Skip to main content

Immune cell metabolic dysfunction in Parkinson’s disease

Abstract

Parkinson’s disease (PD) is a multi-system disorder characterized histopathologically by degeneration of dopaminergic neurons in the substantia nigra pars compacta. While the etiology of PD remains multifactorial and complex, growing evidence suggests that cellular metabolic dysfunction is a critical driver of neuronal death. Defects in cellular metabolism related to energy production, oxidative stress, metabolic organelle health, and protein homeostasis have been reported in both neurons and immune cells in PD. We propose that these factors act synergistically in immune cells to drive aberrant inflammation in both the CNS and the periphery in PD, contributing to a hostile inflammatory environment which renders certain subsets of neurons vulnerable to degeneration. This review highlights the overlap between established neuronal metabolic deficits in PD with emerging findings in central and peripheral immune cells. By discussing the rapidly expanding literature on immunometabolic dysfunction in PD, we aim to draw attention to potential biomarkers and facilitate future development of immunomodulatory strategies to prevent or delay the progression of PD.

Introduction

Parkinson’s disease (PD) is the second most common neurodegenerative disorder after Alzheimer’s disease (AD), affecting millions worldwide [1]. The cardinal motor symptoms of PD, which include tremor, rigidity, akinesia, and postural instability result from degeneration of dopaminergic neurons (DANs) within the substantia nigra pars compacta (SNpc) [2]. The exact cause of neuronal death in PD is not fully understood; however, growing evidence points to bioenergetic insufficiency within DANs as a penultimate step prior to degeneration [3,4,5]. Indeed, many of the known contributors to PD such as normal aging, genetic predisposition, and exposure to environmental toxins converge on metabolic dysfunction [6, 7]. Importantly, recent evidence suggests that many defects in energy homeostasis reported in PD DANs, including poor mitochondrial and lysosomal health, are also observed in immune cells [8,9,10]. Therefore, immune cell metabolic dysfunction in PD may have broad scientific and clinical implications ranging from mechanistic understanding of disease progression to biomarkers for improving patient care. In this review, we will outline the current understanding of immune cell metabolic dysfunction in PD including changes in glycolytic activity, mitochondrial and lysosomal deficits, and disrupted homeostasis of fundamental energetic substrates. We will highlight areas of overlap between the deficits reported in immune cells and those previously reported in neurons, and we will identify gaps in the literature which future research should seek to address.

Glycolysis

Glycolysis is one of the most fundamental pathways of energy production for mammalian cells [11], and glucose utilization is known to be disrupted in the brains of individuals with idiopathic PD (iPD) compared to age-matched non-PD controls [12]. Glycolysis is a highly conserved process across many domains of life that occurs in the cytoplasm, and it consists of a series of reactions that extract energy from glucose [13]. This requires an energy investment phase which consumes ATP to activate glucose, then an energy payoff phase that generates ATP and NADH through substrate-level phosphorylation and redox reactions [13]. The end product of glycolysis is pyruvate, which can be oxidized to Acetyl-CoA and used to generate electron carriers such as NADH and FADH2 in the Krebs cycle [14]. These energy-rich molecules are used in oxidative phosphorylation (OXPHOS) at the inner mitochondrial membrane to generate large amounts of ATP; in contrast to glycolysis alone which produces a net of 2 ATP per glucose, OXPHOS in combination with glycolysis increases the yield to 32–34 ATP per glucose molecule [15]. We direct the reader to other excellent reviews for more detailed explanations on these processes [13, 16, 17]. Growing evidence suggests that neurodegenerative diseases, including PD, are linked to metabolic reprogramming in immune cells that shifts the balance in cellular decision-making between glycolysis and OXPHOS [18, 19]. This could potentially contribute to bioenergetic deficits in PD that prime immune cells to respond with hyperinflammatory responses to stimulation [20].

A wealth of evidence suggests that glycolytic pathways are dysregulated in preclinical models of PD. Induced pluripotent stem cell (iPSC)-derived neurons harboring PD-associated mutations display marked dysregulation in glycolytic gene expression and reduced levels of glucose, ATP, and pyruvate [21]. In addition, treatment with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), a compound frequently used to model PD-like degeneration in animals, causes reductions in glycolytic activity, ATP, and pyruvate in human neuroblastoma cells [22]. Dysregulated glycolysis has also been observed in immune cells in various preclinical PD models (Fig. 1). Mouse microglia treated with α-synuclein preformed fibrils (PFFs) display increased glycolytic activity and capacity with simultaneous reductions in mitochondrial basal respiration and ATP production [23]. Furthermore, inhibition of glycolytic activity has been shown to ameliorate microglial activation and TH+ neuronal loss in both lipopolysaccharide (LPS)- and MPTP-induced mouse models of PD-like nigrostriatal degeneration [24]. These findings suggest that increased glycolysis in microglia is associated with a pro-inflammatory state, potentially exacerbating the vulnerability of nearby neurons and contributing to eventual neuronal degeneration in PD [20]. Supporting this, reports indicate that primary microglia from mice with heterozygous knockout (KO) of Clk1, a gene encoding a mitochondrial hydroxylase, exhibit elevated production of proinflammatory cytokines and reactive oxygen species (ROS) [25]. Consistent with this finding, Clk1−/− BV2 microglia display a hyperinflammatory phenotype and a shift towards aerobic glycolysis that is abrogated by pharmacologic glycolysis inhibitors [25]. Although not of immune cell origin, recent evidence suggests that in PD, astrocytes which are capable of mounting inflammatory responses [26, 27], may show similar pro-glycolytic reprogramming. Specifically, astrocytes derived from PD patients carrying mutations in leucine-rich repeat kinase 2 (LRRK2) display significantly increased basal and compensatory glycolytic activity compared to controls [28], and LRRK2 PD astrocytes produce higher levels of IL-6 than control astrocytes after exposure to tumor necrosis factor (TNF) and interleukin-1 beta (IL-1β) [29]. In summary, these studies suggest a role for CNS metabolic deficits in immune and astroglia activation within the context of PD neurodegeneration, making it a promising target for future clinical development.

Similar findings have been reported in the peripheral immune system, with peripheral blood mononuclear cells (PBMCs) from individuals with iPD exhibiting increased glycolytic activity, capacity, and reserve relative to controls [30]. Intriguingly, these same glycolytic changes are observed in PBMCs from those with rapid eye movement (REM) sleep behavior disorder (RBD), a group generally considered to approximate prodromal PD due to their high likelihood of converting to PD or related synucleinopathies [31]. RBD is characterized by loss of muscle atonia during REM sleep and the physical acting out of dreams that are often intense or violent [32]. Approximately 80% of individuals with RBD will develop a neurodegenerative disease within 10.5 years of RBD diagnosis, and the plurality (43%) of those who convert will develop PD [31]. Therefore, the finding that peripheral immune cell glycolytic activity is dysregulated prior to the motor stage of PD has significant implications for understanding disease progression. Increased glycolytic activity in PBMCs may contribute to a hyperinflammatory response in the periphery, mirroring the findings in microglia [24, 25] (Fig. 1). Indeed, plasma levels of the cytokine TNF are increased in RBD patients [33], and early-stage PD patients show elevated serum levels of IL-1β34. Moreover, pharmacologic inhibition of glycolysis was shown to attenuate inflammatory responses in several mouse models of peripheral inflammation including irritable bowel disease, rheumatoid arthritis, and acute respiratory distress syndrome [35]. In sum, these results suggest that glycolytic activity is increased in both central and peripheral immune cells in individuals with non-motor features of PD, and that targeting glycolysis may be an effective strategy to mitigate aberrant inflammation in the prodromal stage of PD.

Fig. 1
figure 1

Proposed bidirectional and cyclical relationship between altered immune activation and glycolytic reprogramming in PD. This figure presents a schematic overview of factors which contribute to dysregulated immune function and increased glycolytic activity in the context of PD. Exposure to immune stimuli causes pro-glycolytic reprogramming in both central and peripheral immune cell subtypes. Interventions and underlying deficits which increase cellular utilization of glycolytic pathways have been shown to prime enhanced pro-inflammatory responses to secondary insults. These elements combine to create a feed-forwards cycle that contribute to dopaminergic neuron loss in PD. Abbreviations: lipopolysaccharide (LPS); tyrosine hydroxylase (TH). Created with BioRender.com

Mitochondrial dysfunction

Mitochondria are rod-shaped organelles surrounded by a double membrane and are colloquially known as the “powerhouse of the cell” due to their crucial role in generating energy through OXPHOS and the electron transport chain [17]. As mentioned above, OXPHOS occurs at the inner mitochondrial membrane and constitutes the largest source of energy production in eukaryotic cells [36]. This process involves a series of redox reactions where electrons are transferred from high-energy carriers including NADH and FADH2 through protein complexes (I-IV) and mobile carriers like ubiquinone and cytochrome c [17, 37]. As electrons move through the electron transport chain (ETC), the energy released at each step is used to actively pump protons into the intermembrane space and generate an electrochemical gradient [17]. Protons are allowed to flow back into the matrix through a molecular turbine known as ATP synthase (complex V), which harnesses the proton motive force to phosphorylate ADP to ATP [37]. The final step of the ETC involves molecular oxygen acting as the terminal electron acceptor [37]; this represents a crucial difference from glycolysis which can occur in the absence of oxygen [13, 17]. For additional details on mitochondrial function and dysfunction, we direct the reader to comprehensive reviews on mitochondrial energy production [17, 37].

Genetic findings have extensively implicated mitochondrial dysfunction in PD pathogenesis. For example, Parkin, an E3 ubiquitin ligase encoded by the PRKN gene, ubiquitinates damaged mitochondria to induce degradation and mitophagy [38], and mutations in Parkin are the most common cause of autosomal recessive early-onset PD [38]. Parkin deficiency compromises mitochondrial function in Drosophila [39], mice [40], and iPSC-derived neurons from individuals with Parkin PD [41]. Further strengthening the connection between mitochondria and PD, pathogenic variants in the Parkin regulatory protein PTEN-induced kinase 1 (PINK1) form the second most common cause of autosomal recessive PD [42]. PINK1 recruits Parkin to damaged mitochondria and phosphorylates Parkin along with its ubiquitin conjugates on the outer mitochondrial membrane to regulate autophagic clearance of damaged mitochondria [43]. Sporadic forms of PD have also been linked to deficits in mitochondrial health, with iPD neurons showing decreased expression of mitochondrial complexes I–V [44] (Fig. 2). In addition, fibroblasts from individuals with iPD display decreased mitochondrial connectivity, reduced branching, decreased mean mitochondrial size, and increased fragmentation [45]. Mitochondrial deficits in non-immune cells have been extensively reviewed by others, and we refer the reader to these articles for a more comprehensive breakdown of the subject [46, 47] in addition to detailed reviews discussing the known genetic risk factors for PD [48, 49]. Here, we focus on the growing evidence implicating mitochondrial dysfunction in immune cells as a driver of PD pathogenesis.

Electron transport chain defects

The ETC is an essential component of effective mitochondrial function, and growing evidence suggests that disrupted ETC function is linked to PD pathogenesis. For instance, SNpc samples from individuals with iPD show significantly decreased activity of mitochondrial complex I relative to controls [50], while PD astrocytes exhibit reduced expression of complexes I-V [51]. While the effects of reduced ETC activity and energy production may directly compromise neuronal survival, recent findings also suggest that ETC dysfunction in microglia may contribute to PD through increasing neuroinflammation (Fig. 2). Work by Sarkar et al. demonstrated that LPS-primed primary mouse microglia significantly upregulated IL-1β secretion in a dose-dependent manner after treatment with the mitochondrial complex I inhibitors rotenone and tebufenpyrad [52]. Rotenone is a pesticide epidemiologically linked to increased PD risk [53, 54]; however it did not increase inflammatory cytokine secretion in unprimed microglia in this study [52]. This suggests that inhibition of mitochondrial respiration is not sufficient to drive inflammation but may potentiate the proinflammatory response to a secondary insult. Importantly, secretion of IL-1β, but not TNF, was significantly affected by rotenone treatment, indicating that complex I inhibition in primed microglia likely acts through inflammasome signaling. To confirm this, pre-treatment with an NLRP3 inflammasome inhibitor completely blocked the rotenone-induced increase in IL-1β secretion but did not alter secretion of IL-12, as IL-12 processing is independent of NLRP3 activation [52]. Similar findings were reported by Won et al. who demonstrated that rotenone treatment of mouse bone marrow-derived macrophages (BMDMs) primed NLRP3 inflammasome activation [55]. Importantly, co-treatment with ATP and rotenone resulted in NLRP3-dependent caspase-1 activation whereas rotenone treatment alone had no significant effect [55]. Overall, these results point to a mechanism whereby baseline reductions in mitochondrial complex I activity in PD contribute to excessive inflammation through upregulation of inflammasome activity.

In a recent study, peripheral blood lymphocytes from individuals with iPD were found to have diminished mitochondrial complex II activity relative to age-matched controls [56]. The activity of complexes I, III, and IV was similar across PD lymphocytes and controls; however, activity of complex I was significantly decreased in PD platelets [56]. On the other hand, Müftüoglu et al. observed reduced activity of complexes I and IV in iPD PBMCs, in addition to decreased complex I activity in Parkin PD relative to controls [57]. Discrepancies between these studies may arise from the use of PBMCs versus specifically lymphocytes. Therefore, further study is necessary to develop a consensus on which ETC complexes are disrupted in peripheral immune cells in PD and to what extent. Interestingly, iPD PBMCs display a significant upregulation in mRNA expression of the respiratory chain subunits that compose complexes I–V [58]. This may represent a compensatory response to ETC deficits in these cells, suggesting that strategies to rescue ETC activity could have therapeutic potential for restoring peripheral immune function in PD.

Fig. 2
figure 2

Known facets of mitochondrial dysfunction in PD patients and preclinical models which converge on immune dysregulation. Widespread mitochondrial dysfunction is established to occur within neurons in the etiology of PD, including deficits in electron transport chain activity, respiration, ROS homeostasis, and mitochondrial membrane potential. Several of these deficits have been reflected in immune cells derived from PD patients and preclinical models of PD-like degeneration. Mitochondrial dysfunction can increase inflammasome-mediated cytokine secretion and pro-inflammatory response to aggregated α-synuclein. Inflammation promotes increased central-peripheral immune crosstalk and the formation of an inflammatory milieu that is hostile to vulnerable neurons. ROS from dopamine metabolism may render dopaminergic neurons particularly susceptible to degeneration, and a “second hit” of mitochondrial dysfunction overwhelms their capability to survive an inflammatory, neurotoxic environment. Abbreviations: mitochondrial membrane potential (MMP); reactive oxygen species (ROS); electron transport chain (ETC); inducible nitric oxide synthase (iNOS); cyclooxygenase (COX); tumor necrosis factor (TNF); lipopolysaccharide (LPS). Created with BioRender.com.

Changes in mitochondrial respiration

Deficits in mitochondrial respiration and oxygen consumption have been a focus of PD research since the discovery that inhibition of mitochondrial complex I causes parkinsonian symptoms in humans and animal models [59, 60]. Basal respiration and passive proton leak across the mitochondrial membrane are increased in iPSC-derived neural cells from PINK1 PD patients, while neural cells from those with LRRK2 G2019S PD show significantly lower basal respiration relative to controls [61]. Meanwhile, PINK1 KO and chronic MPTP treatment in aged mice have been shown to significantly reduce the maximal respiratory rate of mitochondria from the striatum [62, 63].

More recent studies have sought to determine if PD is associated with respiratory defects in CNS immune cells, given the possibility that energy insufficiency could contribute to inflammatory changes. Lu et al. reported that microglia treated with α-synuclein PFFs underwent significant metabolic reprogramming, demonstrating a reduction in basal and maximal respiratory capacity measured by oxygen consumption rate (OCR) [23]. Concurrently, PFF treatment caused a decrease in microglial ATP production and an increase in extracellular acidification rate (ECAR), indicating a shift from OXPHOS towards energy-inefficient glycolysis. These findings indicate that exposure to immunogenic stimuli such as misfolded α-synuclein can drive reductions in mitochondrial respiration in microglia. Moreover, impaired OXPHOS in microglia is accompanied by increased microglial reactivity, with PFF treatment also causing increased expression of TNF and IL-1β23 (Fig. 2). Microglial metabolic deficits were rescued by pre-treatment with capsaicin [23], a TRPV1 agonist which has previously been shown to inhibit oxidative stress and attenuate DAN degeneration in MPTP-treated mice [64]. These findings were supported by another group who reported that inhibition of mitochondrial complex V in IL-4 treated microglia caused a downregulation of anti-inflammatory markers, leading to an overall more pro-inflammatory phenotype [65]. Sonninen et al. reported that iPSC-derived astrocytes from individuals with LRRK2 PD displayed decreased maximal and spare respiration [29], reflecting a similar pattern of metabolic changes observed by Lu et al. involving α-synuclein PFF treated microglia [23]. Jointly, these reports suggest that preventing maladaptive metabolic reprogramming of inflammatory cells in the CNS and preserving mitochondrial respiration may help to combat neuroinflammation in PD.

Defects in mitochondrial respiration in some peripheral cell types like fibroblasts [66, 67] have been documented in PD, but a consensus has yet to develop on whether this extends to peripheral immune cells. Smith et al. reported that PBMCs from individuals with RBD and iPD showed no significant differences in basal respiration, maximal respiration, spare capacity, ATP production, and nonmitochondrial respiration relative to controls [30]. In contrast, Schirinzi et al. observed increased maximal and spare respiratory capacity in iPD PBMCs compared to neurologically healthy controls (NHCs) [68]. Furthermore, maximal and spare respiratory capacity in PD PBMCs were significantly positively correlated with disease duration and degree of motor impairment as measured by MDS-UPDRS and Hoehn & Yahr scores [69]. It is possible that PBMCs adaptively increase their total mitochondrial content over the course of disease as a compensatory response to intermittent spikes in energetic demand. A separate study by Annesley et al. evaluated immortalized lymphoblastoid cell lines derived from iPD PBMCs and reported elevated basal mitochondrial activity and ATP synthesis [70]. This pattern of mitochondrial dysfunction differs from that observed by Schirinzi and colleagues, and instead suggests a higher baseline energetic demand in PD. The variability observed across reports could be attributed to multiple factors including insufficient power and the possibility that different subsets of PBMCs are differentially affected by PD status in terms of respiratory defects. For example, lymphocytes have been shown to exhibit a higher OCR/ECAR ratio than monocytes, indicating a greater utilization of oxidative phosphorylation in lymphocytes [71]. Therefore, future studies should address the possibility of underlying differences in PBMC population frequencies to reduce this potential source of variability.

Oxidative stress and mitochondrial reactive oxygen species

The mitochondrial ETC is the major source of intracellular ROS production [72], and mitochondrial dysfunction can lead to increased levels of superoxides, hydrogen peroxide, and hydroxyl radicals [73]. These ROS can cause significant damage to the cell by oxidizing proteins, DNA, and lipids, thereby compromising vital cell functions and viability [74]. Recent studies have revealed increased ROS levels [75], reduced function of antioxidant peroxidases [76], and decreased levels of glutathione, the primary endogenous antioxidant molecule, in the SNpc of PD patients relative to age-matched controls [77] (Fig. 2). Furthermore, PD-associated mutations in LRRK2, ATP13A2, DJ-1, SNCA, Parkin, and PINK1 are associated with increased levels of ROS in cellular and animal models [78,79,80,81,82,83,84,85]. While excessive ROS is known to be directly neurotoxic, there is a growing recognition for the role of ROS in immune activation. Here, we will discuss the evidence supporting dysregulated ROS homeostasis in immune cells in PD, as well as potential mechanisms connecting oxidative stress to immune-mediated neurodegeneration.

Markers of elevated oxidative stress have been described in CNS immune cells from individuals with PD and animal models of disease. Hunot et al. reported that microglia in the SNpc and ventral tegmental area (VTA) of individuals with PD express increased levels of inducible nitric oxide synthase (iNOS) [86]. Knott et al. replicated these findings and observed that SNpc microglia from PD patients but not controls displayed upregulation of iNOS, cyclo-oxygenase (COX)-1, and COX-287. In mouse models of PD-like degeneration, administration of LPS and nitrated α-synuclein led to increased ROS production in microglia [88,89,90]. Furthermore, mice with microglial-specific overexpression of α-synuclein showed upregulation of SOD1 and SOD2 in microglia, indicating a disrupted oxidative status [90]. Microglial ROS production in response to α-synuclein can be attenuated through knockout of NADPH oxidase [88], an enzyme whose function is to catalyze the production of superoxide and hydrogen peroxide by transferring electrons to molecular oxygen. Interestingly, α-synuclein-induced ROS production was also attenuated by blockade of voltage-gated potassium and chloride currents in microglia [91], suggesting that excitotoxicity can exacerbate microglial oxidative stress. Limited evidence suggests that PD is also associated with disrupted ROS homeostasis in astrocytes, with one group reporting that iPSC-derived astrocytes from PD patients express increased levels of oxidized proteins compared to controls [28].

Recent findings suggest that drivers of oxidative stress in microglia lead to downstream immune activation, creating a cytotoxic milieu that drives DAN degeneration in PD. Mouse microglia overexpressing human α-synuclein exhibit increased ROS and expression of CD68, a marker of activated microglia, compared to wild type cells [90]. Accumulation of α-synuclein in microglia was also accompanied by upregulation of proinflammatory chemokines and infiltration of T cells into the SNpc [90]. Furthermore, chronic rotenone treatment of microglia in vitro leads to increased ROS production and microglial activation [92, 93], consistent with prior research showing that mitochondrial ROS directly activate the NLRP3 inflammasome [94, 95]. Administration of iNOS inhibitors in mice with microglial-specific overexpression of human α-synuclein mitigated ROS generation, microglial activation, and TH+ DAN loss [90]. These findings highlight microglial ROS production as a relevant therapeutic target for future studies. Overall, these results suggest that oxidative stress in microglia is a critical component of the pathogenic cascade, upstream of immune activation and DAN degeneration in PD.

Studies thus far suggest that disrupted ROS homeostasis is not limited to CNS immune cells but may be present in peripheral blood immune cells. Prigione et al. described increased intracellular ROS in PD PBMCs compared to matched controls, with no change in glutathione reductase activity [96]. The authors also observed that daily levodopa dosage in the PD group was inversely correlated with ROS production, suggesting that levodopa may play a protective role in the redox status of PBMCs [96]. Similarly, lymphoblastoid cell lines derived from individuals with iPD displayed elevated ROS production compared to those derived from controls [70]. On the other hand, Smith et al. observed that PD monocytes displayed increased absolute ROS levels and ROS normalized to mitochondrial content, but PBMCs as a whole did not show significant differences [30]. ROS homeostasis is particularly relevant to monocytes due to their employment of ROS as a bactericidal agent against engulfed pathogens [97]. It will be vital to determine whether ROS homeostasis in PD shows cell-type specific differences, as focusing on the most disease-relevant cell types could enhance the sensitivity of future studies to detect immunometabolic deficits.

Disruption in mitochondrial membrane potential

Mitochondrial health can be measured in many ways, but one of the most common metrics is mitochondrial membrane potential (MMP). The maintenance of an electric potential across the inner mitochondrial membrane (negatively charged in the matrix) is necessary to power ATP synthase, and sustained decreases in MMP trigger mitophagy and overall decreases in cell viability [98]. A number of PD-related genes are known to maintain MMP in neurons. For example, knockdown (KD) of PINK1 in mouse DANs has been shown to cause disrupted calcium homeostasis, increased ROS production, and a reduced MMP [99]. Similarly, loss of DJ-1 causes reduced membrane potential and increased mitochondrial fragmentation [100].

Emerging literature suggests that immune cell mitochondrial health is compromised in PD models. For example, iPSC-derived astrocytes from individuals with LRRK2 PD have reduced MMP and increased mitochondrial fragmentation relative to controls [28]. Experimentally, chronic treatment with α-synuclein PFFs can reduce microglial MMP [23], indicating that aggregated α-synuclein may compromise mitochondrial membrane integrity. Indeed, Choi et al. reported that the mitochondrial lipid cardiolipin can catalyze the oligomerization of A53T α-synuclein, and the buildup of α-synuclein aggregates at mitochondria causes permeabilization of mitochondrial membranes and cell death [101]. Mouse models of PARK7/DJ-1 KO show no differences in microglial MMP at baseline, however LPS treatment significantly reduces microglial MMP in DJ-1 KO relative to wild type [102]. These results imply that loss of DJ-1 impairs mitophagy and renders microglia vulnerable to other stresses, which is consistent with previous reports in neurons describing DJ-1 as a mediator of PINK1/Parkin-dependent mitophagy [103]. In the periphery, Smith et al. reported that iPD PBMCs do not exhibit alterations in MMP relative to controls [30], in contrast to Qadri et al. who observed significantly reduced MMP in PD PBMCs relative to controls in both stimulated and unstimulated conditions [104]. This discrepancy could be due to the different probes used by each group; Smith et al. used tetramethylrhodamine methyl ester, while Qadri et al. used JC-130,104. The possibility of using peripheral immune cell MMP as a biomarker for PD is exciting, however it will be vital for the field to develop a consistent methodology for evaluating MMP before these findings can be translated into providing clinical benefit.

Mitophagy defects

Mitophagy is an essential quality-control process wherein damaged mitochondria are targeted, engulfed by autophagosomes, and ultimately degraded by lysosomes [105]. Defects in mitophagy have been heavily implicated in PD [103, 106,107,108], and mitophagy in immune cells has received attention due to studies linking the accumulation of damaged mitochondria to inflammasome activation [109]. Therefore, it is possible that disease-modifying therapies aimed at enhancing mitophagy in PD will both improve neuronal viability and combat aberrant immune activation. In a study by Singh et al., mice carrying the pathogenic LRRK2 G2019S mutation showed reduced basal mitophagy in both neurons and microglia, while LRRK2 KO was associated with increased mitophagy [107]. Furthermore, pharmacologic inhibition of LRRK2 kinase activity rescued these mitophagy defects in vivo [107]. It has also been shown that mice with a microglia-specific knockout of autophagy protein-5 (Atg5), a protein known to mediate mitophagy [110], exhibit poorer performance in Rotarod and Morris water-maze tasks as well as loss of TH+ neurons111. Microglia with Atg5 KO showed increased activation of the NLRP3 inflammasome, and pharmacologic inhibition of inflammasome activity caused reduced immune activation and reduced TH+ neuronal loss [111]. Studies using LPS and MPTP mouse models have reported similar findings, with pharmacologic activators of mitophagy causing reduced microglial activation and improved DAN viability [112, 113]. Taken together, these findings provide strong evidence that mitophagy deficits specifically in microglia can contribute to PD-like degeneration, and this is likely driven by inflammasome-mediated immune activation.

Mitochondrial DNA defects

Mitochondria are unique among cellular organelles as they contain their own DNA, and the cellular mechanisms for replicating and proofreading mitochondrial DNA (mtDNA) are distinct from those used for nuclear DNA [114]. Postmortem samples of the SNpc from PD donors show an increased number of mtDNA deletions and rearrangements relative to NHCs and AD, suggesting that these classes of mtDNA defects may be specific to PD pathogenesis [115,116,117]. Notably, mtDNA defects in PBMCs may have clinical value as a blood-based biomarker for PD. Qi et al. developed a PCR-based assay based on the principle that lesions in mtDNA block the ability of DNA polymerase to replicate, allowing for quantification of mtDNA damage [118]. Using this assay, the authors determined that iPD and LRRK2 PD PBMCs display significantly increased levels of mtDNA defects, and pharmacologic inhibition of LRRK2 kinase activity effectively reduced the level of mtDNA defects in iPD PBMCs to the same levels observed in NHCs [118]. This promising technique could be leveraged to identify PD endophenotypes using the accessibility of PBMCs, with the potential to improve patient classification into clinical trials for therapeutics aimed at rescuing mitochondrial function. Future efforts should be directed towards investigating if mtDNA defects occur prior to the manifestation of motor symptoms in prodromal PD, and additionally to explore the mechanism by which LRRK2 kinase activity contributes to mtDNA damage.

Defects in mtDNA have also been shown to play a role in immune activation and inflammation. When mitochondria are damaged, mtDNA can leak into the cytoplasm due to mitochondrial outer membrane permeabilization or defective mitophagy [119]. The cytosolic DNA sensor cyclic GMP-AMP synthase (cGAS) recognizes this leaked mtDNA as a danger-associated molecular pattern (DAMP) and catalyzes the production of cyclic GMP-AMP (cGAMP) [119, 120]. cGAMP then binds to and activates stimulator of interferon genes (STING) [120], a key adaptor protein located on the endoplasmic reticulum. STING activation ultimately results in the production of type I interferons and pro-inflammatory cytokines [120]. Hinkle et al. recently observed that double stranded DNA breaks in an α-synuclein PFF mouse model of PD triggered dopaminergic degeneration in a STING-dependent manner [121]. Importantly, STING-deficient mice were protected from dopaminergic neuron loss and motor deficits induced by α-synuclein PFFs [121]. In addition, Hancock-Cerutti et al. reported that KO of VPS13C, a gene associated with autosomal recessive PD, also leads to inflammation mediated by mtDNA and the cGAS-STING pathway [122]. The VPS13 gene family is involved in lipid transport to mitochondria and lysosomes [123], and VPS13KO HeLa cells show increased cytosolic mtDNA levels and simultaneous cGAS-STING activation [122]. Depletion of mtDNA with ethidium bromide in these cells was effective in reversing the increased expression of IFN-stimulated genes [122]. These results support a mechanism whereby mtDNA defects and cGAS-STING signaling link mitochondrial dysfunction to immune activation in PD.

Lysosomal dysfunction and autophagy

Lysosomal dysfunction has emerged as a key player in the pathogenesis of PD. Lysosomes are responsible for the degradation and recycling of intracellular waste, damaged proteins, and cellular debris. One of the hallmark pathological features of PD is the accumulation of abnormal protein aggregates, particularly α-synuclein, within neurons, leading to the formation of Lewy bodies. Growing evidence suggests that impaired lysosomal function plays a key role in the failure to clear these toxic protein aggregates, contributing to the progressive neurodegeneration observed in PD [124, 125]. Some of the most commonly mutated genes in familial cases of PD, including LRRK2 and GBA1, have key functions in regulating lysosomal health [126,127,128] and both are highly expressed in immune cells [9, 129, 130]. Furthermore, lysosomal and autophagic deficits have been shown to contribute to NF-κB signaling and inflammasome activation [131]. Therefore, emerging literature has begun to identify if immune cells in PD show dysfunction in autophagic flux, vesicle trafficking, and lysosomal degradative capacity.

Autophagic deficits in PD

Autophagy is the process by which cells break down damaged or unneeded components so that these materials can be repurposed for new cell parts. Autophagic activity protects cellular energy homeostasis by decreasing bioenergetic demand [132], and it plays a vital role in host defense against microbial pathogens following phagocytosis [133]. Neuronal autophagic deficits have been implicated in PD, with SNpc samples showing that PD neurons have increased autophagic degeneration including vacuolation of the endoplasmic reticulum and increased lysosome-like vacuoles [134]. In addition, autophagic-lysosomal pathway impairments are known to contribute to the buildup of misfolded proteins including α-synuclein and tau [135, 136]. Furthermore, PD brains display reduced expression of lysosomal-associated membrane protein 2 A (LAMP2A) and heat shock cognate 70 protein (hsc70), which suggests that chaperone-mediated autophagy is reduced in PD [137]. Mutations in several PD-related genes have been linked to autophagic deficits in neurons, including SNCA, LRRK2, PARK7/DJ-1, and VPS35138,139,140,141. Autophagic deficits in PD are also found in immune cells, which is particularly relevant because microglia are the predominant cell type responsible for clearing aggregated α-synuclein in the CNS [142].

Both in vitro and in vivo experiments have demonstrated that α-synuclein exposure suppresses autophagic flux in microglia [90, 143]. Consequently, underlying autophagic deficits may compound upon themselves with the buildup of intracellular α-synuclein aggregates causing further disruption in microglial clearance of misfolded proteins. Nash et al. demonstrated that KD of DJ-1 in microglia significantly decreased phagocytosis of α-synuclein and reduced expression of autophagy-related markers including p62 and LC3-II [144]. DJ-1 KD microglia also showed an enhanced pro-inflammatory response to α-synuclein compared to controls, displaying increased secretion of IL-6, IL-1β, and nitric oxide (NO) [144]. Similarly, autophagic deficits caused by a microglial-specific deletion of Atg7 in mice caused increased microglial activation [145]. Moreover, microglial-specific KO of Atg7 exacerbated the pathologic spreading of misfolded tau injected into the brains of mice expressing the P301S mutant form of human tau. Tau is a microtubule associated protein that functions to stabilize microtubules; however, abnormal aggregation of tau has been reported in a variety of neurodegenerative diseases including AD, frontotemporal dementia (FTD), progressive supranuclear palsy, LRRK2 PD, and others [146, 147]. These results indicate that defective microglial autophagy can drive inflammation in the context of pathologic protein aggregation, thereby promoting DAN degeneration. Indeed, mice with microglial-specific KO of Atg5 were found to be significantly more vulnerable to MPTP-induced toxicity and exhibited increased DAN degeneration and poorer performance on motor tests compared to WT mice [148]. Mechanistically, it was found that blocking microglial autophagy in vitro with Atg5 siRNA led to enhanced NLRP3 inflammasome activation after LPS treatment [148].

Several autophagy-related proteins are known to be differentially expressed in peripheral immune cells from individuals with PD. PD PBMCs exhibit increased mRNA and protein expression of autophagy-related p62 relative to matched controls [149, 150]. Additionally, expression of LAMP2 is significantly decreased in PBMCs from individuals with sporadic PD [151], which may indicate impaired fusion of autophagosomes and lysosomes. A number of other transcripts associated with autophagosome formation (ULK3, ATG2A, ATG4B, ATG5, ATG16L1 and HDAC6) were found to be downregulated in iPD PBMCs, while protein levels of ULK1, Beclin-1, and autophagy/beclin-1 regulator 1 were increased [152]. Expression of these autophagy related proteins positively correlated with increased levels of α-synuclein in PBMCs [152], suggesting that this may represent a compensatory response to trigger increased protein degradation.

The microtubule-associated protein light chain 3 (LC3) is used as a general marker of autophagic activity, with the conversion of cytosolic LC3-I to lipidated LC3-II associated with increased autophagosome formation [153]. In astrocytes, overexpression of WT, A30P, and A53T α-synuclein promoted decreased LC3-II and increased p62 expression, suggesting reduced autophagic activity [154]. Increased LC3 gene expression and LC3-II protein levels has been found in iPD PBMCs [151], and two studies have reported an increased LC3-II/LC3-I ratio [151, 155]. On the other contrary, Miki et al. observed no significant difference in LC3 mRNA or LC3-II/LC3-I ratio in PD PBMCs relative to NHCs [152]. Elevated LC3-II can arise from impaired fusion of autophagosomes with lysosomes [151], therefore it is possible that both LC3-I and LC3-II levels are increased in PD PBMCs due to disruption of downstream clearance of autophagosomes.

Endolysosomal and vesicle trafficking

LRRK2

Mutations in LRRK2 are the most frequent cause of late-onset autosomal dominant and sporadic PD [156]. Physiologically, LRRK2 is known to participate in a variety of functions related to endolysosomal maturation and trafficking [8, 157]. One of the potential mechanisms by which LRRK2 mutations may contribute to neuronal dysfunction is through impairment of vesicular trafficking. This was established by studies demonstrating that Rab GTPases are bona fide substrates of LRRK2 kinase activity [158,159,160]. Rab family proteins are known to regulate early and late endocytic trafficking [161], thus changes in their phosphorylation levels caused by gain-of-kinase mutations in LRRK2 may disrupt endosome transport. Additionally, LRRK2 directly interacts with a number of proteins associated with vesicle transport such as actin [162], JIP4163, ARFGAP1164, VPS52165, Sec16a [166], and N-ethylmaleimide sensitive fusion protein (an ATPase that facilitates SNARE complex disassembly) [167].

Similarly, work by our group demonstrated that G2019S mutation in murine macrophages caused an increase in phagocytosis and MHC-II trafficking from the perinuclear lysosome to the plasma membrane [168]. We found that macrophages from G2019S knock-in mice exhibit an increased ratio of extracellular: intracellular MHC-II which was mitigated by Lrrk2 kinase inhibition. Furthermore, we determined that Lrrk2 modulates antigen presentation through mTOR dependent lysosomal-tubule-formation (LTF). Macrophages with the G2019S mutation showed a significant increase in LTF relative to WT, and macrophages nucleofected with Lrrk2 anti-sense oligonucleotide (ASO) to cause Lrrk2 KD showed a significant reduction in LTF. Lastly, gene ontology analysis revealed that vesicular trafficking and lysosomal positioning pathways were associated with the response to IFN-γ treatment in Lrrk2 ASO-nucleofected macrophages. These results highlight that LRRK2 is a key regulator of lysosome and vesicle trafficking, and maintaining proper immune function will be an important consideration for therapeutics which target LRRK2 activity.

VPS35

VPS35 is an essential component of the retromer cargo recognition complex which functions in membrane protein transport and trafficking in the trans-Golgi network [169]. Mutations in VPS35, especially the D620N mutation, are known causes of familial PD [170, 171], and mouse models carrying VPS35 D620N have been shown to exhibit Parkinsonian motor features, tau neuropathology, and dopaminergic degeneration [172, 173]. Intriguingly, monocytes and neutrophils derived from individuals with VPS35 D620N display increased LRRK2-mediated Rab10 phosphorylation compared to controls and iPD, and this difference is abrogated upon LRRK2 kinase inhibition with MLi-2174. VPS35 mutation is not sufficient to cause neuroinflammation, as D620N knock-in mice do not exhibit increased astrogliosis or microgliosis [172, 173], but D620N exacerbates microgliosis secondary to MPTP treatment [172]. Similarly, VPS35 KD in mouse microglia causes an enhanced response to LPS treatment with increased expression of iNOS and IL-6175. VPS35 deficiency is also shown to impair Trem2 trafficking back to the trans-golgi network, leading to increased Trem2 accumulation in the lysosome [175], but overexpression of Trem2 mitigates the proinflammatory phenotype in VPS35 KO microglia [175]. In addition, Pal et al. demonstrated that VPS35 D620N mouse embryonic fibroblasts show broad changes in lysosomal protein contents and increased recruitment of the phospho-Rab effector protein RILPL1 to the lysosome [176]. VPS35 D620N lysosomes also showed enhanced recruitment of LRRK2176, which may be a potential mechanism to explain the immune priming observed by other groups. Bu et al. reported deficits in dopamine transporter expression in D620N knock-in mice that was rescued by LRRK2 kinase inhibitors [177], but the authors did not investigate immune cell specific effects. Collectively, these results suggest that VPS35 and LRRK2 converge on similar pathways related to immune activation and endolysosomal function. Future effort is warranted to investigate how VPS35 mutations modulate immune cell function in humans, and whether therapeutics targeting LRRK2 kinase activity may offer clinical benefit to this patient subset.

Lysosomal degradative capacity

Decreased activity of lysosomal hydrolase enzymes has been reported in postmortem analyses of the SNpc of PD patients [178, 179] and in mouse models of SNCA and GBA1 mutations180-181. Multiple lines of evidence now suggest that deficits in lysosomal degradative capacity in PD extend to immune cells in both the CNS and the periphery (Fig. 3). The pathogenic LRRK2 G2019S mutation is associated with reduced expression of lysosomal hydrolases in macrophages and microglia [182], and astrocytes from Lrrk2 G2019S mice have reduced cathepsin B activity [183]. Pharmacologic inhibition of LRRK2 kinase activity rescues these phenotypes, indicating that the effects of G2019S on lysosomal degradative capacity are kinase-specific [182, 183]. Human iPSC-derived macrophages with G2019S mutation also show deficits in both internalization and degradation of tau fibrils compared to wild type [184]. Furthermore, non-degraded tau fibrils from LRRK2 KO macrophages did not retain seeding competency 24 h post incubation, but fibrils from G2019S macrophages retained seeding competency and were able to induce aggregation of naïve tau monomers [184]. These findings suggest that lysosomal degradative activity in immune cells is critical for modulation of toxic protein aggregation in PD.

The vast majority of lysosomal enzymes exhibit optimal function at acidic pH levels [185], thus it is unsurprising that deficits in lysosomal acidification result in decreased degradative capacity [186]. The SNpc from individuals with iPD shows significant dysregulation in expression and activity of proteins linked to lysosomal acidification including ATP13A2 and LRRK2 relative to control brains [187, 188]. Intriguingly, LRRK2 G2019S is associated with reduced lysosomal pH in astrocytes [183] but increased pH in neurons [189], indicating cell-type specific effects. Thus far, it has been shown that increased endogenous expression of α-synuclein disrupts lysosomal pH in iPSC-derived macrophages [190], and impaired lysosomal acidification predisposes microglia to enhanced secretion of proinflammatory cytokines [191]. TREM2 is a gene involved in lysosomal acidification [192] that has been linked to increased risk for AD [193, 194], with mixed evidence supporting a possible association between TREM2 variants and increased PD susceptibility [193, 195,196,197]. TREM2 deficiency was recently shown to cause defective vesicle acidification and heightened immune activation in microglia [192, 198], and KO of Trem2 in mice severely dysregulates neuronal OXPHOS and mitochondrial structure [199]. Together, these results suggest that deficits in microglial lysosomal acidification may contribute to neuroinflammation in PD and may represent a potential therapeutic target. In support of this, pharmacologic treatment to enhance lysosomal acidification was shown to substantially reduce neurodegeneration and microgliosis in a MPTP model of PD-like degeneration [200].

Mutations in the GBA1 gene which encodes the lysosomal enzyme glucocerebrosidase (GCase) constitute a major genetic risk factor for PD [201, 202], and recent experiments have demonstrated that GCase activity contributes to the clearance of α-synuclein aggregates [203, 204]. These findings have prompted several studies aimed at investigating how GCase activity is altered in microglia and other immune cells in PD. GCase activity has been heavily implicated in microglial activation, with reductions in microglial GCase activity associated with increased neuroinflammation, α-synuclein aggregation, and neurodegeneration [204,205,206,207]. Interestingly, genetic ablation of GBA1 exclusively in midbrain DANs fails to cause degeneration, motor deficits, or α-synuclein aggregation in mice despite significantly increased microglial activation [206]. This suggests that the pathogenic effects of GBA1 deficiency are likely mediated through non-neuronal cell types. Indeed, mouse astrocytes with knock in of the loss-of-function mutation D409V in GBA1 show reductions in lysosomal count, increased lysosomal pH, and reduced cathepsin B activity [208]. In addition, Brunialti et al. reported that microglial-specific inhibition of GCase activity with conduritol-B-epoxide in microglia-neuron co-cultures caused a significant reduction in the expression of nuclear factor erythroid 2-related factor 2 (NFE2L2) activity in neurons [209]. NFE2L2 is a transcription factor which regulates the expression of antioxidant proteins and protects against oxidative damage triggered by inflammation; therefore, these findings suggest that deficits in microglial GCase activity can render nearby neurons vulnerable to oxidative stress. These data collectively suggest that microglial GCase activity regulates lysosomal function and is important for protecting neuronal viability in PD.

A number of PD-associated mutations converge on alterations in GCase activity in immune cells (Fig. 3). PBMCs from PD patients carrying GBA1 and SNCA A53T mutations show significant reductions in GCase protein expression relative to NHCs [10]. In contrast, Kedariti et al. observed that LRRK2 PD PBMCs display increased GCase enzymatic activity relative to both iPD and controls [210]. This implies that LRRK2 kinase activity regulates GCase in a cell-type specific manner, as neurons derived from individuals with LRRK2 PD show reduced GCase activity which can be rescued by inhibition of LRRK2 kinase activity via MLi-2211. Peripheral immune cell subtypes vary in their baseline protein expression of GCase and level of GCase activity [9, 212], and cell-type specific assays may reveal differences that are not detectable in total PBMCs. For example, Atashrazm et al. found that isolated PD monocytes but not lymphocytes displayed reduced GCase activity relative to controls [212]. GBA1 mutant monocyte-derived macrophages exhibit autophagic defects and increased secretion of IL-1β and IL-6213, consistent with the hyperinflammatory phenotype reported in GBA1 mutant microglia. These defects were attenuated by treatment with the pharmacologic GCase activator NCGC758213, indicating that rescue of GCase activity could have therapeutic benefits in preventing aberrant inflammation.

GCase activity in peripheral immune cells may relate to clinical characteristics of PD. Higher GCase activity in blood and peripheral immune cells from individuals with iPD was associated with younger age at onset, longer disease duration, greater levodopa use, and poorer performance on UPDRS and Montreal Cognitive Assessment [214, 215]. This is unexpected given that GCase activity in the blood is modestly lower in those with PD relative to NHCs [214]. The current literature points to GCase activity and lysosomal function as major players in the role of immune cells in PD, however significant gaps in our understanding remain, in particular regarding how different GBA1 mutations may not produce a uniform phenotype. The L444P variant in GBA1 is associated with a complete loss of GCase activity [216], whereas the N370S substitution is associated with retention of approximately 10% of the activity of WT [217]. Many of the studies performed thus far have grouped all GBA1 mutations under a single umbrella due to the difficulty in powering comparisons between genotypes. However, greater consistency in genotyping, reporting, and comparing the effects between GBA1 mutations in study populations moving forwards will be pivotal to defining the mechanisms linking GCase to PD.

Fig. 3
figure 3

Defects in autophagy and lysosomal function in PD show similarities across neurons and immune cells. This figure highlights areas of autophagic and lysosomal dysfunction that show overlap across neurons, microglia, and peripheral immune cells in PD. These similarities may indicate that peripheral immune cells have potential to be used as an accessible tissue source in clinical practice for the investigation of metabolic deficits in the CNS of PD patients. Therapeutic strategies aimed at rescuing lysosomal function in PD may improve neuronal health while also rescuing lysosomal activity in other cell types. Abbreviations: idiopathic Parkinson’s disease (iPD); endoplasmic reticulum (ER); glucocerebrosidase (GCase); central nervous system (CNS). Created with BioRender.com

Ubiquitin-proteasome function

The presence of intracellular inclusions containing aggregated α-synuclein is one of the defining histopathological hallmarks of PD [218], and a wealth of evidence supports the role of dysregulated protein homeostasis in PD pathogenesis. Bioenergetic deficits in neurons have been proposed as a cause of impaired protein quality control in PD [4], and conversely, protein homeostasis is essential for maintaining a number of cellular pathways including autophagy [219] and mitochondrial function [220, 221] which support bioenergetic sufficiency. The ubiquitin-proteasome system (UPS), a major pathway responsible for targeted protein degradation within the cell, has been strongly implicated in the etiology of PD [222, 223]. In this process, proteins are marked for degradation by covalent attachment of ubiquitin molecules and then degraded by a large intracellular protein complex known as a proteasome which releases the reusable ubiquitin [224].

Inhibition of UPS function can compromise cellular metabolism, leading to mitochondrial damage and increased generation of ROS [225]. A number of mutations that lead to familial PD have been linked to alterations in UPS function, including mutations in SNCA [226], UCHL1227, and PRKN [228]. In addition, SNpc samples from individuals with sporadic PD show decreased levels of alpha proteasome subunits [229, 230], decreased expression of the proteasome regulatory molecule PA700222, and reduced hydrolyzing activities of the 20/26S proteasome [223]. Experimental models have further implicated the UPS in PD pathophysiology, with iPSCs from PD patients and pesticide-induced animal models of PD-like degeneration showing impairment in UPS activity [231,232,233]. These findings have contributed to an understanding that aberrant protein aggregation in neurons directly promotes neurodegeneration. However, the UPS has also been shown to regulate immune function with the ability to both activate and dampen inflammasome activation [234]. This has led to a growing understanding that UPS dysfunction in immune cells may contribute to neurodegeneration in PD by driving inflammatory responses [235, 236].

Recent studies have demonstrated that inhibition of proteasome activity causes increased microglial activation in experimental models. KD of HACE1, which encodes an E3 ubiquitin-ligase involved in the UPS, was shown to exacerbate LPS-induced inflammation in BV2 microglia [237]. In addition, BV2 cells treated with siRNA for HACE1 showed increased neurotoxicity during co-culture with SH-SY5Y cells compared to vehicle treated BV2 cells, indicating the neuroprotective effect of this protein [237]. Furthermore, KD of HACE1 in MPTP-treated mice led to more severe motor deficits than MPTP treatment alone [237]. In line with these findings, administration of proteasome inhibitors in rat models of α-synuclein overexpression causes significant microgliosis and α-synuclein aggregation [238, 239]. Even in wild type rats, systemic exposure to proteasome inhibitors was sufficient to recapitulate parkinsonian motor features and DAN degeneration [240]. These findings provide strong evidence that disrupted protein homeostasis in microglia contributes to neuroinflammation and neurotoxicity.

Dysregulated protein homeostasis has also been reported in the peripheral blood of PD patients. Activity of the 20 S proteasome and expression of the ubiquitin conjugating enzyme E2 were found to be lower in PD PBMCs compared to controls [241]. Furthermore, lower 20 S proteasome activity in PBMCs was correlated with longer disease duration and greater severity of motor symptoms. No changes were reported with AD status [241], suggesting that this biomarker of metabolic dysfunction in PBMCs could be specific for PD and have clinical value for predicting disease progression. In addition, DJ-1 isoforms in whole blood from PD patients were shown to have a higher number of 4-hydroxy-2-nonenal post-translational modifications (PTMs) relative to AD patients or NHCs [242]. These PTMs are associated with inhibition of 20 S proteasome activity, and interestingly the frequency of 4-hydroxy-2-nonenal PTMs was positively correlated with PD disease severity [242]. Together, these results point towards systemic UPS defects in PD, and peripheral blood may serve as an accessible biofluid for evaluating PD-related defects in protein homeostasis.

Conclusion

Cellular metabolic dysfunction represents a crucial component of the pathogenic cascade in PD. Mounting literature describes widespread disruption in energy homeostasis in immune cells, and these deficits converge on aberrant inflammation in both the central and peripheral immune systems. The emerging data suggest that metabolic reprogramming towards increased glycolytic activity is likely to be both a cause and a consequence of excessive inflammation in PD, setting the stage for a feed-forwards inflammatory milieu which renders DANs particularly vulnerable to degeneration. A number of genetic and environmental causes of PD converge on mitochondrial dysfunction, including deficits in ETC activity, ROS homeostasis, and mitochondrial respiration. In addition, changes in autophagic activity and lysosomal health have been shown to increase energetic demand. These facets of cellular metabolic dysfunction, many of which were first noted in neurons in PD, have now been shown to extend to immune cells in both the CNS and the periphery. Moreover, energy insufficiency appears to prime immune activation, leading to increased proinflammatory responses to primary insults. As growing attention is directed towards therapeutics targeting metabolism in PD, these strategies may be both directly neuroprotective as well as serve to mitigate the hostile inflammatory environment produced by activated immune cells.

We have highlighted here multiple studies demonstrating the success of pharmacologic interventions targeting glycolysis, mitophagy, and lysosomal pathways in preclinical PD models. However, considerable effort is still required to understand how these metabolic processes in immune cells interact and contribute to neurodegeneration. While significant attention has been directed towards metabolic changes in microglia, astrocytes remain a relatively understudied area with significant gaps in knowledge that must be addressed. Additionally, the field must devote future research towards assessing the feasibility and effectiveness of using immune-targeted therapies to rescue immunometabolic deficits and potentially delay or slow PD progression. Given the extensive parallels in metabolic dysfunction between the peripheral immune system and the central nervous system, the peripheral immune system may provide a non-invasive window into the brain for evaluating central bioenergetic deficits in PD. As we move beyond merely identifying bioenergetic deficits and begin to develop strategies to restore function, the field will need a cell-type-specific understanding of the complex pathways that converge on immunometabolic dysfunction in PD.

Data availability

Not applicable.

References

  1. Ou Z, Pan J, Tang S, et al. Global trends in the incidence, prevalence, and years lived with disability of Parkinson’s disease in 204 countries/territories from 1990 to 2019. Front Public Health. 2021;9:776847.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Surmeier DJ. Determinants of dopaminergic neuron loss in Parkinson’s disease. Febs J. 2018;285(19):3657–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Błaszczyk JW. The emerging role of energy metabolism and neuroprotective strategies in Parkinson’s disease. Front Aging Neurosci. 2018;10:301.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Anandhan A, Jacome MS, Lei S, et al. Metabolic dysfunction in Parkinson’s disease: bioenergetics, redox homeostasis and central carbon metabolism. Brain Res Bull. 2017;133:12–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Duda J, Pötschke C, Liss B. Converging roles of ion channels, calcium, metabolic stress, and activity pattern of substantia Nigra dopaminergic neurons in health and Parkinson’s disease. J Neurochem. 2016;139(1):156–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Haas RH. Mitochondrial dysfunction in aging and diseases of aging. Biology (Basel) 2019;8(2).

  7. Moon HE, Paek SH. Mitochondrial dysfunction in Parkinson’s disease. Exp Neurobiol. 2015;24(2):103–16.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Tansey MG, Wallings RL, Houser MC, Herrick MK, Keating CE, Joers V. Inflammation and immune dysfunction in Parkinson disease. Nat Rev Immunol. 2022;22(11):657–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Hughes LP, Pereira MMM, Hammond DA, et al. Glucocerebrosidase activity is reduced in cryopreserved Parkinson’s disease patient monocytes and inversely correlates with motor severity. J Parkinsons Dis. 2021;11(3):1157–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Papagiannakis N, Xilouri M, Koros C, et al. Autophagy dysfunction in peripheral blood mononuclear cells of Parkinson’s disease patients. Neurosci Lett. 2019;704:112–5.

    Article  CAS  PubMed  Google Scholar 

  11. Lunt SY, Vander Heiden MG. Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annu Rev Cell Dev Biol. 2011;27:441–64.

    Article  CAS  PubMed  Google Scholar 

  12. Borghammer P, Chakravarty M, Jonsdottir KY, et al. Cortical hypometabolism and hypoperfusion in Parkinson’s disease is extensive: probably even at early disease stages. Brain Struct Funct. 2010;214(4):303–17.

    Article  PubMed  Google Scholar 

  13. Chandel NS, Glycolysis. Cold Spring Harb Perspect Biol. 2021;13(5).

  14. Arnold PK, Finley LWS. Regulation and function of the mammalian Tricarboxylic acid cycle. J Biol Chem. 2023;299(2):102838.

    Article  CAS  PubMed  Google Scholar 

  15. Cooper G. The cell: A molecular approach. 2nd ed. Sunderland (MA): Sinauer Associates; 2000.

    Google Scholar 

  16. Kierans SJ, Taylor CT. Glycolysis: A multifaceted metabolic pathway and signaling hub. J Biol Chem. 2024;300(11):107906.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Nolfi-Donegan D, Braganza A, Shiva S. Mitochondrial electron transport chain: oxidative phosphorylation, oxidant production, and methods of measurement. Redox Biol. 2020;37:101674.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sadeghdoust M, Das A, Kaushik DK. Fueling neurodegeneration: metabolic insights into microglia functions. J Neuroinflammation. 2024;21(1):300.

    Article  PubMed Central  Google Scholar 

  19. Han R, Liang J, Zhou B. Glucose metabolic dysfunction in neurodegenerative Diseases-New mechanistic insights and the potential of hypoxia as a prospective therapy targeting metabolic reprogramming. Int J Mol Sci 2021;22(11).

  20. Gao C, Jiang J, Tan Y, Chen S. Microglia in neurodegenerative diseases: mechanism and potential therapeutic targets. Signal Transduct Target Ther. 2023;8(1):359.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Okarmus J, Havelund JF, Ryding M, et al. Identification of bioactive metabolites in human iPSC-derived dopaminergic neurons with PARK2 mutation: altered mitochondrial and energy metabolism. Stem Cell Rep. 2021;16(6):1510–26.

    Article  CAS  Google Scholar 

  22. Cai R, Zhang Y, Simmering JE, et al. Enhancing Glycolysis attenuates Parkinson’s disease progression in models and clinical databases. J Clin Invest. 2019;129(10):4539–49.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Lu J, Wang C, Cheng X, et al. A breakdown in microglial metabolic reprogramming causes internalization dysfunction of α-synuclein in a mouse model of Parkinson’s disease. J Neuroinflammation. 2022;19(1):113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Cheng J, Zhang R, Xu Z, et al. Early glycolytic reprogramming controls microglial inflammatory activation. J Neuroinflammation. 2021;18(1):129.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gu R, Zhang F, Chen G, et al. Clk1 deficiency promotes neuroinflammation and subsequent dopaminergic cell death through regulation of microglial metabolic reprogramming. Brain Behav Immun. 2017;60:206–19.

    Article  CAS  PubMed  Google Scholar 

  26. Dong Y, Benveniste EN. Immune function of astrocytes. Glia. 2001;36(2):180–90.

    Article  CAS  PubMed  Google Scholar 

  27. Rodgers KR, Lin Y, Langan TJ, Iwakura Y, Chou RC. Innate immune functions of astrocytes are dependent upon tumor necrosis Factor-Alpha. Sci Rep. 2020;10(1):7047.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ramos-Gonzalez P, Mato S, Chara JC, Verkhratsky A, Matute C, Cavaliere F. Astrocytic atrophy as a pathological feature of Parkinson’s disease with LRRK2 mutation. NPJ Parkinsons Dis. 2021;7(1):31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sonninen TM, Hämäläinen RH, Koskuvi M, et al. Metabolic alterations in Parkinson’s disease astrocytes. Sci Rep. 2020;10(1):14474.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Smith AM, Depp C, Ryan BJ, et al. Mitochondrial dysfunction and increased Glycolysis in prodromal and early Parkinson’s blood cells. Mov Disord. 2018;33(10):1580–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Galbiati A, Verga L, Giora E, Zucconi M, Ferini-Strambi L. The risk of neurodegeneration in REM sleep behavior disorder: A systematic review and meta-analysis of longitudinal studies. Sleep Med Rev. 2019;43:37–46.

    Article  PubMed  Google Scholar 

  32. St Louis EK, Boeve BF. REM Sleep Behavior Disorder: Diagnosis, Clinical Implications, and Future Directions. Mayo Clin Proc. 2017;92(11):1723–1736.

  33. Zhang H, Wang T, Li Y, et al. Plasma immune markers in an idiopathic REM sleep behavior disorder cohort. Parkinsonism Relat Disord. 2020;78:145–50.

    Article  PubMed  Google Scholar 

  34. Kim R, Kim HJ, Kim A, et al. Peripheral blood inflammatory markers in early Parkinson’s disease. J Clin Neurosci. 2018;58:30–3.

    Article  PubMed  Google Scholar 

  35. Uehara I, Kajita M, Tanimura A, et al. 2-Deoxy-d-glucose induces deglycosylation of Proinflammatory cytokine receptors and strongly reduces immunological responses in mouse models of inflammation. Pharmacol Res Perspect. 2022;10(2):e00940.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Chaban Y, Boekema EJ, Dudkina NV. Structures of mitochondrial oxidative phosphorylation supercomplexes and mechanisms for their stabilisation. Biochim Biophys Acta. 2014;1837(4):418–26.

    Article  CAS  PubMed  Google Scholar 

  37. Zhao RZ, Jiang S, Zhang L, Yu ZB. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int J Mol Med. 2019;44(1):3–15.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Dawson TM, Dawson VL. The role of parkin in Familial and sporadic Parkinson’s disease. Mov Disord. 2010;25(0 1):S32–39.

    PubMed  PubMed Central  Google Scholar 

  39. Greene JC, Whitworth AJ, Kuo I, Andrews LA, Feany MB, Pallanck LJ. Mitochondrial pathology and apoptotic muscle degeneration in drosophila parkin mutants. Proc Natl Acad Sci U S A. 2003;100(7):4078–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Palacino JJ, Sagi D, Goldberg MS, et al. Mitochondrial dysfunction and oxidative damage in parkin-deficient mice. J Biol Chem. 2004;279(18):18614–22.

    Article  CAS  PubMed  Google Scholar 

  41. Shaltouki A, Sivapatham R, Pei Y, et al. Mitochondrial alterations by PARKIN in dopaminergic neurons using PARK2 patient-specific and PARK2 knockout isogenic iPSC lines. Stem Cell Rep. 2015;4(5):847–59.

    Article  CAS  Google Scholar 

  42. Jia F, Fellner A, Kumar KR. Monogenic Parkinson’s Disease: Genotype, Phenotype, Pathophysiology, and Genetic Testing. Genes (Basel) 2022;13(3).

  43. Pickrell AM, Youle RJ. The roles of PINK1, Parkin, and mitochondrial fidelity in Parkinson’s disease. Neuron. 2015;85(2):257–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Chen C, McDonald D, Blain A, et al. Imaging mass cytometry reveals generalised deficiency in OXPHOS complexes in Parkinson’s disease. NPJ Parkinsons Dis. 2021;7(1):39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Antony PMA, Kondratyeva O, Mommaerts K, et al. Fibroblast mitochondria in idiopathic Parkinson’s disease display morphological changes and enhanced resistance to depolarization. Sci Rep. 2020;10(1):1569.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bose A, Beal MF. Mitochondrial dysfunction in Parkinson’s disease. J Neurochem. 2016;139(Suppl 1):216–31.

    Article  CAS  PubMed  Google Scholar 

  47. Henrich MT, Oertel WH, Surmeier DJ, Geibl FF. Mitochondrial dysfunction in Parkinson’s disease - a key disease hallmark with therapeutic potential. Mol Neurodegener. 2023;18(1):83.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Klein C, Westenberger A. Genetics of Parkinson’s disease. Cold Spring Harb Perspect Med. 2012;2(1):a008888.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Funayama M, Nishioka K, Li Y, Hattori N. Molecular genetics of Parkinson’s disease: contributions and global trends. J Hum Genet. 2023;68(3):125–30.

    Article  PubMed  Google Scholar 

  50. Schapira AH, Cooper JM, Dexter D, Clark JB, Jenner P, Marsden CD. Mitochondrial complex I deficiency in Parkinson’s disease. J Neurochem. 1990;54(3):823–7.

    Article  CAS  PubMed  Google Scholar 

  51. Chen C, Mossman E, Malko P, et al. Astrocytic changes in mitochondrial oxidative phosphorylation protein levels in Parkinson’s disease. Mov Disord. 2022;37(2):302–14.

    Article  CAS  PubMed  Google Scholar 

  52. Sarkar S, Malovic E, Harishchandra DS, et al. Mitochondrial impairment in microglia amplifies NLRP3 inflammasome Proinflammatory signaling in cell culture and animal models of Parkinson’s disease. NPJ Parkinsons Dis. 2017;3:30.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Dhillon AS, Tarbutton GL, Levin JL, et al. Pesticide/environmental exposures and Parkinson’s disease in East Texas. J Agromedicine. 2008;13(1):37–48.

    Article  PubMed  Google Scholar 

  54. Pouchieu C, Piel C, Carles C, et al. Pesticide use in agriculture and Parkinson’s disease in the AGRICAN cohort study. Int J Epidemiol. 2018;47(1):299–310.

    Article  PubMed  Google Scholar 

  55. Won JH, Park S, Hong S, Son S, Yu JW. Rotenone-induced impairment of mitochondrial electron transport chain confers a selective priming signal for NLRP3 inflammasome activation. J Biol Chem. 2015;290(45):27425–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Yoshino H, Nakagawa-Hattori Y, Kondo T, Mizuno Y. Mitochondrial complex I and II activities of lymphocytes and platelets in Parkinson’s disease. J Neural Transm Park Dis Dement Sect. 1992;4(1):27–34.

    Article  CAS  PubMed  Google Scholar 

  57. Müftüoglu M, Elibol B, Dalmizrak O, et al. Mitochondrial complex I and IV activities in leukocytes from patients with parkin mutations. Mov Disord. 2004;19(5):544–8.

    Article  PubMed  Google Scholar 

  58. Salemi M, Cosentino F, Lanza G, et al. mRNA expression profiling of mitochondrial subunits in subjects with Parkinson’s disease. Arch Med Sci. 2023;19(3):678–86.

    CAS  PubMed  Google Scholar 

  59. Betarbet R, Sherer TB, MacKenzie G, Garcia-Osuna M, Panov AV, Greenamyre JT. Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat Neurosci. 2000;3(12):1301–6.

    Article  CAS  PubMed  Google Scholar 

  60. Langston JW, Ballard P, Tetrud JW, Irwin I. Chronic parkinsonism in humans due to a product of meperidine-analog synthesis. Science. 1983;219(4587):979–80.

    Article  CAS  PubMed  Google Scholar 

  61. Cooper O, Seo H, Andrabi S, et al. Pharmacological rescue of mitochondrial deficits in iPSC-derived neural cells from patients with Familial Parkinson’s disease. Sci Transl Med. 2012;4(141):141ra190.

    Article  Google Scholar 

  62. Gautier CA, Kitada T, Shen J. Loss of PINK1 causes mitochondrial functional defects and increased sensitivity to oxidative stress. Proc Natl Acad Sci U S A. 2008;105(32):11364–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Patki G, Che Y, Lau YS. Mitochondrial dysfunction in the striatum of aged chronic mouse model of Parkinson’s disease. Front Aging Neurosci. 2009;1:3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Chung YC, Baek JY, Kim SR, et al. Capsaicin prevents degeneration of dopamine neurons by inhibiting glial activation and oxidative stress in the MPTP model of Parkinson’s disease. Exp Mol Med. 2017;49(3):e298.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Liu J, Liang Y, Meng Q, et al. Antagonism of β-arrestins in IL-4-driven microglia reactivity via the Samd4/mTOR/OXPHOS axis in Parkinson’s disease. Sci Adv. 2024;10(34):eadn4845.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Deus CM, Pereira SP, Cunha-Oliveira T, Pereira FB, Raimundo N, Oliveira PJ. Mitochondrial remodeling in human skin fibroblasts from sporadic male Parkinson’s disease patients uncovers metabolic and mitochondrial bioenergetic defects. Biochim Biophys Acta Mol Basis Dis. 2020;1866(3):165615.

    Article  CAS  PubMed  Google Scholar 

  67. Haylett W, Swart C, van der Westhuizen F, et al. Altered mitochondrial respiration and other features of mitochondrial function in Parkin-Mutant fibroblasts from Parkinson’s disease patients. Parkinsons Dis. 2016;2016:1819209.

    PubMed  PubMed Central  Google Scholar 

  68. Ugras S, Daniels MJ, Fazelinia H, et al. Induction of the Immunoproteasome subunit Lmp7 links proteostasis and immunity in α-Synuclein aggregation disorders. EBioMedicine. 2018;31:307–19.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Schirinzi T, Salvatori I, Zenuni H, et al. Pattern of mitochondrial respiration in peripheral blood cells of patients with Parkinson’s disease. Int J Mol Sci. 2022;23(18):10863.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Annesley SJ, Lay ST, De Piazza SW, et al. Immortalized Parkinson’s disease lymphocytes have enhanced mitochondrial respiratory activity. Dis Model Mech. 2016;9(11):1295–305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kramer PA, Ravi S, Chacko B, Johnson MS, Darley-Usmar VM. A review of the mitochondrial and glycolytic metabolism in human platelets and leukocytes: implications for their use as bioenergetic biomarkers. Redox Biol. 2014;2:206–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Turrens JF. Mitochondrial formation of reactive oxygen species. J Physiol. 2003;552(Pt 2):335–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Guo C, Sun L, Chen X, Zhang D. Oxidative stress, mitochondrial damage and neurodegenerative diseases. Neural Regen Res. 2013;8(21):2003–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Azzouz D, Khan MA, Palaniyar N. ROS induces NETosis by oxidizing DNA and initiating DNA repair. Cell Death Discovery. 2021;7(1):113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Spencer JP, Jenner P, Daniel SE, Lees AJ, Marsden DC, Halliwell B. Conjugates of catecholamines with cysteine and GSH in Parkinson’s disease: possible mechanisms of formation involving reactive oxygen species. J Neurochem. 1998;71(5):2112–22.

    Article  CAS  PubMed  Google Scholar 

  76. Angeles DC, Gan BH, Onstead L, et al. Mutations in LRRK2 increase phosphorylation of Peroxiredoxin 3 exacerbating oxidative stress-induced neuronal death. Hum Mutat. 2011;32(12):1390–7.

    Article  CAS  PubMed  Google Scholar 

  77. Ehrhart J, Zeevalk GD. Cooperative interaction between ascorbate and glutathione during mitochondrial impairment in mesencephalic cultures. J Neurochem. 2003;86(6):1487–97.

    Article  CAS  PubMed  Google Scholar 

  78. Heo HY, Park JM, Kim CH, Han BS, Kim KS, Seol W. LRRK2 enhances oxidative stress-induced neurotoxicity via its kinase activity. Exp Cell Res. 2010;316(4):649–56.

    Article  CAS  PubMed  Google Scholar 

  79. Irrcher I, Aleyasin H, Seifert EL, et al. Loss of the Parkinson’s disease-linked gene DJ-1 perturbs mitochondrial dynamics. Hum Mol Genet. 2010;19(19):3734–46.

    Article  CAS  PubMed  Google Scholar 

  80. Billia F, Hauck L, Konecny F, Rao V, Shen J, Mak TW. PTEN-inducible kinase 1 (PINK1)/Park6 is indispensable for normal heart function. Proc Natl Acad Sci U S A. 2011;108(23):9572–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Zhong Z, Umemura A, Sanchez-Lopez E, et al. NF-κB restricts inflammasome activation via elimination of damaged mitochondria. Cell. 2016;164(5):896–910.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Mortiboys H, Thomas KJ, Koopman WJ, et al. Mitochondrial function and morphology are impaired in parkin-mutant fibroblasts. Ann Neurol. 2008;64(5):555–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Zambon F, Cherubini M, Fernandes HJR, et al. Cellular α-synuclein pathology is associated with bioenergetic dysfunction in Parkinson’s iPSC-derived dopamine neurons. Hum Mol Genet. 2019;28(12):2001–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Deas E, Cremades N, Angelova PR, et al. Alpha-Synuclein oligomers interact with metal ions to induce oxidative stress and neuronal death in Parkinson’s disease. Antioxid Redox Signal. 2016;24(7):376–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Gusdon AM, Zhu J, Van Houten B, Chu CT. ATP13A2 regulates mitochondrial bioenergetics through macroautophagy. Neurobiol Dis. 2012;45(3):962–72.

    Article  CAS  PubMed  Google Scholar 

  86. Hunot S, Boissière F, Faucheux B, et al. Nitric oxide synthase and neuronal vulnerability in Parkinson’s disease. Neuroscience. 1996;72(2):355–63.

    Article  CAS  PubMed  Google Scholar 

  87. Knott C, Stern G, Wilkin GP. Inflammatory regulators in Parkinson’s disease: iNOS, lipocortin-1, and cyclooxygenases-1 and– 2. Mol Cell Neurosci. 2000;16(6):724–39.

    Article  CAS  PubMed  Google Scholar 

  88. Zhang W, Wang T, Pei Z, et al. Aggregated alpha-synuclein activates microglia: a process leading to disease progression in Parkinson’s disease. Faseb J. 2005;19(6):533–42.

    Article  CAS  PubMed  Google Scholar 

  89. Liu B, Gao HM, Wang JY, Jeohn GH, Cooper CL, Hong JS. Role of nitric oxide in inflammation-mediated neurodegeneration. Ann N Y Acad Sci. 2002;962:318–31.

    Article  CAS  PubMed  Google Scholar 

  90. Bido S, Muggeo S, Massimino L, et al. Microglia-specific overexpression of α-synuclein leads to severe dopaminergic neurodegeneration by phagocytic exhaustion and oxidative toxicity. Nat Commun. 2021;12(1):6237.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Thomas MP, Chartrand K, Reynolds A, Vitvitsky V, Banerjee R, Gendelman HE. Ion channel Blockade attenuates aggregated alpha synuclein induction of microglial reactive oxygen species: relevance for the pathogenesis of Parkinson’s disease. J Neurochem. 2007;100(2):503–19.

    Article  CAS  PubMed  Google Scholar 

  92. Shaikh SB, Nicholson LF. Effects of chronic low dose rotenone treatment on human microglial cells. Mol Neurodegener. 2009;4:55.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Yu YX, Li YP, Gao F, et al. Vitamin K2 suppresses rotenone-induced microglial activation in vitro. Acta Pharmacol Sin. 2016;37(9):1178–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Nakahira K, Haspel JA, Rathinam VA, et al. Autophagy proteins regulate innate immune responses by inhibiting the release of mitochondrial DNA mediated by the NALP3 inflammasome. Nat Immunol. 2011;12(3):222–30.

    Article  CAS  PubMed  Google Scholar 

  95. Zhou R, Yazdi AS, Menu P, Tschopp J. A role for mitochondria in NLRP3 inflammasome activation. Nature. 2011;469(7329):221–5.

    Article  CAS  PubMed  Google Scholar 

  96. Prigione A, Begni B, Galbussera A, et al. Oxidative stress in peripheral blood mononuclear cells from patients with Parkinson’s disease: negative correlation with Levodopa dosage. Neurobiol Dis. 2006;23(1):36–43.

    Article  CAS  PubMed  Google Scholar 

  97. Paiva CN, Bozza MT. Are reactive oxygen species always detrimental to pathogens? Antioxid Redox Signal. 2014;20(6):1000–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Zorova LD, Popkov VA, Plotnikov EY, et al. Mitochondrial membrane potential. Anal Biochem. 2018;552:50–9.

    Article  CAS  PubMed  Google Scholar 

  99. Gandhi S, Wood-Kaczmar A, Yao Z, et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Mol Cell. 2009;33(5):627–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. McCoy MK, Cookson MR. DJ-1 regulation of mitochondrial function and autophagy through oxidative stress. Autophagy. 2011;7(5):531–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Choi ML, Chappard A, Singh BP, et al. Pathological structural conversion of α-synuclein at the mitochondria induces neuronal toxicity. Nat Neurosci. 2022;25(9):1134–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Lind-Holm Mogensen F, Sousa C, Ameli C, et al. PARK7/DJ-1 deficiency impairs microglial activation in response to LPS-induced inflammation. J Neuroinflammation. 2024;21(1):174.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Imberechts D, Kinnart I, Wauters F, et al. DJ-1 is an essential downstream mediator in PINK1/parkin-dependent mitophagy. Brain. 2022;145(12):4368–84.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Qadri R, Namdeo M, Behari M, Goyal V, Sharma S, Mukhopadhyay AK. Alterations in mitochondrial membrane potential in peripheral blood mononuclear cells in Parkinson’s disease: potential for a novel biomarker. Restor Neurol Neurosci. 2018;36(6):719–27.

    CAS  PubMed  Google Scholar 

  105. Malpartida AB, Williamson M, Narendra DP, Wade-Martins R, Ryan BJ. Mitochondrial dysfunction and mitophagy in Parkinson’s disease: from mechanism to therapy. Trends Biochem Sci. 2021;46(4):329–43.

    Article  CAS  PubMed  Google Scholar 

  106. Hsieh CH, Shaltouki A, Gonzalez AE, et al. Functional impairment in Miro degradation and mitophagy is a shared feature in Familial and sporadic Parkinson’s disease. Cell Stem Cell. 2016;19(6):709–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Singh F, Prescott AR, Rosewell P, Ball G, Reith AD, Ganley IG. Pharmacological rescue of impaired mitophagy in Parkinson’s disease-related LRRK2 G2019S knock-in mice. Elife 2021;10.

  108. Walter J, Bolognin S, Antony PMA, et al. Neural stem cells of Parkinson’s disease patients exhibit aberrant mitochondrial morphology and functionality. Stem Cell Rep. 2019;12(5):878–89.

    Article  CAS  Google Scholar 

  109. Kim MJ, Yoon JH, Ryu JH. Mitophagy: a balance regulator of NLRP3 inflammasome activation. BMB Rep. 2016;49(10):529–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Zhao W, Li Y, Jia L, Pan L, Li H, Du J. Atg5 deficiency-mediated mitophagy aggravates cardiac inflammation and injury in response to angiotensin II. Free Radic Biol Med. 2014;69:108–15.

    Article  CAS  PubMed  Google Scholar 

  111. Cheng J, Liao Y, Dong Y, et al. Microglial autophagy defect causes Parkinson disease-like symptoms by accelerating inflammasome activation in mice. Autophagy. 2020;16(12):2193–205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Han X, Xu T, Fang Q, et al. Quercetin hinders microglial activation to alleviate neurotoxicity via the interplay between NLRP3 inflammasome and mitophagy. Redox Biol. 2021;44:102010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Qiu J, Chen Y, Zhuo J, et al. Urolithin A promotes mitophagy and suppresses NLRP3 inflammasome activation in lipopolysaccharide-induced BV2 microglial cells and MPTP-induced Parkinson’s disease model. Neuropharmacology. 2022;207:108963.

    Article  CAS  PubMed  Google Scholar 

  114. Copeland WC. The mitochondrial DNA polymerase in health and disease. Subcell Biochem. 2010;50:211–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Gu G, Reyes PE, Golden GT, et al. Mitochondrial DNA deletions/rearrangements in Parkinson disease and related neurodegenerative disorders. J Neuropathol Exp Neurol. 2002;61(7):634–9.

    Article  CAS  PubMed  Google Scholar 

  116. Nido GS, Dölle C, Flønes I, et al. Ultradeep mapping of neuronal mitochondrial deletions in Parkinson’s disease. Neurobiol Aging. 2018;63:120–7.

    Article  CAS  PubMed  Google Scholar 

  117. Bender A, Krishnan KJ, Morris CM, et al. High levels of mitochondrial DNA deletions in substantia Nigra neurons in aging and Parkinson disease. Nat Genet. 2006;38(5):515–7.

    Article  CAS  PubMed  Google Scholar 

  118. Qi R, Sammler E, Gonzalez-Hunt CP, et al. A blood-based marker of mitochondrial DNA damage in Parkinson’s disease. Sci Transl Med. 2023;15(711):eabo1557.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Nakahira K, Hisata S, Choi AM. The roles of mitochondrial Damage-Associated molecular patterns in diseases. Antioxid Redox Signal. 2015;23(17):1329–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Zanini G, Selleri V, Lopez Domenech S, et al. Mitochondrial DNA as inflammatory DAMP: a warning of an aging immune system? Biochem Soc Trans. 2023;51(2):735–45.

    Article  CAS  PubMed  Google Scholar 

  121. Hinkle JT, Patel J, Panicker N, et al. STING mediates neurodegeneration and neuroinflammation in nigrostriatal α-synucleinopathy. Proc Natl Acad Sci U S A. 2022;119(15):e2118819119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Hancock-Cerutti W, Wu Z, Xu P et al. ER-lysosome lipid transfer protein VPS13C/PARK23 prevents aberrant mtDNA-dependent STING signaling. J Cell Biol 2022;221(7).

  123. Kumar N, Leonzino M, Hancock-Cerutti W, et al. VPS13A and VPS13C are lipid transport proteins differentially localized at ER contact sites. J Cell Biol. 2018;217(10):3625–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Moors T, Paciotti S, Chiasserini D, et al. Lysosomal dysfunction and α-Synuclein aggregation in Parkinson’s disease: diagnostic links. Mov Disord. 2016;31(6):791–801.

    Article  CAS  PubMed  Google Scholar 

  125. Mazzulli JR, 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 U S A. 2016;113(7):1931–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Erb ML, Moore DJ. LRRK2 and the endolysosomal system in Parkinson’s disease. J Parkinsons Dis. 2020;10(4):1271–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Avenali M, Blandini F, Cerri S. Glucocerebrosidase defects as a major risk factor for Parkinson’s disease. Front Aging Neurosci. 2020;12:97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Navarro-Romero A, Montpeyó M, Martinez-Vicente M. The Emerging Role of the Lysosome in Parkinson’s Disease. Cells 2020;9(11).

  129. Dzamko N, Halliday GM. An emerging role for LRRK2 in the immune system. Biochem Soc Trans. 2012;40(5):1134–9.

    Article  CAS  PubMed  Google Scholar 

  130. Mullin S, Stokholm MG, Hughes D, et al. Brain microglial activation increased in glucocerebrosidase (GBA) mutation carriers without Parkinson’s disease. Mov Disord. 2021;36(3):774–9.

    Article  CAS  PubMed  Google Scholar 

  131. Levine B, Mizushima N, Virgin HW. Autophagy in immunity and inflammation. Nature. 2011;469(7330):323–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Feng Y, Chen Y, Wu X, et al. Interplay of energy metabolism and autophagy. Autophagy. 2024;20(1):4–14.

    Article  CAS  PubMed  Google Scholar 

  133. Cui B, Lin H, Yu J, Yu J, Hu Z. Autophagy and the immune response. Adv Exp Med Biol. 2019;1206:595–634.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Anglade P, Vyas S, Javoy-Agid F, et al. Apoptosis and autophagy in nigral neurons of patients with Parkinson’s disease. Histol Histopathol. 1997;12(1):25–31.

    CAS  PubMed  Google Scholar 

  135. Bellomo G, Paciotti S, Gatticchi L, Parnetti L. The vicious cycle between α-synuclein aggregation and autophagic-lysosomal dysfunction. Mov Disord. 2020;35(1):34–44.

    Article  CAS  PubMed  Google Scholar 

  136. Hou X, Watzlawik JO, Fiesel FC, Springer W. Autophagy in Parkinson’s disease. J Mol Biol. 2020;432(8):2651–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Alvarez-Erviti L, Rodriguez-Oroz MC, Cooper JM, et al. Chaperone-mediated autophagy markers in Parkinson disease brains. Arch Neurol. 2010;67(12):1464–72.

    Article  PubMed  Google Scholar 

  138. Chen L, Xie Z, Turkson S, Zhuang X. A53T human α-synuclein overexpression in Transgenic mice induces pervasive mitochondria macroautophagy defects preceding dopamine neuron degeneration. J Neurosci. 2015;35(3):890–905.

    Article  PubMed  Google Scholar 

  139. Plowey ED, Cherra SJ 3rd, Liu YJ, Chu CT. Role of autophagy in G2019S-LRRK2-associated neurite shortening in differentiated SH-SY5Y cells. J Neurochem. 2008;105(3):1048–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Wang B, Cai Z, Tao K, et al. Essential control of mitochondrial morphology and function by chaperone-mediated autophagy through degradation of PARK7. Autophagy. 2016;12(8):1215–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Zavodszky E, Seaman MN, Moreau K, et al. Mutation in VPS35 associated with Parkinson’s disease impairs WASH complex association and inhibits autophagy. Nat Commun. 2014;5:3828.

    Article  CAS  PubMed  Google Scholar 

  142. Lee HJ, Suk JE, Bae EJ, Lee SJ. Clearance and deposition of extracellular alpha-synuclein aggregates in microglia. Biochem Biophys Res Commun. 2008;372(3):423–8.

    Article  CAS  PubMed  Google Scholar 

  143. Tu HY, Yuan BS, Hou XO, et al. α-synuclein suppresses microglial autophagy and promotes neurodegeneration in a mouse model of Parkinson’s disease. Aging Cell. 2021;20(12):e13522.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Nash Y, Schmukler E, Trudler D, Pinkas-Kramarski R, Frenkel D. DJ-1 deficiency impairs autophagy and reduces alpha-synuclein phagocytosis by microglia. J Neurochem. 2017;143(5):584–94.

    Article  CAS  PubMed  Google Scholar 

  145. Xu Y, Propson NE, Du S, Xiong W, Zheng H. Autophagy deficiency modulates microglial lipid homeostasis and aggravates Tau pathology and spreading. Proc Natl Acad Sci U S A. 2021;118:27.

    Article  Google Scholar 

  146. Henderson MX, Sengupta M, Trojanowski JQ, Lee VMY. Alzheimer’s disease Tau is a prominent pathology in LRRK2 Parkinson’s disease. Acta Neuropathol Commun. 2019;7(1):183.

    Article  PubMed  PubMed Central  Google Scholar 

  147. Gao YL, Wang N, Sun FR, Cao XP, Zhang W, Yu JT. Tau in neurodegenerative disease. Ann Transl Med. 2018;6(10):175.

    Article  PubMed  PubMed Central  Google Scholar 

  148. Qin Y, Qiu J, Wang P, et al. Impaired autophagy in microglia aggravates dopaminergic neurodegeneration by regulating NLRP3 inflammasome activation in experimental models of Parkinson’s disease. Brain Behav Immun. 2021;91:324–38.

    Article  CAS  PubMed  Google Scholar 

  149. Vavilova JD, Boyko AA, Troyanova NI et al. Alterations in proteostasis system components in peripheral blood mononuclear cells in Parkinson disease: focusing on the HSP70 and p62 levels. Biomolecules 2022;12(4).

  150. El Haddad S, Serrano A, Moal F, et al. Disturbed expression of autophagy genes in blood of Parkinson’s disease patients. Gene. 2020;738:144454.

    Article  PubMed  Google Scholar 

  151. Wu G, Wang X, Feng X, et al. Altered expression of autophagic genes in the peripheral leukocytes of patients with sporadic Parkinson’s disease. Brain Res. 2011;1394:105–11.

    Article  CAS  PubMed  Google Scholar 

  152. Miki Y, Shimoyama S, Kon T, et al. Alteration of autophagy-related proteins in peripheral blood mononuclear cells of patients with Parkinson’s disease. Neurobiol Aging. 2018;63:33–43.

    Article  CAS  PubMed  Google Scholar 

  153. Zhang Z, Singh R, Aschner M. Methods for the detection of autophagy in mammalian cells. Curr Protoc Toxicol. 2016;69:201221–6.

    Article  Google Scholar 

  154. Erustes AG, Stefani FY, Terashima JY, et al. Overexpression of α-synuclein in an astrocyte cell line promotes autophagy Inhibition and apoptosis. J Neurosci Res. 2018;96(1):160–71.

    Article  CAS  PubMed  Google Scholar 

  155. Prigione A, Piazza F, Brighina L, et al. Alpha-synuclein nitration and autophagy response are induced in peripheral blood cells from patients with Parkinson disease. Neurosci Lett. 2010;477(1):6–10.

    Article  CAS  PubMed  Google Scholar 

  156. Kumari U, Tan EK. Leucine-rich repeat kinase 2-linked Parkinson’s disease: clinical and molecular findings. J Mov Disord. 2010;3(2):25–31.

    Article  PubMed  PubMed Central  Google Scholar 

  157. Wallings RL, Tansey MG. LRRK2 regulation of immune-pathways and inflammatory disease. Biochem Soc Trans. 2019;47(6):1581–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Steger M, Tonelli F, Ito G et al. Phosphoproteomics reveals that Parkinson’s disease kinase LRRK2 regulates a subset of Rab GTPases. Elife 2016;5.

  159. Vieira OV. Rab3a and Rab10 are regulators of lysosome exocytosis and plasma membrane repair. Small GTPases. 2018;9(4):349–51.

    Article  CAS  PubMed  Google Scholar 

  160. Cardoso CM, Jordao L, Vieira OV. Rab10 regulates phagosome maturation and its overexpression rescues Mycobacterium-containing phagosomes maturation. Traffic. 2010;11(2):221–35.

    Article  CAS  PubMed  Google Scholar 

  161. Wandinger-Ness A, Zerial M. Rab proteins and the compartmentalization of the endosomal system. Cold Spring Harb Perspect Biol. 2014;6(11):a022616.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Parisiadou L, Cai H. LRRK2 function on actin and microtubule dynamics in Parkinson disease. Commun Integr Biol. 2010;3(5):396–400.

    Article  PubMed  PubMed Central  Google Scholar 

  163. Bonet-Ponce L, Beilina A, Williamson CD et al. LRRK2 mediates tubulation and vesicle sorting from lysosomes. Sci Adv 2020;6(46).

  164. Xiong Y, Yuan C, Chen R, Dawson TM, Dawson VL. ArfGAP1 is a GTPase activating protein for LRRK2: reciprocal regulation of ArfGAP1 by LRRK2. J Neurosci. 2012;32(11):3877–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Beilina A, Bonet-Ponce L, Kumaran R, et al. The Parkinson’s disease protein LRRK2 interacts with the GARP complex to promote retrograde transport to the trans-Golgi network. Cell Rep. 2020;31(5):107614.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Cho HJ, Yu J, Xie C, et al. Leucine-rich repeat kinase 2 regulates Sec16A at ER exit sites to allow ER-Golgi export. Embo J. 2014;33(20):2314–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Belluzzi E, Gonnelli A, Cirnaru MD, et al. LRRK2 phosphorylates pre-synaptic N-ethylmaleimide sensitive fusion (NSF) protein enhancing its ATPase activity and SNARE complex disassembling rate. Mol Neurodegener. 2016;11:1.

    Article  PubMed  PubMed Central  Google Scholar 

  168. Wallings RL, Mark JR, Staley HA, et al. ASO-mediated knockdown or kinase Inhibition of G2019S-Lrrk2 modulates lysosomal tubule-associated antigen presentation in macrophages. Mol Ther Nucleic Acids. 2023;34:102064.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Hierro A, Rojas AL, Rojas R, et al. Functional architecture of the retromer cargo-recognition complex. Nature. 2007;449(7165):1063–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Zimprich A, Benet-Pagès A, Struhal W, et al. A mutation in VPS35, encoding a subunit of the retromer complex, causes late-onset Parkinson disease. Am J Hum Genet. 2011;89(1):168–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Vilariño-Güell C, Wider C, Ross OA, et al. VPS35 mutations in Parkinson disease. Am J Hum Genet. 2011;89(1):162–7.

    Article  PubMed  PubMed Central  Google Scholar 

  172. Niu M, Zhao F, Bondelid K, et al. VPS35 D620N knockin mice recapitulate Cardinal features of Parkinson’s disease. Aging Cell. 2021;20(5):e13347.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Chen X, Kordich JK, Williams ET, et al. Parkinson’s disease-linked D620N VPS35 knockin mice manifest Tau neuropathology and dopaminergic neurodegeneration. Proc Natl Acad Sci U S A. 2019;116(12):5765–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Mir R, Tonelli F, Lis P, et al. The Parkinson’s disease VPS35[D620N] mutation enhances LRRK2-mediated Rab protein phosphorylation in mouse and human. Biochem J. 2018;475(11):1861–83.

    Article  CAS  PubMed  Google Scholar 

  175. Yin J, Liu X, He Q, Zhou L, Yuan Z, Zhao S. Vps35-dependent recycling of Trem2 regulates microglial function. Traffic. 2016;17(12):1286–96.

    Article  CAS  PubMed  Google Scholar 

  176. Pal P, Taylor M, Lam PY, et al. Parkinson’s VPS35[D620N] mutation induces LRRK2-mediated lysosomal association of RILPL1 and TMEM55B. Sci Adv. 2023;9(50):eadj1205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Bu M, Follett J, Deng I, et al. Inhibition of LRRK2 kinase activity rescues deficits in striatal dopamine physiology in VPS35 p.D620N knock-in mice. NPJ Parkinsons Dis. 2023;9(1):167.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Moors TE, Paciotti S, Ingrassia A, et al. Characterization of brain lysosomal activities in GBA-Related and sporadic Parkinson’s disease and dementia with lewy bodies. Mol Neurobiol. 2019;56(2):1344–55.

    Article  CAS  PubMed  Google Scholar 

  179. Nelson MP, Boutin M, Tse TE, et al. The lysosomal enzyme alpha-Galactosidase A is deficient in Parkinson’s disease brain in association with the pathologic accumulation of alpha-synuclein. Neurobiol Dis. 2018;110:68–81.

    Article  CAS  PubMed  Google Scholar 

  180. Yun SP, Kim D, Kim S, et al. α-Synuclein accumulation and GBA deficiency due to L444P GBA mutation contributes to MPTP-induced parkinsonism. Mol Neurodegener. 2018;13(1):1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Drobny A, Boros FA, Balta D, et al. Reciprocal effects of alpha-synuclein aggregation and lysosomal homeostasis in synucleinopathy models. Transl Neurodegener. 2023;12(1):31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Yadavalli N, Ferguson SM. LRRK2 suppresses lysosome degradative activity in macrophages and microglia through MiT-TFE transcription factor Inhibition. Proc Natl Acad Sci U S A. 2023;120(31):e2303789120.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Henry AG, Aghamohammadzadeh S, Samaroo H, et al. Pathogenic LRRK2 mutations, through increased kinase activity, produce enlarged lysosomes with reduced degradative capacity and increase ATP13A2 expression. Hum Mol Genet. 2015;24(21):6013–28.

    Article  CAS  PubMed  Google Scholar 

  184. Karabova MK, Cowley SA, James WS. The role of human microglia and microglial LRRK2 in Tau pathogenesis. Alzheimer’s Dement. 2022;18(S4):e069393.

    Article  Google Scholar 

  185. Mindell JA. Lysosomal acidification mechanisms. Annu Rev Physiol. 2012;74:69–86.

    Article  CAS  PubMed  Google Scholar 

  186. Lo CH, Zeng J. Defective lysosomal acidification: a new prognostic marker and therapeutic target for neurodegenerative diseases. Transl Neurodegener. 2023;12(1):29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Dehay B, Ramirez A, Martinez-Vicente M, et al. Loss of P-type ATPase ATP13A2/PARK9 function induces general lysosomal deficiency and leads to Parkinson disease neurodegeneration. Proc Natl Acad Sci U S A. 2012;109(24):9611–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Di Maio R, Hoffman EK, Rocha EM et al. LRRK2 activation in idiopathic Parkinson’s disease. Sci Transl Med 2018;10(451).

  189. Schapansky J, Khasnavis S, DeAndrade MP, et al. Familial knockin mutation of LRRK2 causes lysosomal dysfunction and accumulation of endogenous insoluble α-synuclein in neurons. Neurobiol Dis. 2018;111:26–35.

    Article  CAS  PubMed  Google Scholar 

  190. Haenseler W, Zambon F, Lee H, et al. Excess α-synuclein compromises phagocytosis in iPSC-derived macrophages. Sci Rep. 2017;7(1):9003.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Campagno KE, Mitchell CH. The P2X(7) Receptor in Microglial Cells Modulates the Endolysosomal Axis, Autophagy, and Phagocytosis. Front Cell Neurosci. 2021;15:645244.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Filipello F, You SF, Mirfakhar FS, et al. Defects in lysosomal function and lipid metabolism in human microglia harboring a TREM2 loss of function mutation. Acta Neuropathol. 2023;145(6):749–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Lill CM, Rengmark A, Pihlstrøm L, et al. The role of TREM2 R47H as a risk factor for Alzheimer’s disease, frontotemporal Lobar degeneration, amyotrophic lateral sclerosis, and Parkinson’s disease. Alzheimers Dement. 2015;11(12):1407–16.

    Article  PubMed  Google Scholar 

  194. Lu Y, Liu W, Wang X. TREM2 variants and risk of Alzheimer’s disease: a meta-analysis. Neurol Sci. 2015;36(10):1881–8.

    Article  PubMed  Google Scholar 

  195. Rayaprolu S, Mullen B, Baker M, et al. TREM2 in neurodegeneration: evidence for association of the p.R47H variant with frontotemporal dementia and Parkinson’s disease. Mol Neurodegener. 2013;8:19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Liu G, Liu Y, Jiang Q, et al. Convergent genetic and expression datasets highlight TREM2 in Parkinson’s disease susceptibility. Mol Neurobiol. 2016;53(7):4931–8.

    Article  CAS  PubMed  Google Scholar 

  197. Mengel D, Thelen M, Balzer-Geldsetzer M, et al. TREM2 rare variant p.R47H is not associated with Parkinson’s disease. Parkinsonism Relat Disord. 2016;23:109–11.

    Article  PubMed  Google Scholar 

  198. Guo Y, Wei X, Yan H, et al. TREM2 deficiency aggravates α-synuclein-induced neurodegeneration and neuroinflammation in Parkinson’s disease models. Faseb J. 2019;33(11):12164–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Tagliatti E, Desiato G, Mancinelli S, et al. Trem2 expression in microglia is required to maintain normal neuronal bioenergetics during development. Immunity. 2024;57(1):86–e105109.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Vest RT, Chou CC, Zhang H, et al. Small molecule C381 targets the lysosome to reduce inflammation and ameliorate disease in models of neurodegeneration. Proc Natl Acad Sci U S A. 2022;119(11):e2121609119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Sidransky E, Nalls MA, Aasly JO, et al. Multicenter analysis of glucocerebrosidase mutations in Parkinson’s disease. N Engl J Med. 2009;361(17):1651–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Anheim M, Elbaz A, Lesage S, et al. Penetrance of Parkinson disease in glucocerebrosidase gene mutation carriers. Neurology. 2012;78(6):417–20.

    Article  CAS  PubMed  Google Scholar 

  203. Mazzulli JR, Zunke F, Tsunemi T, et al. Activation of β-Glucocerebrosidase reduces pathological α-Synuclein and restores lysosomal function in Parkinson’s patient midbrain neurons. J Neurosci. 2016;36(29):7693–706.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Rocha EM, Smith GA, Park E, et al. Sustained systemic glucocerebrosidase Inhibition induces brain α-Synuclein aggregation, microglia and complement C1q activation in mice. Antioxid Redox Signal. 2015;23(6):550–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Ginns EI, Mak SK, Ko N, et al. Neuroinflammation and α-synuclein accumulation in response to glucocerebrosidase deficiency are accompanied by synaptic dysfunction. Mol Genet Metab. 2014;111(2):152–62.

    Article  CAS  PubMed  Google Scholar 

  206. Soria FN, Engeln M, Martinez-Vicente M, et al. Glucocerebrosidase deficiency in dopaminergic neurons induces microglial activation without neurodegeneration. Hum Mol Genet. 2017;26(14):2603–15.

    Article  CAS  PubMed  Google Scholar 

  207. Mus L, Siani F, Giuliano C, Ghezzi C, Cerri S, Blandini F. Development and biochemical characterization of a mouse model of Parkinson’s disease bearing defective glucocerebrosidase activity. Neurobiol Dis. 2019;124:289–96.

    Article  CAS  PubMed  Google Scholar 

  208. Sanyal A, DeAndrade MP, Novis HS, et al. Lysosome and inflammatory defects in GBA1-Mutant astrocytes are normalized by LRRK2 Inhibition. Mov Disord. 2020;35(5):760–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Brunialti E, Villa A, Mekhaeil M, et al. Inhibition of microglial β-glucocerebrosidase hampers the microglia-mediated antioxidant and protective response in neurons. J Neuroinflammation. 2021;18(1):220.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Kedariti M, Frattini E, Baden P, et al. LRRK2 kinase activity regulates GCase level and enzymatic activity differently depending on cell type in Parkinson’s disease. NPJ Parkinsons Dis. 2022;8(1):92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Ysselstein D, Nguyen M, Young TJ, et al. LRRK2 kinase activity regulates lysosomal glucocerebrosidase in neurons derived from Parkinson’s disease patients. Nat Commun. 2019;10(1):5570.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Atashrazm F, Hammond D, Perera G, et al. Reduced glucocerebrosidase activity in monocytes from patients with Parkinson’s disease. Sci Rep. 2018;8(1):15446.

    Article  PubMed  PubMed Central  Google Scholar 

  213. Aflaki E, Moaven N, Borger DK, et al. Lysosomal storage and impaired autophagy lead to inflammasome activation in gaucher macrophages. Aging Cell. 2016;15(1):77–88.

    Article  CAS  PubMed  Google Scholar 

  214. Alcalay RN, Levy OA, Waters CC, et al. Glucocerebrosidase activity in Parkinson’s disease with and without GBA mutations. Brain. 2015;138(Pt 9):2648–58.

    Article  PubMed  PubMed Central  Google Scholar 

  215. Kim HJ, Jeon B, Song J, Lee WW, Park H, Shin CW. Leukocyte glucocerebrosidase and β-hexosaminidase activity in sporadic and genetic Parkinson disease. Parkinsonism Relat Disord. 2016;23:99–101.

    Article  PubMed  Google Scholar 

  216. Grabowski GA. Gaucher disease: gene frequencies and genotype/phenotype correlations. Genet Test. 1997;1(1):5–12.

    Article  CAS  PubMed  Google Scholar 

  217. Grace ME, Newman KM, Scheinker V, Berg-Fussman A, Grabowski GA. Analysis of human acid beta-glucosidase by site-directed mutagenesis and heterologous expression. J Biol Chem. 1994;269(3):2283–91.

    Article  CAS  PubMed  Google Scholar 

  218. Calabresi P, Mechelli A, Natale G, Volpicelli-Daley L, Di Lazzaro G, Ghiglieri V. Alpha-synuclein in Parkinson’s disease and other synucleinopathies: from overt neurodegeneration back to early synaptic dysfunction. Cell Death Dis. 2023;14(3):176.

    Article  PubMed  PubMed Central  Google Scholar 

  219. Vilchez D, Saez I, Dillin A. The role of protein clearance mechanisms in organismal ageing and age-related diseases. Nat Commun. 2014;5:5659.

    Article  CAS  PubMed  Google Scholar 

  220. Moehle EA, Shen K, Dillin A. Mitochondrial proteostasis in the context of cellular and organismal health and aging. J Biol Chem. 2019;294(14):5396–407.

    Article  CAS  PubMed  Google Scholar 

  221. Jadiya P, Tomar D. Mitochondrial protein quality control mechanisms. Genes (Basel) 2020;11(5).

  222. McNaught KS, Belizaire R, Isacson O, Jenner P, Olanow CW. Altered proteasomal function in sporadic Parkinson’s disease. Exp Neurol. 2003;179(1):38–46.

    Article  CAS  PubMed  Google Scholar 

  223. McNaught KS, Jenner P. Proteasomal function is impaired in substantia Nigra in Parkinson’s disease. Neurosci Lett. 2001;297(3):191–4.

    Article  CAS  PubMed  Google Scholar 

  224. Tanaka K. The proteasome: overview of structure and functions. Proc Jpn Acad Ser B Phys Biol Sci. 2009;85(1):12–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Maharjan S, Oku M, Tsuda M, Hoseki J, Sakai Y. Mitochondrial impairment triggers cytosolic oxidative stress and cell death following proteasome Inhibition. Sci Rep. 2014;4:5896.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. McKinnon C, De Snoo ML, Gondard E, et al. Early-onset impairment of the ubiquitin-proteasome system in dopaminergic neurons caused by α-synuclein. Acta Neuropathol Commun. 2020;8(1):17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Wintermeyer P, Krüger R, Kuhn W, et al. Mutation analysis and association studies of the UCHL1 gene in German Parkinson’s disease patients. NeuroReport. 2000;11(10):2079–82.

    Article  CAS  PubMed  Google Scholar 

  228. Chan NC, Salazar AM, Pham AH, et al. Broad activation of the ubiquitin-proteasome system by parkin is critical for mitophagy. Hum Mol Genet. 2011;20(9):1726–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Bukhatwa S, Zeng BY, Rose S, Jenner P. A comparison of changes in proteasomal subunit expression in the substantia Nigra in Parkinson’s disease, multiple system atrophy and progressive supranuclear palsy. Brain Res. 2010;1326:174–83.

    Article  CAS  PubMed  Google Scholar 

  230. McNaught KS, Belizaire R, Jenner P, Olanow CW, Isacson O. Selective loss of 20S proteasome alpha-subunits in the substantia Nigra Pars compacta in Parkinson’s disease. Neurosci Lett. 2002;326(3):155–8.

    Article  CAS  PubMed  Google Scholar 

  231. Chang KH, Lee-Chen GJ, Wu YR, et al. Impairment of proteasome and anti-oxidative pathways in the induced pluripotent stem cell model for sporadic Parkinson’s disease. Parkinsonism Relat Disord. 2016;24:81–8.

    Article  PubMed  Google Scholar 

  232. Betarbet R, Canet-Aviles RM, Sherer TB, et al. Intersecting pathways to neurodegeneration in Parkinson’s disease: effects of the pesticide rotenone on DJ-1, alpha-synuclein, and the ubiquitin-proteasome system. Neurobiol Dis. 2006;22(2):404–20.

    Article  CAS  PubMed  Google Scholar 

  233. Zeng BY, Iravani MM, Lin ST, et al. MPTP treatment of common marmosets impairs proteasomal enzyme activity and decreases expression of structural and regulatory elements of the 26S proteasome. Eur J Neurosci. 2006;23(7):1766–74.

    Article  PubMed  Google Scholar 

  234. Cockram PE, Kist M, Prakash S, Chen SH, Wertz IE, Vucic D. Ubiquitination in the regulation of inflammatory cell death and cancer. Cell Death Differ. 2021;28(2):591–605.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Jansen AH, Reits EA, Hol EM. The ubiquitin proteasome system in glia and its role in neurodegenerative diseases. Front Mol Neurosci. 2014;7:73.

    Article  PubMed  PubMed Central  Google Scholar 

  236. Davidson K, Pickering AM. The proteasome: A key modulator of nervous system function, brain aging, and neurodegenerative disease. Front Cell Dev Biol. 2023;11:1124907.

    Article  PubMed  PubMed Central  Google Scholar 

  237. Zang CX, Wang L, Yang HY, et al. HACE1 negatively regulates neuroinflammation through ubiquitylating and degrading Rac1 in Parkinson’s disease models. Acta Pharmacol Sin. 2022;43(2):285–94.

    Article  CAS  PubMed  Google Scholar 

  238. Pienaar IS, Harrison IF, Elson JL, et al. An animal model mimicking pedunculopontine nucleus cholinergic degeneration in Parkinson’s disease. Brain Struct Funct. 2015;220(1):479–500.

    Article  CAS  PubMed  Google Scholar 

  239. Elson JL, Yates A, Pienaar IS. Pedunculopontine cell loss and protein aggregation direct microglia activation in parkinsonian rats. Brain Struct Funct. 2016;221(4):2319–41.

    Article  CAS  PubMed  Google Scholar 

  240. McNaught KS, Perl DP, Brownell AL, Olanow CW. Systemic exposure to proteasome inhibitors causes a progressive model of Parkinson’s disease. Ann Neurol. 2004;56(1):149–62.

    Article  CAS  PubMed  Google Scholar 

  241. Blandini F, Sinforiani E, Pacchetti C, et al. Peripheral proteasome and caspase activity in Parkinson disease and alzheimer disease. Neurology. 2006;66(4):529–34.

    Article  CAS  PubMed  Google Scholar 

  242. Lin X, Cook TJ, Zabetian CP, et al. DJ-1 isoforms in whole blood as potential biomarkers of Parkinson disease. Sci Rep. 2012;2:954.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all the members of the Tansey lab for insightful discussions during the creation of this manuscript, in particular Dr. Rebecca L. Wallings for her constructive feedback and mentorship. Figures were created with BioRender.com.

Funding

This work was partially supported by funds from the Michael J. Fox Foundation (MJFF-019562 and MJFF-023925), the National Institutes of Health (NIH RF1NS128800), and the Parkinson’s Foundation Research Center of Excellence Award (PF-RCE-1945).

Author information

Authors and Affiliations

Authors

Contributions

J.R.M. and M.G.T. contributed to the writing and generating figures. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Malú Gámez Tansey.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing Interests

The authors have no competing interests to disclose related to the content of this paper.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mark, J.R., Tansey, M.G. Immune cell metabolic dysfunction in Parkinson’s disease. Mol Neurodegeneration 20, 36 (2025). https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13024-025-00827-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doiorg.publicaciones.saludcastillayleon.es/10.1186/s13024-025-00827-y