungs and kidneys, where the obvious degeneration phenotypes are observed, show a strong LRRK2 expression [35]. Fibroblasts derived from patients with G2019S-associated PD also display lysosomal aggregation and enlargement [36]. In addition to the lysosomal function and trafficking, LRRK2 was revealed to be involved in autophagy. A study using cultured neurons demonstrated that LRRK2 G2019S or its upstream regulator Rab29 enhances JIP4 recruitment to autophagic vesicles via the phosphorylation of Rab10 and activates the kinesin motor, thereby disrupting the retrograde transport and maturation of autophagic vesicles [37]. LRRK2 is also reported to be a substrate for chaperone-mediated autophagy (CMA), which is disrupted by the I2020T and G2019S mutations [38].
LRRK2 was reported to show microtubule-associated localization, forming a skein-like structure [39,40]. Cryo-electron microscopic analysis further revealed that LRRK2 is polymerized on the surface of microtubules in a helical fashion [41,42]. Furthermore, LRRK2 was shown to be associated with non-acetylated microtubules [43]. In Drosophila, LRRK2 pathogenic mutants in the ROC-COR domain exhibited a strong interference in axonal transport, which was reversed by the promotion of microtubule acetylation [43]. In vitro studies showed that a truncated LRRK2 with the ROC-COR, kinase, and WD40 domains inhibited the motility of both kinesin and dynein motors on the microtubules in a concentration-dependent manner [41]. Moreover, LRRK2 interfered with primary cilia formation in a kinase-dependent manner by enhancing the binding of phospho-Rab10 with its effector, Rab-interacting lysosomal protein-like protein [44,45]. Additionally, Hedgehog signaling, which is sensed by the primary cilia, was altered in LRRK2 G2019 and R1441C knock-in mice, suggesting that the synaptic activity and survival of the striatal neurons and astrocytes are compromised by cilia loss [45,46].
The ROC domain has a Rab-like small GTPase structure and functions as a GTPase switch to regulate LRRK2 kinase activity [47,48]. Several phosphorylation targets of LRRK2 have been determined, which include Rab3A/B/C/D, Rab8A/B, Rab10, Rab12, Rab35, and Rab43 as endogenous substrates, Rab5B/C and Rab29 as potential substrates [19,44,49,50,51,52,53], along with nuclear translocation of nuclear factor of activated T-cells 1 (NFATc2/NFAT1), an inflammation-related transcription factor [14], and Wiskott–Aldrich syndrome protein family verprolin-homologous protein 2 (WAVE2), a regulator of actin dynamics [54]. Rab29, the gene of which is located in the PARK16 locus [55], was characterized as an upstream molecule of LRRK2, recruiting LRRK2 to the vesicle membrane [56,57]. However, the involvement of Rab29 in LRRK2 signaling is under debate because LRRK2 was also activated in Rab29 knock-out cells [58]. Many Rab proteins are involved in vesicular trafficking, such as synaptic dense core vesicles, endolysosomes, and trans-Golgi network. Thus, the alteration of vesicular trafficking may be a major pathogenic mechanism of LRRK2 mutation-associated PD [59].

3. Mammalian Model Studies of LRRK2

Studies of various transgenic LRRK2 mouse models revealed a possible role for LRRK2 in the dysfunction of striatal neurotransmission and dopaminergic homeostasis, yet the results are not consistent. For example, studies of mice overexpressing LRRK2 G2019S or R1441G in dopaminergic neurons showed reduced dopamine production or release and loss of the midbrain dopaminergic neurons [60,61]. Others reported an altered morphology of dopaminergic neurons without neuronal loss [62,63]. These differences could arise from the insertion loci of LRRK2 transgenes or the expression levels of the transgene itself, which may confound the LRRK2 phenotypes observed in different models and hinder our understanding of the roles of LRRK2 expressed at physiological levels. On the other hand, LRRK2 knock-in models, which are expected to have physiological levels of expression and distribution of gene expression, show only minor overall abnormalities in neurological function, limiting their value as disease models [64,65]. Surprisingly, LRRK2 knockout mice and rats show abnormalities in the lungs and kidneys, and the role of LRRK2 in non-neural tissues is currently being elucidated [64,65]. Importantly, in one study, microglial activation was suppressed in LRRK2 knockout rats, making them resistant to toxicity induced by α-synuclein and lipopolysaccharide (LPS) [66]. This observation strongly suggests a role for LRRK2 in glial cells’ physiology. Furthermore, LRRK2 is highly expressed in monocytes and macrophages, and the brain infiltration of macrophages differentiated from monocytes could contribute to pathology [10,67,68,69]. After birth, macrophages’ and microglia’s developmental trajectory diverges and the two populations may play different roles in the brain [70]. However, the contribution of monocytes and macrophages to LRRK2 pathology cannot be ignored. Thus, this review will also mention studies on these cells along with those on astrocytes and microglia (Figure 2).

4. Astrocytes

4.1. Astrocytes and PD

Astrocytes are the major glial cells in the central nervous system and are essential for providing the appropriate environment for neurons in the adult brain [73]. Astrocytes’ roles include the metabolic exchange with neurons, the regulation of neurotransmitter cycle in the synaptic cleft, and the maintenance of the blood–brain barrier [74,75]. Recently, changes in the astrocytic function have been suggested to lead to a pathological environment in several neurodegenerative diseases, such as Alzheimer’s and Huntington’s diseases [76,77]. Astrocytes are also of interest in PD because of the postulation of inflammatory astrocyte involvement in PD neurodegeneration [78,79,80]. The astrocytic expression of PD causative genes, including LRRK2, also strongly suggests the involvement of astrocytes in PD pathogenesis [81].

4.2. Role of LRRK2 in Astrocytes

As mentioned in Section 2, LRRK2 may regulate Rab-mediated vesicular trafficking. In mouse primary astrocytes, the lysosome number and size are dependent on LRRK2 kinase activity [11]. Lysosomal hypertrophy was observed in the presence of LRRK2 G2019S, R1441C, and Y1699C, which exhibit increased kinase activity [11]. Among the lysosome-associated molecules, mutations of Beta-glucocerebrosidase 1 (GBA1) have been identified as a risk factor for PD [82,83]. Lysosomal alkalinization and dysfunction were observed in primary astrocytes from GBA1 D409V knock-in mice, which were improved by the treatment with an LRRK2 inhibitor, suggesting that GBA1 is a lysosomal target of LRRK2 [84]. Lysosomal membrane permeabilization by l-leucyl-l-leucine methyl ester (LLOME), a lysosomotropic reagent, recruited LRRK2 to ruptured lysosomes in primary astrocytes, where activated LRRK2 phosphorylated Rab10 and recruited the motor adaptor protein JIP4 to the microtubules, leading to lysosomal tubulation and budding [23]. Although the physiological significance of this phenomenon is unclear, lysosomal tubulation and budding may act as a protective response by sorting undigested proteins to other active lysosomes for proper degradation [85]. Microtubule-dependent lysosomal positioning is determined by the kinesin motor-associated BLOC-one-related complex and Arf-like small GTPase (ARL8B) [86] and JIP4, the dynein motor-associated protein [87]. The movement of lysosomes to the perinuclear region is linked to autophagy activation [88] and also occurred in response to LLOME treatment [89]. The positioning of LRRK2-positive lysosomes determined the substrate specificity of Rab GTPase [89]. In astrocytes and HEK293FT cells, the recruitment and phosphorylation of Rab10 by LRRK2 only occurred in the perinuclear lysosomes. In contrast, Rab12 was recruited and phosphorylated by LRRK2 in both the perinuclear and the peripheral lysosomes [89]. Although PPM1H, the phosphatase of Rab GTPases, may be involved in this phenomenon, the detailed mechanism requires further analysis [90].
The role of LRRK2 in vesicular trafficking has been reported in a variety of experimental astrocyte systems, which suggest that both α-synuclein-associated and -non-associated mechanisms are involved in non-autonomous neuronal cell death. Endolysosomal dysfunction and impaired clearance of α-synuclein due to decreased phagocytosis-related annexin A2 have been demonstrated in primary astrocytes harboring LRRK2 G2019S [91]. In a study using induced pluripotent stem (iPS) cells derived from patients with LRRK2 G2019S-associated PD, the accumulation of α-synuclein in astrocytes due to impaired autophagy was also observed [12]. In a co-culture of LRRK2 G2019S astrocytes with dopaminergic neurons from healthy controls, astrocyte-derived pathogenic α-synuclein was transported into the dopaminergic neurons, resulting in neurodegeneration [12].
Changes in extracellular vesicles (EVs)/multivesicular bodies (MVBs) have also been observed in LRRK2 G2019S astrocytes derived from human iPS cells. Compared to normal astrocytes, LRRK2 G2019S astrocytes have smaller MVBs, accompanied with the accumulation of phospho-S129 α-synuclein and a lower EV release, while astrocyte–dopaminergic neuron co-culture systems have shown that the alteration of EV biogenesis in LRRK2 G2019S astrocytes induced neurite shortening and cell death in dopaminergic neurons [13]. Another human iPS cell study reported that LRRK2 G2019S astrocytes from patients with PD have reduced mitochondrial function and impaired neuronal interactions, mainly, synaptic interactions [92].
In summary, LRRK2 function in astrocytes may be closely related to organelle dynamics (Figure 3). Additionally, Rab proteins may very likely play an important role because of their downstream location. In LRRK2 pathological mutations, non-autonomous neuronal cell death may also be associated with altered organelle dynamics, suggesting that research on this mechanism may lead to a better understanding of PD pathogenesis. Recent single-cell analyses suggest that astrocytes may differ in function and properties in different brain regions [93,94]. Elucidating the possibility that LRRK2 may be of particular importance for local astrocytes in the substantia nigra and striatum will be an important future challenge.

5. Microglia

5.1. Microglial Cells and PD Etiology

Inflammation-responsive microglia with human leukocyte antigen-DR isotype have been observed in the brain pathology of PD; the extent to which reactive microglia contribute to PD pathogenesis remains undetermined [95]. Microglial innate immune signaling in mouse models of Alzheimer’s disease was found to cause synapse loss [96,97]. A pathway enrichment analysis from a meta-genome-wide association study (meta-GWAS) suggested at least some immune component involvement in PD development risk [98]. Considering these findings, the relation between PD etiology and immune responses is still a debated topic.
Cx3cr1-, CD11b-, Iba1-, and F4/80-positive microglial cells are a subset of glia that occupy approximately 10% of the brain. Microglia are resident intracranial immune cells with functions similar to those of macrophages [99]. They are mesenchymal and myeloid cells derived from the yolk sac and have self-renewal capability [100]. Physiologically, their main functions are sensing and housekeeping of the surrounding environment and protection against toxic substances [101]. Microglial uptake and degradation of extracellular α-synuclein aggregates is the most effective among all central nervous system cells in vitro, suggesting that microglia are major clearance cells for α-synuclein [4]. Extracellular α-synuclein released from neurons may be phagocytosed via toll-like receptor (TLR) 4, wherein TLR4 transcriptionally upregulated p62/SQSTM1, an autophagy receptor, via nuclear factor (NF)-κB signaling to lead to the degradation of α-synuclein by the autophagy–lysosomal pathway in microglia [102]. Although TLR4 exerted protective effects on dopaminergic neurons in terms of α-synuclein degradation, the TLR2–MyD88–NF-κB pathway was shown to promote α-synuclein propagation [103]. TLR2 recognized α-synuclein preformed fibrils because they could be a strong ligand of TLR2 [103]. α-synuclein preformed fibrils then activated the microglial TLR2–MyD88–NF-κB pathway and stimulated the secretion of proinflammatory cytokines, such as interleukin (IL)-1β and tumor necrosis factor-α (TNF-α) [103]. These microglia-derived proinflammatory molecules in turn transcriptionally upregulated α-synuclein levels in neurons, promoting α-synuclein propagation [103]. Conversely, the blocking of the TLR2 cascade was shown to ameliorate neuroinflammation and prevent α-synuclein fibril propagation in a mouse model of α-synuclein propagation [103]. This differential responsiveness of microglia to physiological and pathogenic α-synuclein via TLR2 and TLR4 suggests that microglia are involved in the pathological mechanisms of PD. Moreover, the activation of NF-κB signaling, which plays a central role in immune responses, was suggested to be protective in neurons and to contribute to neurotoxicity in glia [104]. These contradictory observations may be due to a focus on different effects of the NF-κB pathway.

5.2. Role of LRRK2 in Microglia

LRRK2 has been implicated in inflammation because LRRK2 gene variants are linked to autoimmune diseases, such as Crohn’s disease and systemic lupus erythematosus [105,106]. LRRK2 is abundantly expressed in innate immune cells, including monocytes and neutrophils [107], where its expression is upregulated by interferon (IFN) γ [108,109]. At the cellular level, the amplification of inflammatory signals via LRRK2 mutations was observed in microglia as well as in monocytes and macrophages, with similar mechanisms [10,110]. Inflammatory cytokines are elevated in sporadic and LRRK2-linked PD [111]. Moreover, a correlation between disease severity and elevated levels of inflammatory cytokines was observed in the patients with LRRK2-linked PD [112].
Single-cell analysis also revealed a variant in the 5′ untranslated region of the LRRK2 gene isolated in a GWAS of sporadic PD that was enhanced by microglia-specific LRRK2 expression [113]. Human iPS cell-derived monocytes and macrophages with LRRK2 G2019S were shown to potentiate the production of inflammatory cytokines, such as IL-1β, TNF-α, and IFNγ, following the activation of TLRs in a LRRK2 kinase activity-independent manner [110]. IFNγ also increased LRRK2 expression in neurons [15]. Moreover, conditioned medium from LPS-activated LRRK2 G2019S microglia, in which the nuclear translocation of NFATc2/NFAT1 was impaired, suppressed neurite outgrowth in both normal and G2019S neurons [15]. Protein kinase A-mediated phosphorylation of p50, the NF-κB inhibitory subunit, was negatively regulated by LRRK2, thereby contributing to the production of IL-1β in microglia [114].
Amoeboid-like immune cells, such as microglia and macrophages, are constantly changing their morphology in response to foreign entities. Considering that the lack of GbpC, a ROCO family protein, affects chemotaxis in Dictyostelium, an amoeboid organism, the ROCO-family proteins may be involved in the functions and motility of amoeboid-like cells [115]. In Dictyostelium, the cGMP-binding GbpC is involved in the chemotaxis mechanism via phosphorylation of myosin II [116].
LRRK2 phosphorylated and stabilized WAVE2, a regulator of the actin cytoskeleton, leading to a positive regulation of phagocytosis in microglia [54]. LPS-mediated activation of microglia has been demonstrated to promote dopaminergic neuron death via the LRRK2–WAVE2 pathway [54]. Apart from the positive regulation of phagocytosis via WAVE2 phosphorylation, LRRK2 was also shown to negatively regulate the migration of microglia through the suppression of focal adhesion kinase via phosphorylation [117]. These results are indirectly supported by the observation that CX3CL1, which activates microglial motility, showed increased levels in LRRK2-null microglia [118]. In contrast to that observed in microglia, the inhibition of LRRK2 kinase attenuated chemotaxis to α-synuclein fibrils in primary cultured macrophages and the infiltration of peripheral monocytes into the central nervous system [69]. Moreover, the PD-linked LRRK2 G2294R mutation in the WD40 domain, which destabilized the LRRK2 protein, resulted in an attenuated macrophage uptake of α-synuclein fibrils [119].
In the study analyzing the inflammatory cascade, NFATc2 was reported to be phosphorylated by LRRK2 in microglia exposed to α-synuclein and translocated into the nucleus for transcriptional regulation, and this mechanism was inhibited in the primary microglia of LRRK2 knock-out mice [14]. However, another study of iPS microglia derived from patients with the G2019S mutation showed inhibition of the nuclear migration of NFATc2 [15]. Additionally, the study also reported that the microtubule-dependent nuclear translocation of NFATc4 was impaired in neurons, leading to defects in neurite outgrowth [15].
LRRK2 pathogenic mutants were demonstrated to sequester Rab8 to the lysosomes via phosphorylation, directing the transferrin-mediated iron endocytosis pathway in microglia from recycling to degradation [16]. Mislocalization of transferrin at the lysosomes was also observed in LPS-induced inflammatory microglia [16]. Moreover, iron and ferritin deposition in inflammatory microglia was observed in the striatum of G2019S knock-in mice [16].
In the macrophage cell line RAW264.7, LRRK2 accumulated and recruited Rab8 and Rab10 under lysosomal stress in a Rab29-dependent manner, which led to the extracellular release of the lysosomal contents [22,23]. LRRK2 was also mobilized to the lysosomes damaged by LLOME. Moreover, Rab10-mediated lysosomal tubulation and budding to the plasma membrane were detected in astrocytes [23]. Furthermore, lysosome enlargement due to LRRK2 inhibition in RAW264.7 cells suggested that LRRK2 is involved in lysosome quality control, as proposed in the previously mentioned astrocyte study that revealed LRRK2-dependent lysosomal budding [22].
LRRK2 was also translocated to maturing phagosomes in human iPS cell-derived macrophages in a kinase activity-independent manner [109]. Although LRRK2 was not involved in the initial phagocytosis, LRRK2 kinase activity was required for the mobilization of Rab8a and Rab10 to the phagosomes, suggesting that LRRK2 is required in the phagosome maturation pathway [109].
Microglia exchange mitochondria and materials through F-actin-associated nanotubes [31]. Microglia that had taken up excess amounts of α-synuclein fibrils transported these fibrils to other microglia through F-actin-associated nanotubes, while inflammatory microglia accepted healthy mitochondria from naive microglia to improve survival and diminish the inflammatory profile [31]. In this context, LRRK2 G2019S disturbed this transport in a kinase-activity-independent manner, which could contribute to the exacerbation of inflammation by α-synuclein fibrils [31].
In addition to the abnormalities in vesicle trafficking and lysosomal function, microglial studies indicated a link between LRRK2 and both inflammation and phagocytosis (Figure 4). As noted in the section on astrocytes, microglia may also be involved in non-autonomous neuronal cell death.

6. Conclusions and Perspective

We reviewed LRRK2 studies in glial cells, wherein LRRK2 is suggested to have diverse roles in non-autonomous neuronal cell death/neuroinflammation. The dysregulation of organelle dynamics, including endolysosomes, was demonstrated in LRRK2-mutated astrocytes, while the disturbance of the inflammatory cascade and phagocytosis/migration mechanism has been proposed in LRRK2-mutated microglia in addition to the dysfunction of organelle dynamics and mitochondria. In contrast, the pathological analyses of PD brain with LRRK2 R1441H [120] and G2385R [121] indicated that the inflammatory profile is not as prominent as observed in the basic experiments. Although these results can be attributed to the fact that the pathological analysis studies may have detected results after the inflammatory response had terminated (at the time when neuronal death was already complete), this issue is still unresolved.
Recently, parenchymal border macrophages have been reported to regulate the flow dynamics of the cerebrospinal fluid [122]. Given the diversity of the LRRK2 pathology, the disruption of the glymphatic system by macrophages with LRRK2 mutations may result in an impaired efflux of aberrant proteins [123].
Phenotypic analyses of LRRK2 knock-out animals showed degeneration of the renal proximal tubules and pulmonary epithelial cells [124], enlargement of lysosomes [125,126], altered axonal transport [127], and synaptic dysregulation [24,26]. In contrast, the nigrostriatal dopaminergic system was not functionally compromised in LRRK2 knock-out mice [32] and human cases with LRRK2 loss-of-function variants [128]. LRRK2 kinase inhibitors [129] are currently being investigated based on the results that the causes of human LRRK2 loss of function are not strongly associated with any specific phenotype or disease state [128]. Especially, a phase III trial has begun for an LRRK2 inhibitor, BIIB122, that is being developed by Denali and Biogen [130]. This drug is the most advanced Parkinson’s disease modifying drug in development. However, considering that the broad physiological role of LRRK2 is potentially not limited to kinase activity, further analysis may be needed for a more optimal drug development.

Author Contributions

Writing—original draft preparation, T.I.; writing—review and editing, Y.I.; supervision, N.H.; and funding acquisition, Y.I. and N.H. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Grants-in-Aid for Scientific Research (17H04049 and 20H03453 to Y.I.) from JSPS in Japan, a grant from the Japan Agency for Medical Research and Development, Research on Development of New Drugs, GAPFREE (JP19ak0101112 to Y.I. and N.H.), and was partly supported by a grant from Otsuka Pharmaceutical (to Y.I. and N.H.) and AbbVie GK (to Y.I. and N.H.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We thank Daisuke Taniguchi and Tosiki Tezuka for their discussion of the research associated with this review.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; collection, analysis, and interpretation of the data; writing of the manuscript; or decision to publish the results.

Abbreviations

CMAChaperone-mediated autophagy
CORCarboxyl-terminal of ROC
EVExtracellular vesicle
GWASGenome-wide association study
IFNInterferon
ILInterleukin
iPSInduced pluripotent stem
LLOMEl-leucyl-l-leucine methyl ester
LPS Lipopolysaccharide
LRRK2Leucine rich-repeat kinase 2
MVBsMultivesicular bodies
NFAT Nuclear translocation of nuclear factor of activated T-cells
NF-κB Nuclear factor-κB
PDParkinson’s disease
ROCRas of complex protein
TLRToll-like receptor
TNF-αTumor necrosis factor-α
WAVE2Wiskott–Aldrich syndrome protein family verprolin-homologous protein 2

References

  1. 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]
  2. Paisán-Ruíz, C.; Jain, S.; Evans, E.W.; Gilks, W.P.; Simón, J.; van der Brug, M.; López 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]
  3. 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] [PubMed] [Green Version]
  4. Healy, D.G.; Falchi, M.; O’Sullivan, S.S.; Bonifati, V.; Durr, A.; Bressman, S.; Brice, A.; Aasly, J.; Zabetian, C.P.; Goldwurm, S.; et al. Phenotype, genotype, and worldwide genetic penetrance of LRRK2-associated Parkinson’s disease: A case-control study. Lancet Neurol. 2008, 7, 583–590. [Google Scholar] [CrossRef] [Green Version]
  5. Kalia, L.V.; Lang, A.E.; Hazrati, L.N.; Fujioka, S.; Wszolek, Z.K.; Dickson, D.W.; Ross, O.A.; Van Deerlin, V.M.; Trojanowski, J.Q.; Hurtig, H.I.; et al. Clinical correlations with Lewy body pathology in LRRK2-related Parkinson disease. JAMA Neurol. 2015, 72, 100–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hasegawa, K.; Stoessl, A.J.; Yokoyama, T.; Kowa, H.; Wszolek, Z.K.; Yagishita, S. Familial parkinsonism: Study of original Sagamihara PARK8 (I2020T) kindred with variable clinicopathologic outcomes. Park. Relat. Disord 2009, 15, 300–306. [Google Scholar] [CrossRef] [Green Version]
  7. Tolosa, E.; Vila, M.; Klein, C.; Rascol, O. LRRK2 in Parkinson disease: Challenges of clinical trials. Nat. Rev. Neurol. 2020, 16, 97–107. [Google Scholar] [CrossRef] [PubMed]
  8. Biskup, S.; Moore, D.J.; Celsi, F.; Higashi, S.; West, A.B.; Andrabi, S.A.; Kurkinen, K.; Yu, S.W.; Savitt, J.M.; Waldvogel, H.J.; et al. Localization of LRRK2 to membranous and vesicular structures in mammalian brain. Ann. Neurol. 2006, 60, 557–569. [Google Scholar] [CrossRef]
  9. Hatano, T.; Kubo, S.; Imai, S.; Maeda, M.; Ishikawa, K.; Mizuno, Y.; Hattori, N. Leucine-rich repeat kinase 2 associates with lipid rafts. Hum. Mol. Genet. 2007, 16, 678–690. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Henry, A.G.; Aghamohammadzadeh, S.; Samaroo, H.; Chen, Y.; Mou, K.; Needle, E.; Hirst, W.D. Pathogenic LRRK2 mutations, through increased kinase activity, produce enlarged lysosomes with reduced degradative capacity and increase ATP13A2 expression. Hum. Mol. Genet. 2015, 24, 6013–6028. [Google Scholar] [CrossRef] [Green Version]
  12. di Domenico, A.; Carola, G.; Calatayud, C.; Pons-Espinal, M.; Muñoz, J.P.; Richaud-Patin, Y.; Fernandez-Carasa, I.; Gut, M.; Faella, A.; Parameswaran, J.; et al. Patient-Specific iPSC-Derived Astrocytes Contribute to Non-Cell-Autonomous Neurodegeneration in Parkinson’s Disease. Stem Cell Rep. 2019, 12, 213–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. de Rus Jacquet, A.; Tancredi, J.L.; Lemire, A.L.; DeSantis, M.C.; Li, W.P.; O’Shea, E.K. The LRRK2 G2019S mutation alters astrocyte-to-neuron communication via extracellular vesicles and induces neuron atrophy in a human iPSC-derived model of Parkinson’s disease. Elife 2021, 10, e73062. [Google Scholar] [CrossRef] [PubMed]
  14. Kim, C.; Beilina, A.; Smith, N.; Li, Y.; Kim, M.; Kumaran, R.; Kaganovich, A.; Mamais, A.; Adame, A.; Iba, M.; et al. LRRK2 mediates microglial neurotoxicity via NFATc2 in rodent models of synucleinopathies. Sci. Transl. Med. 2020, 12, eaay0399. [Google Scholar] [CrossRef]
  15. Panagiotakopoulou, V.; Ivanyuk, D.; De Cicco, S.; Haq, W.; Arsić, A.; Yu, C.; Messelodi, D.; Oldrati, M.; Schöndorf, D.C.; Perez, M.J.; et al. Interferon-γ signaling synergizes with LRRK2 in neurons and microglia derived from human induced pluripotent stem cells. Nat. Commun. 2020, 11, 5163. [Google Scholar] [CrossRef] [PubMed]
  16. Mamais, A.; Kluss, J.H.; Bonet-Ponce, L.; Landeck, N.; Langston, R.G.; Smith, N.; Beilina, A.; Kaganovich, A.; Ghosh, M.C.; Pellegrini, L.; et al. Mutations in LRRK2 linked to Parkinson disease sequester Rab8a to damaged lysosomes and regulate transferrin-mediated iron uptake in microglia. PLoS Biol. 2021, 19, e3001480. [Google Scholar] [CrossRef] [PubMed]
  17. Marín, I. The Parkinson disease gene LRRK2: Evolutionary and structural insights. Mol. Biol. Evol. 2006, 23, 2423–2433. [Google Scholar] [CrossRef] [Green Version]
  18. Vides, E.G.; Adhikari, A.; Chiang, C.Y.; Lis, P.; Purlyte, E.; Limouse, C.; Shumate, J.L.; Spínola-Lasso, E.; Dhekne, H.S.; Alessi, D.R.; et al. A feed-forward pathway drives LRRK2 kinase membrane recruitment and activation. Elife 2022, 11, e79771. [Google Scholar] [CrossRef]
  19. McGrath, E.; Waschbüsch, D.; Baker, B.M.; Khan, A.R. LRRK2 binds to the Rab32 subfamily in a GTP-dependent manner via its armadillo domain. Small GTPases 2021, 12, 133–146. [Google Scholar] [CrossRef]
  20. Berger, Z.; Smith, K.A.; Lavoie, M.J. Membrane localization of LRRK2 is associated with increased formation of the highly active LRRK2 dimer and changes in its phosphorylation. Biochemistry 2010, 49, 5511–5523. [Google Scholar] [CrossRef]
  21. Zhang, P.; Fan, Y.; Ru, H.; Wang, L.; Magupalli, V.G.; Taylor, S.S.; Alessi, D.R.; Wu, H. Crystal structure of the WD40 domain dimer of LRRK2. Proc. Natl. Acad. Sci. USA 2019, 116, 1579–1584. [Google Scholar] [CrossRef] [Green Version]
  22. Eguchi, T.; Kuwahara, T.; Sakurai, M.; Komori, T.; Fujimoto, T.; Ito, G.; Yoshimura, S.I.; Harada, A.; Fukuda, M.; Koike, M.; et al. LRRK2 and its substrate Rab GTPases are sequentially targeted onto stressed lysosomes and maintain their homeostasis. Proc. Natl. Acad. Sci. USA 2018, 115, E9115–E9124. [Google Scholar] [CrossRef] [Green Version]
  23. Bonet-Ponce, L.; Beilina, A.; Williamson, C.D.; Lindberg, E.; Kluss, J.H.; Saez-Atienzar, S.; Landeck, N.; Kumaran, R.; Mamais, A.; Bleck, C.K.E.; et al. LRRK2 mediates tubulation and vesicle sorting from lysosomes. Sci. Adv. 2020, 6, eabb2454. [Google Scholar] [CrossRef]
  24. 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]
  25. Piccoli, G.; Onofri, F.; Cirnaru, M.D.; Kaiser, C.J.; Jagtap, P.; Kastenmuller, A.; Pischedda, F.; Marte, A.; von Zweydorf, F.; Vogt, A.; et al. Leucine-rich repeat kinase 2 binds to neuronal vesicles through protein interactions mediated by its C-terminal WD40 domain. Mol. Cell Biol. 2014, 34, 2147–2161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. 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] [PubMed] [Green Version]
  27. Imai, Y.; Kobayashi, Y.; Inoshita, T.; Meng, H.; Arano, T.; Uemura, K.; Asano, T.; Yoshimi, K.; Zhang, C.L.; Matsumoto, G.; et al. The Parkinson’s Disease-Associated Protein Kinase LRRK2 Modulates Notch Signaling through the Endosomal Pathway. PLoS Genet. 2015, 11, e1005503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. 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]
  29. Wauters, F.; Cornelissen, T.; Imberechts, D.; Martin, S.; Koentjoro, B.; Sue, C.; Vangheluwe, P.; Vandenberghe, W. LRRK2 mutations impair depolarization-induced mitophagy through inhibition of mitochondrial accumulation of RAB10. Autophagy 2020, 16, 203–222. [Google Scholar] [CrossRef]
  30. Liu, H.; Ho, P.W.; Leung, C.T.; Pang, S.Y.; Chang, E.E.S.; Choi, Z.Y.; Kung, M.H.; Ramsden, D.B.; Ho, S.L. Aberrant mitochondrial morphology and function associated with impaired mitophagy and DNM1L-MAPK/ERK signaling are found in aged mutant Parkinsonian LRRK2(R1441G) mice. Autophagy 2021, 17, 3196–3220. [Google Scholar] [CrossRef] [PubMed]
  31. Scheiblich, H.; Dansokho, C.; Mercan, D.; Schmidt, S.V.; Bousset, L.; Wischhof, L.; Eikens, F.; Odainic, A.; Spitzer, J.; Griep, A.; et al. Microglia jointly degrade fibrillar alpha-synuclein cargo by distribution through tunneling nanotubes. Cell 2021, 184, 5089–5106.e5021. [Google Scholar] [CrossRef] [PubMed]
  32. Hinkle, K.M.; Yue, M.; Behrouz, B.; Dächsel, J.C.; Lincoln, S.J.; Bowles, E.E.; Beevers, J.E.; Dugger, B.; Winner, B.; Prots, I.; et al. LRRK2 knockout mice have an intact dopaminergic system but display alterations in exploratory and motor co-ordination behaviors. Mol. Neurodegener. 2012, 7, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. 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]
  34. Baptista, M.A.S.; Merchant, K.; Barrett, T.; Bhargava, S.; Bryce, D.K.; Ellis, J.M.; Estrada, A.A.; Fell, M.J.; Fiske, B.K.; Fuji, R.N.; et al. LRRK2 inhibitors induce reversible changes in nonhuman primate lungs without measurable pulmonary deficits. Sci. Transl. Med. 2020, 12, eaav0820. [Google Scholar] [CrossRef] [PubMed]
  35. Biskup, S.; Moore, D.J.; Rea, A.; Lorenz-Deperieux, B.; Coombes, C.E.; Dawson, V.L.; Dawson, T.M.; West, A.B. Dynamic and redundant regulation of LRRK2 and LRRK1 expression. BMC Neurosci. 2007, 8, 102. [Google Scholar] [CrossRef] [Green Version]
  36. Hockey, L.N.; Kilpatrick, B.S.; Eden, E.R.; Lin-Moshier, Y.; Brailoiu, G.C.; Brailoiu, E.; Futter, C.E.; Schapira, A.H.; Marchant, J.S.; Patel, S. Dysregulation of lysosomal morphology by pathogenic LRRK2 is corrected by TPC2 inhibition. J. Cell Sci. 2015, 128, 232–238. [Google Scholar] [CrossRef] [Green Version]
  37. Boecker, C.A.; Goldsmith, J.; Dou, D.; Cajka, G.G.; Holzbaur, E.L.F. Increased LRRK2 kinase activity alters neuronal autophagy by disrupting the axonal transport of autophagosomes. Curr. Biol. 2021, 31, 2140–2154.e2146. [Google Scholar] [CrossRef]
  38. Orenstein, S.J.; Kuo, S.H.; Tasset, I.; Arias, E.; Koga, H.; Fernandez-Carasa, I.; Cortes, E.; Honig, L.S.; Dauer, W.; Consiglio, A.; et al. Interplay of LRRK2 with chaperone-mediated autophagy. Nat. Neurosci. 2013, 16, 394–406. [Google Scholar] [CrossRef] [Green Version]
  39. Kett, L.R.; Boassa, D.; Ho, C.C.; Rideout, H.J.; Hu, J.; Terada, M.; Ellisman, M.; Dauer, W.T. LRRK2 Parkinson disease mutations enhance its microtubule association. Hum. Mol. Genet. 2012, 21, 890–899. [Google Scholar] [CrossRef] [Green Version]
  40. Law, B.M.; Spain, V.A.; Leinster, V.H.; Chia, R.; Beilina, A.; Cho, H.J.; Taymans, J.M.; Urban, M.K.; Sancho, R.M.; Blanca Ramírez, M.; et al. A direct interaction between leucine-rich repeat kinase 2 and specific β-tubulin isoforms regulates tubulin acetylation. J. Biol. Chem. 2014, 289, 895–908. [Google Scholar] [CrossRef]
  41. Deniston, C.K.; Salogiannis, J.; Mathea, S.; Snead, D.M.; Lahiri, I.; Matyszewski, M.; Donosa, O.; Watanabe, R.; Böhning, J.; Shiau, A.K.; et al. Structure of LRRK2 in Parkinson’s disease and model for microtubule interaction. Nature 2020, 588, 344–349. [Google Scholar] [CrossRef] [PubMed]
  42. Watanabe, R.; Buschauer, R.; Böhning, J.; Audagnotto, M.; Lasker, K.; Lu, T.W.; Boassa, D.; Taylor, S.; Villa, E. The In Situ Structure of Parkinson’s Disease-Linked LRRK2. Cell 2020, 182, 1508–1518.e1516. [Google Scholar] [CrossRef]
  43. 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]
  44. Steger, M.; Diez, F.; Dhekne, H.S.; Lis, P.; Nirujogi, R.S.; Karayel, O.; Tonelli, F.; Martinez, T.N.; Lorentzen, E.; Pfeffer, S.R.; et al. Systematic proteomic analysis of LRRK2-mediated Rab GTPase phosphorylation establishes a connection to ciliogenesis. Elife 2017, 6, e31012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Dhekne, H.S.; Yanatori, I.; Gomez, R.C.; Tonelli, F.; Diez, F.; Schüle, B.; Steger, M.; Alessi, D.R.; Pfeffer, S.R. A pathway for Parkinson’s Disease LRRK2 kinase to block primary cilia and Sonic hedgehog signaling in the brain. Elife 2018, 7, e40202. [Google Scholar] [CrossRef] [PubMed]
  46. Khan, S.S.; Sobu, Y.; Dhekne, H.S.; Tonelli, F.; Berndsen, K.; Alessi, D.R.; Pfeffer, S.R. Pathogenic LRRK2 control of primary cilia and Hedgehog signaling in neurons and astrocytes of mouse brain. Elife 2021, 10, e67900. [Google Scholar] [CrossRef]
  47. Ito, G.; Okai, T.; Fujino, G.; Takeda, K.; Ichijo, H.; Katada, T.; Iwatsubo, T. GTP binding is essential to the protein kinase activity of LRRK2, a causative gene product for familial Parkinson’s disease. Biochemistry 2007, 46, 1380–1388. [Google Scholar] [CrossRef]
  48. Li, X.; Tan, Y.C.; Poulose, S.; Olanow, C.W.; Huang, X.Y.; Yue, Z. Leucine-rich repeat kinase 2 (LRRK2)/PARK8 possesses GTPase activity that is altered in familial Parkinson’s disease R1441C/G mutants. J. Neurochem. 2007, 103, 238–247. [Google Scholar] [CrossRef] [Green Version]
  49. Heo, H.Y.; Kim, K.S.; Seol, W. Coordinate Regulation of Neurite Outgrowth by LRRK2 and Its Interactor, Rab5. Exp. Neurobiol. 2010, 19, 97–105. [Google Scholar] [CrossRef]
  50. 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]
  51. MacLeod, D.A.; Rhinn, H.; Kuwahara, T.; Zolin, A.; Di Paolo, G.; McCabe, B.D.; Marder, K.S.; Honig, L.S.; Clark, L.N.; Small, S.A.; et al. RAB7L1 interacts with LRRK2 to modify intraneuronal protein sorting and Parkinson’s disease risk. Neuron 2013, 77, 425–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. 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] [PubMed]
  53. Jeong, G.R.; Jang, E.H.; Bae, J.R.; Jun, S.; Kang, H.C.; Park, C.H.; Shin, J.H.; Yamamoto, Y.; Tanaka-Yamamoto, K.; Dawson, V.L.; et al. Dysregulated phosphorylation of Rab GTPases by LRRK2 induces neurodegeneration. Mol. Neurodegener. 2018, 13, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kim, K.S.; Marcogliese, P.C.; Yang, J.; Callaghan, S.M.; Resende, V.; Abdel-Messih, E.; Marras, C.; Visanji, N.P.; Huang, J.; Schlossmacher, M.G.; et al. Regulation of myeloid cell phagocytosis by LRRK2 via WAVE2 complex stabilization is altered in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2018, 115, E5164–E5173. [Google Scholar] [CrossRef] [Green Version]
  55. Kuwahara, T.; Inoue, K.; D′Agati, V.D.; Fujimoto, T.; Eguchi, T.; Saha, S.; Wolozin, B.; Iwatsubo, T.; Abeliovich, A. LRRK2 and RAB7L1 coordinately regulate axonal morphology and lysosome integrity in diverse cellular contexts. Sci. Rep. 2016, 6, 29945. [Google Scholar] [CrossRef]
  56. Liu, Z.; Bryant, N.; Kumaran, R.; Beilina, A.; Abeliovich, A.; Cookson, M.R.; West, A.B. LRRK2 phosphorylates membrane-bound Rabs and is activated by GTP-bound Rab7L1 to promote recruitment to the trans-Golgi network. Hum. Mol. Genet. 2018, 27, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Purlyte, E.; Dhekne, H.S.; Sarhan, A.R.; Gomez, R.; Lis, P.; Wightman, M.; Martinez, T.N.; Tonelli, F.; Pfeffer, S.R.; Alessi, D.R. Rab29 activation of the Parkinson’s disease-associated LRRK2 kinase. EMBO J. 2018, 37, 1–18. [Google Scholar] [CrossRef]
  58. Kalogeropulou, A.F.; Freemantle, J.B.; Lis, P.; Vides, E.G.; Polinski, N.K.; Alessi, D.R. Endogenous Rab29 does not impact basal or stimulated LRRK2 pathway activity. Biochem. J. 2020, 477, 4397–4423. [Google Scholar] [CrossRef]
  59. Gao, Y.; Wilson, G.R.; Stephenson, S.E.M.; Bozaoglu, K.; Farrer, M.J.; Lockhart, P.J. The emerging role of Rab GTPases in the pathogenesis of Parkinson’s disease. Mov. Disord 2018, 33, 196–207. [Google Scholar] [CrossRef]
  60. Weng, Y.H.; Chen, C.Y.; Lin, K.J.; Chen, Y.L.; Yeh, T.H.; Hsiao, I.T.; Chen, I.J.; Lu, C.S.; Wang, H.L. (R1441C) LRRK2 induces the degeneration of SN dopaminergic neurons and alters the expression of genes regulating neuronal survival in a transgenic mouse model. Exp. Neurol. 2016, 275 Pt 1, 104–115. [Google Scholar] [CrossRef]
  61. Xiong, Y.; Neifert, S.; Karuppagounder, S.S.; Liu, Q.; Stankowski, J.N.; Lee, B.D.; Ko, H.S.; Lee, Y.; Grima, J.C.; Mao, X.; et al. Robust kinase- and age-dependent dopaminergic and norepinephrine neurodegeneration in LRRK2 G2019S transgenic mice. Proc. Natl. Acad. Sci. USA 2018, 115, 1635–1640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Li, Y.; Liu, W.; Oo, T.F.; Wang, L.; Tang, Y.; Jackson-Lewis, V.; Zhou, C.; Geghman, K.; Bogdanov, M.; Przedborski, S.; et al. Mutant LRRK2(R1441G) BAC transgenic mice recapitulate cardinal features of Parkinson’s disease. Nat. Neurosci. 2009, 12, 826–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Tsika, E.; Kannan, M.; Foo, C.S.; Dikeman, D.; Glauser, L.; Gellhaar, S.; Galter, D.; Knott, G.W.; Dawson, T.M.; Dawson, V.L.; et al. Conditional expression of Parkinson’s disease-related R1441C LRRK2 in midbrain dopaminergic neurons of mice causes nuclear abnormalities without neurodegeneration. Neurobiol. Dis. 2014, 71, 345–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Seegobin, S.P.; Heaton, G.R.; Liang, D.; Choi, I.; Blanca Ramirez, M.; Tang, B.; Yue, Z. Progress in LRRK2-Associated Parkinson’s Disease Animal Models. Front. Neurosci. 2020, 14, 674. [Google Scholar] [CrossRef] [PubMed]
  65. Chang, E.E.S.; Ho, P.W.; Liu, H.F.; Pang, S.Y.; Leung, C.T.; Malki, Y.; Choi, Z.Y.; Ramsden, D.B.; Ho, S.L. LRRK2 mutant knock-in mouse models: Therapeutic relevance in Parkinson’s disease. Transl. Neurodegener. 2022, 11, 10. [Google Scholar] [CrossRef] [PubMed]
  66. Daher, J.P.; Volpicelli-Daley, L.A.; Blackburn, J.P.; Moehle, M.S.; West, A.B. Abrogation of alpha-synuclein-mediated dopaminergic neurodegeneration in LRRK2-deficient rats. Proc Natl Acad Sci U S A 2014, 111, 9289–9294. [Google Scholar] [CrossRef] [Green Version]
  67. Thevenet, J.; Pescini Gobert, R.; Hooft van Huijsduijnen, R.; Wiessner, C.; Sagot, Y.J. Regulation of LRRK2 expression points to a functional role in human monocyte maturation. PLoS ONE 2011, 6, e21519. [Google Scholar] [CrossRef]
  68. 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]
  69. Xu, E.; Boddu, R.; Abdelmotilib, H.A.; Sokratian, A.; Kelly, K.; Liu, Z.; Bryant, N.; Chandra, S.; Carlisle, S.M.; Lefkowitz, E.J.; et al. Pathological α-synuclein recruits LRRK2 expressing pro-inflammatory monocytes to the brain. Mol. Neurodegener. 2022, 17, 7. [Google Scholar] [CrossRef]
  70. Yamasaki, R.; Lu, H.; Butovsky, O.; Ohno, N.; Rietsch, A.M.; Cialic, R.; Wu, P.M.; Doykan, C.E.; Lin, J.; Cotleur, A.C.; et al. Differential roles of microglia and monocytes in the inflamed central nervous system. J. Exp. Med. 2014, 211, 1533–1549. [Google Scholar] [CrossRef]
  71. Liu, Z.; Lee, J.; Krummey, S.; Lu, W.; Cai, H.; Lenardo, M.J. The kinase LRRK2 is a regulator of the transcription factor NFAT that modulates the severity of inflammatory bowel disease. Nat. Immunol. 2011, 12, 1063–1070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hui, K.Y.; Fernandez-Hernandez, H.; Hu, J.; Schaffner, A.; Pankratz, N.; Hsu, N.Y.; Chuang, L.S.; Carmi, S.; Villaverde, N.; Li, X.; et al. Functional variants in the LRRK2 gene confer shared effects on risk for Crohn’s disease and Parkinson’s disease. Sci. Transl. Med. 2018, 10, eaai7795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta. Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Daneman, R.; Prat, A. The blood-brain barrier. Cold Spring Harb. Perspect. Biol. 2015, 7, a020412. [Google Scholar] [CrossRef] [Green Version]
  75. Allen, N.J.; Eroglu, C. Cell Biology of Astrocyte-Synapse Interactions. Neuron 2017, 96, 697–708. [Google Scholar] [CrossRef] [PubMed]
  76. Diaz-Castro, B.; Gangwani, M.R.; Yu, X.; Coppola, G.; Khakh, B.S. Astrocyte molecular signatures in Huntington’s disease. Sci. Transl. Med. 2019, 11, eaaw8546. [Google Scholar] [CrossRef] [Green Version]
  77. Habib, N.; McCabe, C.; Medina, S.; Varshavsky, M.; Kitsberg, D.; Dvir-Szternfeld, R.; Green, G.; Dionne, D.; Nguyen, L.; Marshall, J.L.; et al. Disease-associated astrocytes in Alzheimer’s disease and aging. Nat. Neurosci. 2020, 23, 701–706. [Google Scholar] [CrossRef]
  78. Miklossy, J.; Doudet, D.D.; Schwab, C.; Yu, S.; McGeer, E.G.; McGeer, P.L. Role of ICAM-1 in persisting inflammation in Parkinson disease and MPTP monkeys. Exp. Neurol. 2006, 197, 275–283. [Google Scholar] [CrossRef]
  79. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Zhang, X.; Zhang, R.; Nisa Awan, M.U.; Bai, J. The Mechanism and Function of Glia in Parkinson’s Disease. Front. Cell Neurosci. 2022, 16, 903469. [Google Scholar] [CrossRef]
  81. Booth, H.D.E.; Hirst, W.D.; Wade-Martins, R. The Role of Astrocyte Dysfunction in Parkinson’s Disease Pathogenesis. Trends Neurosci 2017, 40, 358–370. [Google Scholar] [CrossRef] [Green Version]
  82. Sidransky, E.; Nalls, M.A.; Aasly, J.O.; Aharon-Peretz, J.; Annesi, G.; Barbosa, E.R.; Bar-Shira, A.; Berg, D.; Bras, J.; Brice, A.; et al. Multicenter analysis of glucocerebrosidase mutations in Parkinson’s disease. N. Engl. J. Med. 2009, 361, 1651–1661. [Google Scholar] [CrossRef] [Green Version]
  83. Neumann, J.; Bras, J.; Deas, E.; O′Sullivan, S.S.; Parkkinen, L.; Lachmann, R.H.; Li, A.; Holton, J.; Guerreiro, R.; Paudel, R.; et al. Glucocerebrosidase mutations in clinical and pathologically proven Parkinson’s disease. Brain 2009, 132, 1783–1794. [Google Scholar] [CrossRef] [Green Version]
  84. 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] [PubMed]
  85. Bonet-Ponce, L.; Cookson, M.R. The endoplasmic reticulum contributes to lysosomal tubulation/sorting driven by LRRK2. Mol. Biol. Cell 2022, 33, ar124. [Google Scholar] [CrossRef] [PubMed]
  86. Guardia, C.M.; Farias, G.G.; Jia, R.; Pu, J.; Bonifacino, J.S. BORC Functions Upstream of Kinesins 1 and 3 to Coordinate Regional Movement of Lysosomes along Different Microtubule Tracks. Cell Rep. 2016, 17, 1950–1961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kumar, G.; Chawla, P.; Dhiman, N.; Chadha, S.; Sharma, S.; Sethi, K.; Sharma, M.; Tuli, A. RUFY3 links Arl8b and JIP4-Dynein complex to regulate lysosome size and positioning. Nat. Commun. 2022, 13, 1540. [Google Scholar] [CrossRef] [PubMed]
  88. Korolchuk, V.I.; Rubinsztein, D.C. Regulation of autophagy by lysosomal positioning. Autophagy 2011, 7, 927–928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Kluss, J.H.; Beilina, A.; Williamson, C.D.; Lewis, P.A.; Cookson, M.R.; Bonet-Ponce, L. Lysosomal positioning regulates Rab10 phosphorylation at LRRK2(+) lysosomes. Proc. Natl. Acad. Sci. USA 2022, 119, e2205492119. [Google Scholar] [CrossRef]
  90. Berndsen, K.; Lis, P.; Yeshaw, W.M.; Wawro, P.S.; Nirujogi, R.S.; Wightman, M.; Macartney, T.; Dorward, M.; Knebel, A.; Tonelli, F.; et al. PPM1H phosphatase counteracts LRRK2 signaling by selectively dephosphorylating Rab proteins. Elife 2019, 8, e50416. [Google Scholar] [CrossRef] [PubMed]
  91. Streubel-Gallasch, L.; Giusti, V.; Sandre, M.; Tessari, I.; Plotegher, N.; Giusto, E.; Masato, A.; Iovino, L.; Battisti, I.; Arrigoni, G.; et al. Parkinson’s Disease-Associated LRRK2 Interferes with Astrocyte-Mediated Alpha-Synuclein Clearance. Mol. Neurobiol. 2021, 58, 3119–3140. [Google Scholar] [CrossRef] [PubMed]
  92. Ramos-Gonzalez, P.; Mato, S.; Chara, J.C.; Verkhratsky, A.; Matute, C.; Cavaliere, F. Astrocytic atrophy as a pathological feature of Parkinson’s disease with LRRK2 mutation. NPJ Park. Dis. 2021, 7, 31. [Google Scholar] [CrossRef] [PubMed]
  93. Batiuk, M.Y.; Martirosyan, A.; Wahis, J.; de Vin, F.; Marneffe, C.; Kusserow, C.; Koeppen, J.; Viana, J.F.; Oliveira, J.F.; Voet, T.; et al. Identification of region-specific astrocyte subtypes at single cell resolution. Nat. Commun. 2020, 11, 1220. [Google Scholar] [CrossRef] [Green Version]
  94. Sofroniew, M.V. Astrocyte Reactivity: Subtypes, States, and Functions in CNS Innate Immunity. Trends Immunol. 2020, 41, 758–770. [Google Scholar] [CrossRef]
  95. McGeer, P.L.; Itagaki, S.; Boyes, B.E.; McGeer, E.G. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology 1988, 38, 1285–1291. [Google Scholar] [CrossRef] [PubMed]
  96. Vainchtein, I.D.; Chin, G.; Cho, F.S.; Kelley, K.W.; Miller, J.G.; Chien, E.C.; Liddelow, S.A.; Nguyen, P.T.; Nakao-Inoue, H.; Dorman, L.C.; et al. Astrocyte-derived interleukin-33 promotes microglial synapse engulfment and neural circuit development. Science 2018, 359, 1269–1273. [Google Scholar] [CrossRef] [Green Version]
  97. Hong, S.; Beja-Glasser, V.F.; Nfonoyim, B.M.; Frouin, A.; Li, S.; Ramakrishnan, S.; Merry, K.M.; Shi, Q.; Rosenthal, A.; Barres, B.A.; et al. Complement and microglia mediate early synapse loss in Alzheimer mouse models. Science 2016, 352, 712–716. [Google Scholar] [CrossRef] [Green Version]
  98. Nalls, M.A.; Blauwendraat, C.; Vallerga, C.L.; Heilbron, K.; Bandres-Ciga, S.; Chang, D.; Tan, M.; Kia, D.A.; Noyce, A.J.; Xue, A.; et al. Identification of novel risk loci, causal insights, and heritable risk for Parkinson’s disease: A meta-analysis of genome-wide association studies. Lancet Neurol. 2019, 18, 1091–1102. [Google Scholar] [CrossRef] [PubMed]
  99. Ransohoff, R.M.; El Khoury, J. Microglia in Health and Disease. Cold Spring Harb. Perspect. Biol. 2015, 8, a020560. [Google Scholar] [CrossRef] [Green Version]
  100. Ginhoux, F.; Lim, S.; Hoeffel, G.; Low, D.; Huber, T. Origin and differentiation of microglia. Front. Cell Neurosci. 2013, 7, 45. [Google Scholar] [CrossRef]
  101. Kabba, J.A.; Xu, Y.; Christian, H.; Ruan, W.; Chenai, K.; Xiang, Y.; Zhang, L.; Saavedra, J.M.; Pang, T. Microglia: Housekeeper of the Central Nervous System. Cell Mol. Neurobiol. 2018, 38, 53–71. [Google Scholar] [CrossRef]
  102. Choi, I.; Zhang, Y.; Seegobin, S.P.; Pruvost, M.; Wang, Q.; Purtell, K.; Zhang, B.; Yue, Z. Microglia clear neuron-released α-synuclein via selective autophagy and prevent neurodegeneration. Nat. Commun. 2020, 11, 1386. [Google Scholar] [CrossRef] [Green Version]
  103. Dutta, D.; Jana, M.; Majumder, M.; Mondal, S.; Roy, A.; Pahan, K. Selective targeting of the TLR2/MyD88/NF-κB pathway reduces α-synuclein spreading in vitro and in vivo. Nat. Commun. 2021, 12, 5382. [Google Scholar] [CrossRef] [PubMed]
  104. Dresselhaus, E.C.; Meffert, M.K. Cellular Specificity of NF-kappaB Function in the Nervous System. Front. Immunol. 2019, 10, 1043. [Google Scholar] [CrossRef] [Green Version]
  105. Barrett, J.C.; Hansoul, S.; Nicolae, D.L.; Cho, J.H.; Duerr, R.H.; Rioux, J.D.; Brant, S.R.; Silverberg, M.S.; Taylor, K.D.; Barmada, M.M.; et al. Genome-wide association defines more than 30 distinct susceptibility loci for Crohn’s disease. Nat. Genet. 2008, 40, 955–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Zhang, Y.M.; Zhou, X.J.; Cheng, F.J.; Qi, Y.Y.; Hou, P.; Zhao, M.H.; Zhang, H. Autophagy-related gene LRRK2 is likely a susceptibility gene for systemic lupus erythematosus in northern Han Chinese. Oncotarget 2017, 8, 13754–13761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Fan, Y.; Howden, A.J.M.; Sarhan, A.R.; Lis, P.; Ito, G.; Martinez, T.N.; Brockmann, K.; Gasser, T.; Alessi, D.R.; Sammler, E.M. Interrogating Parkinson’s disease LRRK2 kinase pathway activity by assessing Rab10 phosphorylation in human neutrophils. Biochem. J. 2018, 475, 23–44. [Google Scholar] [CrossRef] [Green Version]
  108. Gardet, A.; Benita, Y.; Li, C.; Sands, B.E.; Ballester, I.; Stevens, C.; Korzenik, J.R.; Rioux, J.D.; Daly, M.J.; Xavier, R.J.; et al. LRRK2 is involved in the IFN-gamma response and host response to pathogens. J. Immunol. 2010, 185, 5577–5585. [Google Scholar] [CrossRef] [Green Version]
  109. Lee, H.; Flynn, R.; Sharma, I.; Haberman, E.; Carling, P.J.; Nicholls, F.J.; Stegmann, M.; Vowles, J.; Haenseler, W.; Wade-Martins, R.; et al. LRRK2 Is Recruited to Phagosomes and Co-recruits RAB8 and RAB10 in Human Pluripotent Stem Cell-Derived Macrophages. Stem Cell Rep. 2020, 14, 940–955. [Google Scholar] [CrossRef]
  110. Ahmadi Rastegar, D.; Hughes, L.P.; Perera, G.; Keshiya, S.; Zhong, S.; Gao, J.; Halliday, G.M.; Schüle, B.; Dzamko, N. Effect of LRRK2 protein and activity on stimulated cytokines in human monocytes and macrophages. NPJ Park. Dis. 2022, 8, 34. [Google Scholar] [CrossRef]
  111. Brockmann, K.; Apel, A.; Schulte, C.; Schneiderhan-Marra, N.; Pont-Sunyer, C.; Vilas, D.; Ruiz-Martinez, J.; Langkamp, M.; Corvol, J.C.; Cormier, F.; et al. Inflammatory profile in LRRK2-associated prodromal and clinical PD. J. Neuroinflammation 2016, 13, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Brockmann, K.; Schulte, C.; Schneiderhan-Marra, N.; Apel, A.; Pont-Sunyer, C.; Vilas, D.; Ruiz-Martinez, J.; Langkamp, M.; Corvol, J.C.; Cormier, F.; et al. Inflammatory profile discriminates clinical subtypes in LRRK2-associated Parkinson’s disease. Eur. J. Neurol. 2017, 24, 427-e426. [Google Scholar] [CrossRef] [PubMed]
  113. Langston, R.G.; Beilina, A.; Reed, X.; Kaganovich, A.; Singleton, A.B.; Blauwendraat, C.; Gibbs, J.R.; Cookson, M.R. Association of a common genetic variant with Parkinson’s disease is mediated by microglia. Sci. Transl. Med. 2022, 14, eabp8869. [Google Scholar] [CrossRef] [PubMed]
  114. Russo, I.; Berti, G.; Plotegher, N.; Bernardo, G.; Filograna, R.; Bubacco, L.; Greggio, E. Leucine-rich repeat kinase 2 positively regulates inflammation and down-regulates NF-κB p50 signaling in cultured microglia cells. J. Neuroinflammation 2015, 12, 230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Marín, I.; van Egmond, W.N.; van Haastert, P.J. The Roco protein family: A functional perspective. Faseb. J. 2008, 22, 3103–3110. [Google Scholar] [CrossRef] [Green Version]
  116. Bosgraaf, L.; Russcher, H.; Smith, J.L.; Wessels, D.; Soll, D.R.; Van Haastert, P.J. A novel cGMP signalling pathway mediating myosin phosphorylation and chemotaxis in Dictyostelium. EMBO J. 2002, 21, 4560–4570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Choi, I.; Kim, B.; Byun, J.W.; Baik, S.H.; Huh, Y.H.; Kim, J.H.; Mook-Jung, I.; Song, W.K.; Shin, J.H.; Seo, H.; et al. LRRK2 G2019S mutation attenuates microglial motility by inhibiting focal adhesion kinase. Nat. Commun. 2015, 6, 8255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Ma, B.; Xu, L.; Pan, X.; Sun, L.; Ding, J.; Xie, C.; Koliatsos, V.E.; Cai, H. LRRK2 modulates microglial activity through regulation of chemokine (C-X3-C) receptor 1 -mediated signalling pathways. Hum. Mol. Genet. 2016, 25, 3515–3523. [Google Scholar] [CrossRef] [Green Version]
  119. Ogata, J.; Hirao, K.; Nishioka, K.; Hayashida, A.; Li, Y.; Yoshino, H.; Shimizu, S.; Hattori, N.; Imai, Y. A Novel LRRK2 Variant p.G2294R in the WD40 Domain Identified in Familial Parkinson’s Disease Affects LRRK2 Protein Levels. Int. J. Mol. Sci. 2021, 22, 3708. [Google Scholar] [CrossRef]
  120. Takanashi, M.; Funayama, M.; Matsuura, E.; Yoshino, H.; Li, Y.; Tsuyama, S.; Takashima, H.; Nishioka, K.; Hattori, N. Isolated nigral degeneration without pathological protein aggregation in autopsied brains with LRRK2 p.R1441H homozygous and heterozygous mutations. Acta Neuropathol. Commun. 2018, 6, 105. [Google Scholar] [CrossRef]
  121. Tezuka, T.; Taniguchi, D.; Sano, M.; Shimada, T.; Oji, Y.; Tsunemi, T.; Ikeda, A.; Li, Y.; Yoshino, H.; Ogata, J.; et al. Pathophysiological evaluation of the LRRK2 G2385R risk variant for Parkinson’s disease. NPJ Park. Dis. 2022, 8, 97. [Google Scholar] [CrossRef]
  122. Drieu, A.; Du, S.; Storck, S.E.; Rustenhoven, J.; Papadopoulos, Z.; Dykstra, T.; Zhong, F.; Kim, K.; Blackburn, S.; Mamuladze, T.; et al. Parenchymal border macrophages regulate the flow dynamics of the cerebrospinal fluid. Nature 2022, 611, 585–593. [Google Scholar] [CrossRef] [PubMed]
  123. Ishida, K.; Yamada, K.; Nishiyama, R.; Hashimoto, T.; Nishida, I.; Abe, Y.; Yasui, M.; Iwatsubo, T. Glymphatic system clears extracellular tau and protects from tau aggregation and neurodegeneration. J. Exp. Med. 2022, 219, e20211275. [Google Scholar] [CrossRef]
  124. 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]
  125. Dodson, M.W.; Leung, L.K.; Lone, M.; Lizzio, M.A.; Guo, M. Novel ethyl methanesulfonate (EMS)-induced null alleles of the Drosophila homolog of LRRK2 reveal a crucial role in endolysosomal functions and autophagy in vivo. Dis Model Mech. 2014, 7, 1351–1363. [Google Scholar] [CrossRef]
  126. Giaime, E.; Tong, Y.; Wagner, L.K.; Yuan, Y.; Huang, G.; Shen, J. Age-Dependent Dopaminergic Neurodegeneration and Impairment of the Autophagy-Lysosomal Pathway in LRRK-Deficient Mice. Neuron 2017, 96, 796–807.e796. [Google Scholar] [CrossRef] [Green Version]
  127. Inoshita, T.; Liu, J.-Y.; Taniguchi, D.; Ishii, R.; Shiba-Fukushima, K.; Hattori, N.; Imai, Y. Parkinson disease-associated Leucine-rich repeat kinase regulates UNC-104-dependent axonal transport of Arl8-positive vesicles in Drosophila. iScience 2022, 25, 105476. [Google Scholar] [CrossRef] [PubMed]
  128. Whiffin, N.; Armean, I.M.; Kleinman, A.; Marshall, J.L.; Minikel, E.V.; Goodrich, J.K.; Quaife, N.M.; Cole, J.B.; Wang, Q.; Karczewski, K.J.; et al. The effect of LRRK2 loss-of-function variants in humans. Nat. Med. 2020, 26, 869–877. [Google Scholar] [CrossRef] [PubMed]
  129. Jennings, D.; Huntwork-Rodriguez, S.; Henry, A.G.; Sasaki, J.C.; Meisner, R.; Diaz, D.; Solanoy, H.; Wang, X.; Negrou, E.; Bondar, V.V.; et al. Preclinical and clinical evaluation of the LRRK2 inhibitor DNL201 for Parkinson’s disease. Sci. Transl. Med. 2022, 14, eabj2658. [Google Scholar] [CrossRef]
  130. Kingwell, K. LRRK2-targeted Parkinson disease drug advances into phase III. Nat. Rev. Drug Discov. 2022, 22, 3–5. [Google Scholar] [CrossRef]
Figure 1. Domain structure of LRRK2. Sites of representative pathological mutations are indicated [7]. G2385R is considered a risk variant in Asian races [7]. LRRK2, leucine rich-repeat kinase 2; ARM, armadillo domain; ANK, ankyrin repeat domain; LRR, leucine-rich repeat domain; ROC, Ras of complex protein domain; COR, carboxyl-terminal of ROC domain; WD40, WD 40 domain; aa, amino acids.
Figure 1. Domain structure of LRRK2. Sites of representative pathological mutations are indicated [7]. G2385R is considered a risk variant in Asian races [7]. LRRK2, leucine rich-repeat kinase 2; ARM, armadillo domain; ANK, ankyrin repeat domain; LRR, leucine-rich repeat domain; ROC, Ras of complex protein domain; COR, carboxyl-terminal of ROC domain; WD40, WD 40 domain; aa, amino acids.
Biomolecules 13 00178 g001
Figure 2. Overview of LRRK2 research, research problems, and issues to be resolved. In humans, risk SNPs have been found in Crohn’s disease as well as Parkinson’s disease. Currently, the impact of Crohn’s disease-associated SNPs on LRRK2 function is not clear [71,72]. Many in vivo models do not exhibit the pathological changes observed in Parkinson’s disease. However, LRRK2 knockout rodent models are an excellent tool to study the physiological function of LRRK2. In vitro models are superior for analysis at the organelle and molecular levels but have the disadvantage of not being able to reproduce the normal aging process and the brain environment. The results obtained from each of these models should be assessed in light of these considerations. The studies to which this review refers are indicated in red [11,12,13,14,73]. C. elegans and Drosophila models are excellent for molecular genetic analysis but will not be mentioned here. EV, extracellular vesicle.
Figure 2. Overview of LRRK2 research, research problems, and issues to be resolved. In humans, risk SNPs have been found in Crohn’s disease as well as Parkinson’s disease. Currently, the impact of Crohn’s disease-associated SNPs on LRRK2 function is not clear [71,72]. Many in vivo models do not exhibit the pathological changes observed in Parkinson’s disease. However, LRRK2 knockout rodent models are an excellent tool to study the physiological function of LRRK2. In vitro models are superior for analysis at the organelle and molecular levels but have the disadvantage of not being able to reproduce the normal aging process and the brain environment. The results obtained from each of these models should be assessed in light of these considerations. The studies to which this review refers are indicated in red [11,12,13,14,73]. C. elegans and Drosophila models are excellent for molecular genetic analysis but will not be mentioned here. EV, extracellular vesicle.
Biomolecules 13 00178 g002
Figure 3. Reported roles of LRRK2 in astrocytes. (A) LRRK2 accumulates in membrane-damaged lysosomes and phosphorylates downstream Rab10, causing lysosomal tubulation and budding [23]. (B) Microtubule-dependent kinesin–ARL8B and dynein–JIP4 motors move the lysosomes to the peripheral and perinuclear regions, respectively [89]. Rab10 is preferentially phosphorylated in the perinuclear lysosomes by LRRK2 (A), whereas Rab12 is phosphorylated in both the perinuclear and the peripheral lysosomes. (C) Impairment of chaperone-mediated autophagy by LRRK2 G2019S leads to α-synuclein accumulation in astrocytes [12]. Astrocyte-derived pathogenic α-synuclein is transported into dopaminergic neurons [12]. On the other hand, phagocytic defects of α-synuclein fibrils due to decreased annexin A2 by LRRK2 G2019S are also suggested [91]. (D) Compared to normal astrocytes, LRRK2 G2019S astrocytes have smaller MVBs, along with the accumulation of phospho-S129 α-synuclein and lower EV release. The alteration of EV biogenesis in LRRK2 G2019S astrocytes induces dendritic shortening and cell death in dopaminergic neurons [13]. MVB, multivesicular body; EV, extracellular vesicle; P, phosphorylation.
Figure 3. Reported roles of LRRK2 in astrocytes. (A) LRRK2 accumulates in membrane-damaged lysosomes and phosphorylates downstream Rab10, causing lysosomal tubulation and budding [23]. (B) Microtubule-dependent kinesin–ARL8B and dynein–JIP4 motors move the lysosomes to the peripheral and perinuclear regions, respectively [89]. Rab10 is preferentially phosphorylated in the perinuclear lysosomes by LRRK2 (A), whereas Rab12 is phosphorylated in both the perinuclear and the peripheral lysosomes. (C) Impairment of chaperone-mediated autophagy by LRRK2 G2019S leads to α-synuclein accumulation in astrocytes [12]. Astrocyte-derived pathogenic α-synuclein is transported into dopaminergic neurons [12]. On the other hand, phagocytic defects of α-synuclein fibrils due to decreased annexin A2 by LRRK2 G2019S are also suggested [91]. (D) Compared to normal astrocytes, LRRK2 G2019S astrocytes have smaller MVBs, along with the accumulation of phospho-S129 α-synuclein and lower EV release. The alteration of EV biogenesis in LRRK2 G2019S astrocytes induces dendritic shortening and cell death in dopaminergic neurons [13]. MVB, multivesicular body; EV, extracellular vesicle; P, phosphorylation.
Biomolecules 13 00178 g003
Figure 4. Reported roles of LRRK2 in microglia. (A) LRRK2 negatively regulates the migration of microglia through the suppression of focal adhesion kinase (FAK) via phosphorylation [117]. (B) LRRK2 phosphorylates and stabilizes WAVE2, inducing the positive regulation of microglial phagocytosis [54]. (C) LRRK2 kinase activity is required for the mobilization of Rab8a and Rab10 to phagosomes, which leads to phagosome maturation [109]. (D) LRRK2 accumulates and recruits Rab8 and Rab10 under lysosomal stress in a Rab29-dependent manner, which causes the extracellular release of the lysosomal contents [22]. (E) NFATc2 is phosphorylated by LRRK2 in microglia exposed to α-synuclein and then translocated to the nucleus for transcriptional regulation [14]. (F) Nuclear migration of NFATc2 is inhibited in LRRK2 G2019S iPS cell-derived microglia [15]. (G) LRRK2 pathogenic mutants sequester Rab8 to the lysosomes via phosphorylation, directing the transferrin-mediated iron endocytosis pathway from lysosomal recycling to degradation [16]. As a result, iron deposition is accelerated [16]. (H) Microglia that have taken up excess amounts of α-synuclein fibrils transport these fibrils to other microglia through F-actin-associated nanotubes [31]. P, phosphorylation.
Figure 4. Reported roles of LRRK2 in microglia. (A) LRRK2 negatively regulates the migration of microglia through the suppression of focal adhesion kinase (FAK) via phosphorylation [117]. (B) LRRK2 phosphorylates and stabilizes WAVE2, inducing the positive regulation of microglial phagocytosis [54]. (C) LRRK2 kinase activity is required for the mobilization of Rab8a and Rab10 to phagosomes, which leads to phagosome maturation [109]. (D) LRRK2 accumulates and recruits Rab8 and Rab10 under lysosomal stress in a Rab29-dependent manner, which causes the extracellular release of the lysosomal contents [22]. (E) NFATc2 is phosphorylated by LRRK2 in microglia exposed to α-synuclein and then translocated to the nucleus for transcriptional regulation [14]. (F) Nuclear migration of NFATc2 is inhibited in LRRK2 G2019S iPS cell-derived microglia [15]. (G) LRRK2 pathogenic mutants sequester Rab8 to the lysosomes via phosphorylation, directing the transferrin-mediated iron endocytosis pathway from lysosomal recycling to degradation [16]. As a result, iron deposition is accelerated [16]. (H) Microglia that have taken up excess amounts of α-synuclein fibrils transport these fibrils to other microglia through F-actin-associated nanotubes [31]. P, phosphorylation.
Biomolecules 13 00178 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Iseki, T.; Imai, Y.; Hattori, N. Is Glial Dysfunction the Key Pathogenesis of LRRK2-Linked Parkinson’s Disease? Biomolecules 2023, 13, 178. https://doi.org/10.3390/biom13010178

AMA Style

Iseki T, Imai Y, Hattori N. Is Glial Dysfunction the Key Pathogenesis of LRRK2-Linked Parkinson’s Disease? Biomolecules. 2023; 13(1):178. https://doi.org/10.3390/biom13010178

Chicago/Turabian Style

Iseki, Tatou, Yuzuru Imai, and Nobutaka Hattori. 2023. "Is Glial Dysfunction the Key Pathogenesis of LRRK2-Linked Parkinson’s Disease?" Biomolecules 13, no. 1: 178. https://doi.org/10.3390/biom13010178

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