Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Jul 27.
Published in final edited form as: Neurobiol Dis. 2022 Dec 5;176:105940. doi: 10.1016/j.nbd.2022.105940

The complex role of inflammation and gliotransmitters in Parkinson’s disease

Adithya Gopinath 1,*, Phillip M Mackie 1, Leah T Phan 1, Malú Gámez Tansey 1,*, Habibeh Khoshbouei 1,*
PMCID: PMC10372760  NIHMSID: NIHMS1907538  PMID: 36470499

Abstract

Our understanding of the role of innate and adaptive immune cell function in brain health and how it goes awry during aging and neurodegenerative diseases is still in its infancy. Inflammation and immunological dysfunction are common components of Parkinson’s disease (PD), both in terms of motor and non-motor components of PD. In recent decades, the antiquated notion that the central nervous system (CNS) in disease states is an immune-privileged organ, has been debunked. The immune landscape in the CNS influences peripheral systems, and peripheral immunological changes can alter the CNS in health and disease. Identifying immune and inflammatory pathways that compromise neuronal health and survival is critical in designing innovative and effective strategies to limit their untoward effects on neuronal health.

Keywords: Parkinson’s disease, immunity, cytokines, microglia, peripheral immunity, neuroimmunology, dopamine, dopamine transporter

Introduction

Parkinson’s disease (PD) is the second most common neurodegenerative disease worldwide and is associated with extensive motor impairment and loss of quality of life, but the underlying etiology of PD remains enigmatic(Poewe et al., 2017). In some instances, genetic variants conferring increased risk of PD have been identified, but familial PD represents a small subset (>10%) of PD cases, with most cases still considered of unknown origin or idiopathic (Bandres-Ciga et al., 2020; Lesage and Brice, 2009). Regardless of origin, inflammation and immunological dysfunction are common components of PD, both in terms of motor and non-motor components of PD(Barnum and Tansey, 2012; Kim et al., 2022; Lindqvist et al., 2012). In recent decades, the antiquated notion that the central nervous system (CNS) in disease states is an immune-privileged organ, has been debunked(Brioschi and Colonna, 2019; Harris et al., 2014; Louveau et al., 2015). At large, the field agrees that the immune landscape in the CNS influences peripheral systems, and that peripheral immunological changes can alter the CNS in health and disease and for PD pathogenesis, the contribution of peripheral organs has gained much attention in recent years.

The routes by which neuroimmune communication occurs continues to be debated, with many groups having made headway in this area(Abbott et al., 2018; Goldeck et al., 2016; Iliff et al., 2012; Kortekaas et al., 2005; Mogensen et al., 2021; Ransohoff and Brown, 2012; Ransohoff et al., 2003). But the consensus remains that neuroinflammation and peripheral inflammation occurs in PD. CNS glial cells, namely astrocytes and microglia, are frontline responders to brain changes and have been extensively studied in PD. Similarly, peripheral organs and in specific the role of peripheral blood innate and adaptive immune cells have more recently received extensive coverage in the context of neuroinflammation and peripheral inflammation in PD. In this review, we summarize key findings in CNS inflammation, effects on various neuronal populations, as well as accompanying peripheral immune changes which are implicated in the view of PD as a disease of chronic inflammation.

Microglia and astrocytes in neuroimmune communication and neurodegenerative disease

Neuronal functions in healthy and disease states are modulated by both neuronal factors and via cell-to-cell communication with CNS-resident microglia and astrocytes. Emerging clinical and experimental evidence suggests that normal glial function is critical to normal brain function and neuronal homeostasis, and that dysfunction of these cells plays key roles in the development and/or progression of CNS diseases(Frost and Schafer, 2016; Schwartz et al., 2013; Sofroniew and Vinters, 2010; Wyss-Coray and Mucke, 2002). The ability of glia to recognize and respond to disruptions in proteostasis, such as alpha synuclein aggregates, and glial or immune expression of mutant genes implicated in PD(Booth et al., 2017; Jeon et al., 2020; Zhang et al., 2016), have led to extensive investigation of glia in PD. Indeed, immuno-modulation by glia in preclinical PD models continues to be investigated. Microglia, CNS myeloid cells which are the self-renewing tissue resident macrophages, are heterogenous and survey the CNS parenchyma at steady state via a network of tiny processes. At rest, microglia are termed ramified, and phenotypically characterized by a small cell body, with fine processes extending outwards(Hristovska and Pascual, 2015; Lier et al., 2021; Morioka et al., 1993; Streit, 1993; Streit, 2002; Streit and Graeber, 1993). During insult or injury, microglia become active phagocytes and exhibit an activated, macrophage-like ameboid phenotype(Gopinath et al., 2020b; Shaerzadeh et al., 2020; Streit, 2002). Basics of microglial morphology, activation and subsequent functional changes are investigated extensively; for this information we refer readers to Streit et al 2002, Kettenman et al 2011 among many others(Badanjak et al., 2021; Beynon and Walker, 2012; Donat et al., 2017; Kettenmann et al., 2011; Norden et al., 2016; Pike et al., 2022b; Sarkar et al., 2020; Sarlus and Heneka, 2017; Streit, 2002; Streit et al., 2004).

Microglial dysfunction and senescence occur with age, with some groups indicating microglial function or morphology may vary in key regions affected by PD(De Biase et al., 2017; Shaerzadeh et al., 2020; Streit, 2002; Streit et al., 2004). With age as a primary risk factor for PD, it is possible that the aging process confers increased risk of dopamine neuron degeneration that could be linked to senescence of supporting microglia(Shaerzadeh et al., 2020; Streit, 2002; Streit et al., 2004). The concept of microglial activation and senescence is complex. Early studies conducted by McGeer et al. found that microglia showed increased HLA-DR expression (the human equivalent of major histocompatibility complex 2) nearby to dopamine neurons in the midbrains of PD patients postmortem(McGeer et al., 1988). Subsequent studies showed increased MHC-II expression in experimental models not only in the midbrain but in other brain regions as well, suggesting that microglial antigen presentation may play a critical role in PD progression(Imamura et al., 2003; Kurkowska-Jastrzębska et al., 1999; McGeer et al., 1988). These findings were supported by GWAS studies of PD patients relative to healthy controls, in addition to studies showing that peripheral innate immune cells in individuals with PD and a high-risk genotype in a common genetic variant in HLA-DRA implicated in late-onset PD expressed markedly increased MHCII in response to IFNγ stimulation (Hamza et al., 2010; Hill-Burns et al., 2011; Kannarkat et al., 2015; Nalls et al., 2011). In addition, microglia have also been implicated in cell-to-cell seeding of alpha-synuclein aggregates(George et al., 2019) and evidence from postmortem studies and experimental models of PD-like degeneration have shown that microglial activation is also associated with onset and progression of PD-like pathology, with activation in close association with neuroinflammation(Béraud et al., 2013; Block and Hong, 2007; Harms et al., 2018; Kim and Joh, 2006; Lavisse et al., 2021; Lull and Block, 2010; Muzio et al., 2021; Nagatsu and Sawada, 2005; Ouchi et al., 2009; Ouchi et al., 2005; Smajić et al., 2022). In some preclinical models of PD, microglia-mediated inflammation exacerbates dopamine neuron degeneration via inflammasome activation and subsequent cyclical inflammation(Lee et al., 2019; Pike et al., 2022a; Pike et al., 2022c). Nevertheless, the alternative hypothesis that microglia activation is a response to neuronal injury or cell death that initiates a runaway cycle of neuroinflammation, and a feed-forward cascade of neuronal demise should be considered. It is possible that early immune activation in PD may play protective roles, whereas chronic cycles of inflammation resulting from chronically activated microglia could be maladaptive and lead to the demise of selectively vulnerable neuronal populations. Taken together, these findings suggest microglial activation, antigen presentation via MHC-II, abnormal protein accumulation and subsequent inflammatory responses may play a role in the etiology and pathophysiology of PD.

Studies of microglial activation in PD were quickly followed by studies in which microglial activation was inhibited to attenuate the degeneration of dopamine neurons. In vitro studies showed that antagonists of TLR4 and TLR2, both involved in innate immune defenses against pathogens and the associated inflammatory response, promote dopamine neuron survival, but have yet to be studied in vivo in preclinical PD models(Gorecki et al., 2021; Hughes et al., 2019). Examination of microglia as therapeutic targets for PD are ongoing(Liu et al., 2019). As a note of caution, in other diseases processes such as amyotrophic lateral sclerosis (ALS), early preclinical evidence suggested microglial inhibition might prove an important therapeutic target. However, these results have failed to translate into human clinical trials, indicating that preclinical data from disease models should be interpreted with caution when extrapolating to human subjects(Dupuis et al., 2012; Gordon et al., 2007; Wobst et al., 2020). In the investigation of inflammation in PD, relationships between chronic anti-inflammatory drug use (aspirin, ibuprofen, and other non-steroidal anti-inflammatory drugs; NSAIDs) and the risk of developing PD has been investigated, with multiple analyses suggesting an association between inflammation, NSAID use and PD risk. However, the results were mixed with some studies suggesting that some NSAIDs, but not aspirin, confer decreased risk of developing PD, while other studies show NSAIDs, on the whole, do not alter the risk for PD if taken later in life(Gagne and Power, 2010; Poly et al., 2019; Rees et al., 2011; Samii et al., 2009). Results by Hernan et al. in 2006 suggested possible sex-specific effects of NSAIDs on PD risk, possibly explaining the contrasting results from other studies on the topic and warranting further examination(Hernán et al., 2006) during ongoing investigations of the links between inflammation and PD.

Astrocytes regulate synaptic transmission in addition to innate and adaptive immune responses, depending on timing and context of stimuli present during an inflammatory response (Coulter and Eid, 2012; Kirischuk et al., 2016; Newman, 2003). Astrocyte activation can be beneficial or detrimental for tissue repair, depending on the nature of the inflamed milieu. Astrocytes are in close contact with other CNS-resident cells such as microglia, oligodendrocytes, neurons, and blood vessels that maintain the blood brain barrier (BBB) integrity and its permeability. Astrocytes’ proximity to blood vessels in the CNS enables them to regulate immune cell trafficking via secretion of cytokines and chemokines that can activate adaptive or innate immune responses(Phatnani and Maniatis, 2015; Sofroniew, 2015a; Sofroniew, 2015b; Sofroniew and Vinters, 2010). Growth factors such as glial derived neurotrophic factor (GDNF), brain-derived neurotrophic factor (BDNF) and others are generated and released by astrocytes and are critical to the health and maintenance of dopamine neurons(Lein et al., 2007; Lin et al., 1993; Pöyhönen et al., 2019). In the presence of neuroinflammation, astrocytes become ‘reactive’, secreting a variety of bioactive molecules, including chemokines, growth factors and neurotrophins and altering their production of neuron-supporting growth factors(Linnerbauer et al., 2020; Sofroniew, 2015a; Sofroniew, 2015b). In their role as key regulators of the BBB, astrocyte end-feet surround and enclose CNS vasculature and are well positioned to interact with and influence any peripheral immune cells patrolling the CNS environment(Abbott et al., 2018; Abbott et al., 2006; Linnerbauer et al., 2020; Sofroniew, 2015a; Sofroniew, 2015b). During neuroinflammation, astrocytes are simultaneously exposed to an array of stimuli originated from periphery and CNS that can activate different intracellular signaling pathways and responses. Thus, astrocyte responses are key determinants of the net result of CNS and peripheral immune perturbation. Indeed, astrocytes under inflammatory conditions, are capable of upregulating MHC-II, and possibly presenting CNS antigens to patrolling peripheral immune cells, underscoring the importance of CNS-to-peripheral immune system connections in PD and contributions of peripheral immune cells in the context of PD-associated neuroinflammation(Abbott et al., 2006; Sofroniew, 2015a; Sofroniew, 2015b). The identification of key pathways in neuroinflammation together with the timely reconstruction of the inflamed milieu deserves further investigation. Understanding the nature and duration of immune insult(s) that stimulate or inhibit the production of pathogenic or protective astrocytes are critical for the development of new therapeutic approaches in neuroinflammatory and neurodegenerative disorders.

Immunological mediators in neurotransmission

Immunological mediators, originating from microglia, astrocytes, or peripheral immune cells, can act on non-immune cells in the CNS including neurons. A wide range of studies have found neurons express receptors for multiple types of cytokines, including TNF, IL1b, IL4, and IL10, among others(Friedman, 2001; Herz et al., 2021; Huang et al., 2011; Probert, 2015; Zhou et al., 2009). Although cytokines are most studied for their neuroprotective or pathological effects in disease states, cytokines can modulate the fundamental property of neurons—their electric signaling have apoptotic effects on vulnerable populations. Cytokines can modulate the conductance of several channels across a wide variety of neuronal subtypes in both health and disease. This section reviews some examples of cytokine modulation of neural activity and how this may relate to multiple aspects of the pathophysiology in PD.

While the exact effect a given cytokine has on neuronal activity varies by neuronal subtype, the conserved phenomenon of cytokine modulation of neuronal conductance represents a prime example of the integration between immune and neural signaling. Cytokine regulation of neuronal activity in neurodegeneration, especially PD, represents an open field and compelling area of future study.

Perhaps one of the most well-studied areas for cytokine modulation of neural activity is in the context of pain and sensation. The immune system plays a critical role in the development of chronic pain, and recent studies have attempted to highlight the immune system as a potential therapeutic target(Kavelaars and Heijnen, 2021). Intuitively speaking, crosstalk between the immune system and pain-sensing nociceptive neurons seems like a critical aspect of sensing external threats and physical harm. The immune system detects tissue damage or microbial invasion and communicates to nociceptors which, in turn, relay this information to the CNS to register as pain. Inflammatory cytokines such as TNF, IL1, CCL2, IL33, and IL17, increase the activity of sensory neurons. For example, TNF can increase the amplitude of TRPV1 currents in dorsal root ganglia (DRG) neurons, making them more sensitive to cognate stimuli(Nicol et al., 1997; Sommer et al., 1998). IL17 has a similar effect, but this was only observed in females(Luo et al., 2021). Other pro-inflammatory cytokines increase neuronal excitability through their action on voltage-dependent conductance. Both acute and long-term IL1b can increase voltage-gated sodium (NaV) currents and can also increase hyperpolarized cyclic nucleotide (HCN) and voltage-gated calcium (CaV) currents while decreasing voltage-gated potassium (Kv) currents such as K(Ca)1.1 collectively increasing DRG excitability(Noh et al., 2019; Stemkowski et al., 2021; Stemkowski et al., 2015). Yet another mechanism that has been reported for increased DRG activity in inflammatory settings is through synaptic events. Specifically, CCL2 increases excitatory synaptic event frequency and amplitude(Gao et al., 2009). Thus, cytokines directly modulate pain circuitry through diverse mechanisms involving intrinsic excitability, synaptic connectivity, or both.

Cytokine-mediated neuromodulation has been studied in the context of learning and memory given that cytokines can act as gliotransmitters. Multiple groups have consistently shown that immune-depletion paradigms, either general or targeted, impair performance on the Morris Water Maze, Y-Maze, or Contextual Fear Conditioning tasks (reviewed in(Salvador et al., 2021)). Traditionally, glia has been hypothesized as necessary mediators for the immune system’s effect on cognitive behaviors, but immune signals can directly act on central neurons to influence their excitability and/or connectivity. TNF was reported play a role in synaptic scaling, a form of learning and memory and to increase excitatory transmission in hippocampal neurons by increasing AMPAR membrane levels and decreasing GABA-AR surface levels, but IL-1β, another pro-inflammatory cytokine increased GABAergic post-synaptic potentials and decreased mEPSC frequency, suggesting there is no convergent effect of inflammation on hippocampal excitability(Beattie et al., 2002; Hellstrom et al., 2005; Stellwagen et al., 2005; Stellwagen and Malenka, 2006; Tomasoni et al., 2017). Furthermore, IL-17−/− mice were reported to have increased synaptic transmission in the dentate gyrus in slice preparation and increased intrinsic excitability in vitro(Ribeiro et al., 2019). Contrary to these findings, a separate report found that IL-17−/− mice had decreased synaptic transmission in the hippocampal CA1 region after a short-term memory task(Liu et al., 2014). Additionally, blocking IL-17α in an experimental encephalitis model partially rescues LTP deficits(Di Filippo et al., 2021). Similarly, anti-inflammatory cytokines, such as IL-4 and IL-10 have divergent effects on intrinsic hippocampal excitability by either decreasing or increasing K(Ca) currents, respectively(Chen et al., 2020; Levin et al., 2016; Nenov et al., 2019). Together these reports highlight the regional and context specificity that cytokine signaling can have in the CNS at steady and inflammatory states. Thus, in both branches of the nervous system, a motif appears: cytokines act as critical integrators of environmental conditions to neuronal computation. Across neuron subpopulations, cytokine modulation is a conserved phenomenon at the steady state, albeit with divergent effects. This knowledge has already been applied to pain research and shows promising signs of potential clinical translation, but our understanding of this phenomenon in neurodegeneration is limited. There are a few reports of the effect of cytokines in neuroinflammatory conditions but in the context of Alzheimer’s disease (AD) and PD, this is an area that deserves more effort(Di Filippo et al., 2021). Recognizing that cytokine signaling can affect memory formation at baseline, they may play a role in the cognitive decline of AD. Surprisingly, there is a dearth of knowledge on cytokine modulation of midbrain dopamine neurons, representing a critical area for future studies.

A surge of studies that cytokines directly modulate neuronal activity in health and disease have emerged, but there are still outstanding questions on cell-type specific mechanisms and relevance to behavior. Some groups have reported the intracellular signaling cascades responsible for increased or decreased electrical properties of the neurons following stimulation by cytokines, however this information is likely specific to certain neuronal types (Belkouch et al., 2011; Obreja et al., 2002; Schäfers et al., 2003). Given the growing appreciation for neuronal heterogeneity even within a single brain region, it is unknown how some cytokines may have divergent effects on membrane conductance and neuronal excitability(Blum et al., 2021; Kamath et al., 2022; O’Leary et al., 2020; Poulin et al., 2016). Whether cytokines affect expression, trafficking, or gating, these mechanisms must be elucidated to inform clinical implementation of cytokine-targeting therapies. Indeed, IL-4R expression on GABAergic neurons has been shown to decrease learning in a fear conditioning paradigm by altering dentate gyrus synaptic transmission, and, in a similar contextual learning paradigm, CCR5 was shown to modulate ensemble formation in dCA1 thereby affecting learning(Herz et al., 2021; Shen et al., 2022). Collectively these and subsequent studies will generate a more comprehensive atlas of cytokine-mediated modulation of behavior in the steady and inflammatory states, informing the field regarding cytokines as a therapeutic target in PD.

Midbrain dopamine neurons exhibit a canonical pacemaking property, tonically firing at 1–8Hz depending on the system(Grace and Bunney, 1984; Lin et al., 2021; Otomo et al., 2020). Properly regulated firing and neurotransmitter release by Substantia Nigra pars compacta dopamine neurons is critical in initiating goal-directed movement and is lost in PD. Dopamine neurons, particularly in the Substantia Nigra pars compacta, rely on the coordinated activity of a number of voltage and ligand-gated ion channels, including voltage-gated sodium (NaV), voltage-gated potassium (Kv), voltage-gated calcium (Cav), and hyperpolarized cyclic nucleotide-gated (HCN) channels to maintain their pacemaking. Additionally, ionotropic ligand-gated receptors, such as n-methyl-D-aspartate (NMDA) receptors are expressed on dopamine neurons and in the striatum where they regulate bursting activity and play crucial roles in synaptic plasticity. For a more comprehensive review on these channels in dopamine neurons see work by Gantz and colleagues (Gantz et al., 2018). Together, these conductances culminate in the tonic and phasic dopamine activity responsible for movement.

Alterations in channel expression and localization have been noted in PD animal models and post-mortem tissue. CaV channels, specifically CaV1.3, have long been thought to be critical in PD progression, as their inhibition rescued some animal models(Chan et al., 2007; Guzman et al., 2018; Kang et al., 2012; Lee et al., 2014; Liss and Striessnig, 2019; Marras et al., 2012; Schapira et al., 2014; Surmeier et al., 2012; Surmeier et al., 2017). However, it should be noted, that a clinical trial of CaV1.3 inhibitor failed, and a recent animal model suggests that this channel may be downregulated early in disease(2020; González-Rodríguez et al., 2021). Nevertheless, increased calcium influx is a major suspect of bioenergetic imbalances thought to belie dopaminergic vulnerability(González-Rodríguez et al., 2021; Guzman et al., 2010; Lieberman et al., 2017; Sanchez-Padilla et al., 2014). Other pacemaking related channels, particularly potassium channels, have also been implicated in PD(Chen et al., 2018). Specifically K(Ca) channel SK and A-type potassium channels have shown altered expression or potential therapeutic use in a variety of animal models including the A53T mouse model, 6OHDA rat models, and MPTP models in rat and non-human primate, and even in post-mortem tissue(Aidi-Knani et al., 2015; Alvarez-Fischer et al., 2013; Chen et al., 2014; Doo et al., 2010; Haghdoost-Yazdi et al., 2011; Kim et al., 2011; Lu et al., 2019; Mourre et al., 2017; Subramaniam et al., 2014). Again, disrupted pacemaking activity is thought to exacerbate bioenergetic stresses in dopamine neurons potentially leading to their demise. Furthermore, circuit dysfunction is a major cause of motor symptoms as well as L-DOPA-related complications (reviewed in (McGregor and Nelson, 2019)), and changes in NMDA receptor levels in the striatum are thought to contribute to these symptoms(Calabresi et al., 2000a; Calabresi et al., 2000b; Hallett and Standaert). Whether by altering basal ganglia circuitry or by potentially contributing to dopamine neuron vulnerability, dysregulation of membrane conductance represents an important aspect of PD pathophysiology.

While there are several immune-based, progressive, animal models of PD, none have been studied for how immune signaling modulates neural activity in model progression. All the channels above have shown some capacity to be modulated by cytokines. For example, A-type currents in sensory neurons are suppressed by IL-33(Wang et al., 2022b). CaVs are regulated by a host of cytokines—TNF increases L-type currents in hippocampal neurons, IL-1β increases CaV currents in sensory neurons, and IL-17α decreases N-type currents in sympathetic neurons(Chisholm et al., 2012; Furukawa and Mattson, 1998; Noh et al., 2019; Stemkowski et al., 2015). Similarly, ionotropic glutamate receptors, including NMDA receptors, exhibit a diverse sensitivity to various cytokines. Low dose IL-1β increases NMDA-mediated Ca-influx in hippocampal neurons(Viviani et al., 2003), but decreases mEPSCs at a higher dose(Tomasoni et al., 2017). Thus, channel modulation by cytokines exhibits significant cell-type specificity. Currently, there is a lack of information on cytokine modulation of mid-brain dopaminergic neurons. Studies focusing on these neurons specifically will be required before examining immune-mediated changes to conductances in PD models. Nevertheless, such investigations may reveal additional indications for cytokine-targeted therapies in PD.

Peripheral myeloid cells in Parkinson’s disease

While the involvement of peripheral immunity in neuroinflammation has garnered a lot of attention, investigating whether changes in the peripheral immune system are cause or consequence of degeneration of dopamine neurons or other selectively vulnerable populations that degenerate in PD, recent data support the interpretation that loss of catecholaminergic neurons in PD alters immunophenotype of peripheral myeloid and lymphoid cells (Chen et al., 2021; De Virgilio et al., 2016; Gardai et al., 2013; Gopinath et al., 2021; Gopinath et al., 2020b; Gopinath et al., 2020c; Gopinath, 2022; Grozdanov et al., 2019; Kustrimovic et al., 2019; Kuter et al., 2020; Öberg et al., 2021; Sonninen et al., 2020; Sulzer et al., 2017; Tan et al., 2020; Troncoso-Escudero et al., 2018; Vila et al., 2001). These data strongly support the existence of a compensatory immune response in the peripheral immune cells, in PD. Within the myeloid compartment, PD patients exhibit pronounced increase in inflammatory monocytes that is associated with a decrease in resolution-phase immune cells(Gopinath, 2022; Grozdanov et al., 2014). The inflammatory characteristics of these cells underpins an altered dynamic between inflammatory monocytes (that are normally upregulated by inflammatory mediators such as TNF and IL-1β) and resolution-phase monocytes that display a tissue-repairing phenotype.

Macrophages and monocytes are of particular interest among myeloid cells, as these cells are dominant first responders to inflammation(Kapellos et al., 2019; Merah-Mourah et al., 2020; Wijeyekoon et al., 2018; Wong et al., 2012). Blood derived monocytes and macrophages are known to enter the CNS under acute inflammatory conditions in response to CNS lesions and secrete inflammatory factors which can contribute to neuroinflammation and degeneration(Cheng et al., 1998; Harms et al., 2018; Varvel et al., 2016). The recent discovery that parenchymal border macrophages (PBMs), a specialized kind of perivascular and leptomeningeal macrophages, regulate CSF flow and clear toxic aggregates suggests a complex interplay between multiple myeloid cell types in maintaining CNS homeostasis(Drieu et al., 2022; Mazzitelli et al., 2022). Unlike chronic conditions such as PD, acute CNS inflammation in multiple sclerosis (MS) is associated with extensive peripheral immune infiltration into the CNS and subsequent inflammatory damage(Forbes and Miron, 2022; Mikita et al., 2011). Indeed, peripheral myeloid cells acquire and present CNS antigens to peripheral adaptive immune cells, directly contributing to demyelination in MS(Riedhammer and Weissert, 2015; van Langelaar et al., 2020). Whether or not peripheral myeloid cells participate directly in neurodegenerative processes in PD remains to be determined. Recent data from Sulzer and colleagues support the idea that alpha-synuclein-specific T-cells develop in PD patients and may correlate with markers of clinical progression as described in more detail below(Dhanwani et al., 2022; Garretti et al., 2019; Sulzer et al., 2017). Finally, in addition to peripheral myeloid cells, recent reports of reservoirs of myeloid cells in the meninges derived from the skull and vertebral bone raise the possibility that these could play important role in maintaining brain health, and may become dysfunctional with age and contribute to PD (Cugurra et al., 2021).

Dendritic cells (DCs), which arise from myeloid lineage cells in the bone marrow, are uniquely positioned among myeloid cells as professional antigen presentation cells (APCs)(Hilligan and Ronchese, 2020). Foremost among DC roles in immunity is to present peptides from inflammatory or pathogenic states to other immune cells, including B and T cells (adaptive immune system), to activate a systemic immune response. In the context of CNS inflammation, DCs can migrate from the meninges and glymphatic vessels after exposure to CNS antigens, trafficking to nearby lymph nodes (primarily cervical lymph nodes) and present CNS antigens to peripheral lymphoid cells(De Laere et al., 2018; Laman and Weller, 2013; Mohammad et al., 2014; Sagar et al., 2012). Subsequently, peripheral adaptive immune cells can home into the inflamed CNS compartment, potentially contributing to inflammation and neurodegeneration.

Lymphoid cells in Parkinson’s disease

Peripheral lymphoid cells enter the CNS compartment both under steady state and in disease state(Balasa et al., 2021; Cabarrocas et al., 2003; Gross et al., 2020; Kleine and Benes, 2006; Strazielle et al., 2016). Most notably, in PD, T cells are detected near dying dopamine neurons, suggesting T cells may play a role in early degeneration or PD progression(Ambrosi et al., 2017; Brochard et al., 2009; Galiano-Landeira et al., 2020; Garretti et al., 2019; Sulzer et al., 2017). However, whether T cells are responsible for directly damaging dopamine neurons remain less understood. In neurodegeneration and neuroinflammation, T cells which recognize specific antigens (antigen-specific T cells) are of particular interest. CD8 positive T cells, known as cytotoxic T cells, and CD4 positive T cells, known as helper T cells, both can recognize antigens. The investigation of antigen specificity in T cells of PD patients have shown variable results, reflecting the heterogeneity of human PD and the limited interpretation of T cell involvement in the neurodegenerative process(Garretti et al., 2019; Singhania et al., 2021; Sulzer et al., 2017). For example, Sulzer and others have shown that antigen specific T cells in PD patients can recognize alpha synuclein, implicated in familial PD(Bandres-Ciga et al., 2020; Oliveira et al., 2015; Singleton et al., 2003; Sulzer et al., 2017). However, the epitopes recognized by T cells derived from PD patients did not reveal any single antigen dominating the peripheral T cell landscape suggesting while T cells may be involved in PD, a single clonally expanded population of antigen-specific T cells is unlikely to directly underlies degeneration of dopamine neurons(Singhania et al., 2021). Of interest, is that T cells are primary sources of inflammatory cytokine interferon gamma (IFNy) in the periphery, which has been repeatedly shown to play a role in PD and PD-like neurodegeneration(Barcia et al., 2011; Ferrari et al., 2021; Kambayashi et al., 2003; Murray et al., 2002; Panagiotakopoulou et al., 2020; Perussia, 1991).

B cells are also key players in the adaptive immune system but have not been directly detected in the brains of PD patients, PD mice nor in proximity to degenerating neurons(Brochard et al., 2009; McGeer et al., 1988; Wang et al., 2022a). However, given that PD patients experience persistent peripheral inflammation, the link between altered B cells and PD is theorized(Campos-Acuña et al., 2019; Ferrari and Tarelli, 2011; Gopinath et al., 2021; Gopinath et al., 2022; Kustrimovic et al., 2019; Qin et al., 2016; Süβ et al., 2021; Williams et al., 2022). Consistent with this hypothesis, decreased naïve B cells and an increase in memory B cell subsets are reported in PD patients, suggesting an immunological activation state in response to inflammation(Wang et al., 2022a). B cells undergo a process called somatic hypermutation, during which highly specific antigen-specific antibodies are produced(Bräuninger et al., 2003; Küppers, 2003; Wang et al., 2022a). A requirement for somatic hypermutation is an immunological trigger, such as inflammation(Beltrán et al., 2014; Brink and Phan, 2018; Pierre et al., 1997). Since both CNS and peripheral inflammation is thought to occur in PD, both peripheral and CNS locations for B cell affinity maturation have been investigated(Jain and Yong, 2021; Lu and Robinson, 2014). While still uncorroborated, meninges could function as a lymphoid interface in which antigen specific T cells meet and induce B cell somatic hypermutation(Hartlehnert et al., 2021; Serafini et al., 2004). Consistent with this idea, a subset of meningeal B cells were recently shown to derive from a local bone marrow source (the calvaria) and become antigen-experienced in the CNS(Brioschi et al., 2021). Potential roles of B cells in PD warrants further exploration, as antibodies specific to components of dopamine neurons have been detected in the blood of PD patients at higher levels than in control subjects(Besong-Agbo et al., 2013; Carvey et al., 1991; Han et al., 2012; Kannarkat et al., 2013).

Inflammatory mediators in Parkinson’s disease

Beyond antigen presentation and canonical immune functions, both CNS and peripheral immune cells are involved in the inflammatory response, and therefore are thought to be central to neuroinflammation in PD. Mounting evidence suggests that PD involves a chronic inflammatory state, supported by the fact that inflammatory cytokines such as TNF, IL-6, IL-1β and other cytokines are increased in both brain and peripheral immune system(Barnum and Tansey, 2012; Imamura et al., 2003; Kim et al., 2022; Lindqvist et al., 2012; Magnusen et al., 2021; Mogi et al., 1994; Nagatsu et al., 2000; Nagatsu and Sawada, 2005; Sawada et al., 2006). Whether cytokines from the CNS can reach peripheral circulation and vice versa in PD is unclear. Microglia and astrocytes are potent sources of CNS cytokines including TNF and IL6, with activated microglia secreting inflammatory cytokines and upregulating expression of MHC-II(Ezcurra et al., 2010; Hickey and Kimura, 1988; Imamura et al., 2003; Imamura et al., 1990; Perry et al., 1985; Perry et al., 2010). Microglia are particularly dense in the midbrain region(Kim et al., 2000), with different percentages of microglia relative to dopamine neurons in the dopaminergic midbrain(Shaerzadeh et al., 2020). Whether or not microglia are the main drivers of neurodegeneration, dying dopamine neurons release damage-signal molecules such as neuromelanin, ATP, alpha-synuclein, driving microglial activation and subsequent inflammatory cytokine secretion(Davalos et al., 2005; Gao and Hong, 2008; Perry et al., 2010; Wilms et al., 2003; Zhang et al., 2005). In fact, microglial release of inflammatory molecules may influence not only dopamine neuron death but also induce further cyclic microglial activation, potentially leading to a runaway cycle of neuroinflammation(Block and Hong, 2007). While still debated and lacking clinical evidence, the literature supports a role for cytokine signaling in PD.

In the brains of PD patients, expression of IL-1, IL-2, IL-1β, TNF, IL6, TGFβ, and IFNγ have all been associated with gliosis, dopamine neuron loss or both (Collins et al., 2012; Gopinath et al., 2021; Kang et al., 2021; Karpenko et al., 2018; Marogianni et al., 2020). Both CSF and blood of PD patients show increased levels of TNF, IL-1β and IL-6, with increased TGFβ largely restricted to CSF(Imamura et al., 2003; Marogianni et al., 2020). Most compelling are the data indicating that inhibition of cytokine signaling, particularly TNF, in PD models resulted in partial rescue of dopamine neurons, suggesting cytokine signaling originating in the CNS may play a role in PD pathophysiology and etiology(Elyaman and Khoury, 2017; McCoy et al., 2006; McCoy and Tansey, 2008; Mirza et al., 2000). Epidemiological data also support the role of TNF in PD pathogenesis; specifically, individuals with inflammatory bowel disease (IBD) have higher risk for late-onset PD unless treated with anti-TNF biologics (Peter et al., 2018).

Strong evidence in the literature, both experimental and clinical studies, support the notion that TNF is increased in both CNS and periphery in PD(McCoy et al., 2006; McCoy and Tansey, 2008; Mogi et al., 1994; Nagatsu et al., 2000; Qin et al., 2016). While many cells express TNF-receptors, immune cells in both CNS and periphery are the primary source for TNF(Aggarwal, 2003; Bazzoni and Beutler, 1996; Kalliolias and Ivashkiv, 2016; Kraft et al., 2009; MacEwan, 2002; Veroni et al., 2010). Myeloid derived cells such as peripheral monocytes and CNS-resident microglia are particularly sensitive to inflammatory stimuli and secrete TNF upon immune stimulation(Eissner et al., 2000; Gregersen et al., 2000; Hanisch, 2002; Kraft et al., 2009; Segueni et al., 2016). In the CNS, during inflammation, microglia are the main source for TNF(Gregersen et al., 2000; Welser-Alves and Milner, 2013). Whereas in the periphery, monocytes, and other myeloid cells as well as endothelial tissue produce TNF under inflammatory conditions(Gane et al., 2016; Imaizumi et al., 2000; Krishnaswamy et al., 1999; Mandal et al., 2020; Schröder et al., 2018). In addition, recent studies have shown that PD patients’ peripheral immune cells are marked by altered expression of tyrosine hydroxylase, the rate limiting enzyme in dopamine synthesis typically studied in dopamine neurons, with a direct link to increased TNF, suggesting that CNS-originating cytokines such as TNF may influence peripheral immunophenotype and modulate the inflammatory immune landscape in PD(Gopinath et al., 2021).

PD postmortem studies show increased expression of reactive oxygen species (ROS) from mitochondrial stress and inducible nitric oxide synthase, which contribute to increased ROS and can damage dopamine neurons(Anzalone et al.). The significance of oxidative stress is supported by many preclinical and clinical studies revealing the impact of mitochondrial ROS on dopamine neuron degeneration(Dias et al., 2013; Ekstrand and Galter, 2009; Good et al., 2011; Hao et al., 2010; Meredith and Rademacher, 2011; Morais et al., 2009; Simola et al., 2007; Trist et al., 2019; Weng et al., 2018). ROS is considered fundamental to the degenerative process, where the substantia nigra (SN) of PD patients exhibit mitochondrial dysfunction, disturbed calcium homeostasis, and reduced antioxidant molecules(James et al., 2015; Reeve et al., 2014; Trist et al., 2019; Venkateshappa et al., 2012). PD is associated with a significant increase in iron in the degenerating SN dopamine neurons that is measurable in living PD patients and in post-mortem PD brain. Dopamine neurons experience high baseline levels of oxidative stress due to their canonical pace-making activity; dopamine release underlie the higher energy demands of dopamine neurons than other neurons(Dagra et al., 2021; Dias et al., 2013; Lin et al., 2021; Pissadaki and Bolam, 2013). Therefore, dopamine neurons are uniquely vulnerable to ROS-mediated damage(Surmeier et al., 2012). Mitochondria are a potent source of ROS at steady state and during degeneration of dopamine neurons(Brand et al., 2004). During PD progression, extracellular ROS produced by glia may be an additional source of ROS. In healthy states, mitochondrial enzymes counterbalance excess ROS production, while mutations in ROS-scavenging mitochondrial enzymes confer a PD phenotype in PD animal models and in PD patients. Induced radical species are expressed on demand in myeloid cells in all tissue compartments in both CNS and periphery(Anzalone et al.; Brigelius-Flohé and Maiorino, 2013; Karnati et al., 2013; Kawamata and Manfredi, 2010; Ruszkiewicz and Albrecht, 2015). During systemic inflammation, both microglia and peripheral myeloid cells are abundant sources of oxygen radicals(Kraft and Harry, 2011; Kumar et al., 2014; Taylor et al., 2003). In animal models of PD, including 6-hydroxydopamine (6-OHDA) and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), microglial oxygen radicals accelerate the toxin-mediated dopamine neuron lesion(Gao et al., 2003a; Gao et al., 2003b; Zhang et al., 2005). Though microglial-originating radical species are unlikely to reach peripheral circulation in sufficient concentrations to influence peripheral immune cells, cytokines produced by activated microglia may enter peripheral circulation via glymphatic drainage and/or subarachnoid granulations and alter the inflammatory state of peripheral myeloid cells(Norden and Godbout, 2013). Since cytokines are BBB permeable, peripheral cytokines can influence the CNS neuroinflammatory landscape in PD(Norden and Godbout, 2013). This bidirectional mechanism may be central to a feed-forward CNS-to-periphery loop of runaway systemic inflammation and by extension, neurodegeneration(Gopinath, 2022). Microglial and peripheral immune responses to inflammatory challenges are also exaggerated in aged subjects compared to healthy adult subjects, aligning with the overarching hypothesis in the PD field that the disease results, at least in part, from the intersection of age and inflammation in brain and periphery (Buchanan et al., 2008; Rosczyk et al., 2008; Sierra et al., 2007). We therefore speculate that a combination of oxidative stress, lysosomal and/or mitochondrial dysfunction in combination with an inflammatory environment related to aging, lifestyle choices, etc., create the “perfect storm” for PD pathogenesis and progression (Tansey et al., 2022).

Dopamine transporter as a nexus for CNS-periphery inflammation in PD

Recent studies show dopamine, and the associated synthesis, storage and clearance mechanisms are dysregulated in PD(Gopinath et al., 2021; Gopinath, 2022; Mackie et al., 2022). Over the last three decades, PD diagnosis has relied on the loss of dopamine transporter (DAT) in the brains of PD patients(Brooks, 2010; Haavik and Toska, 1998; Murtomäki et al., 2022; Nagatsu et al., 2019; Nutt et al., 2004; Tabrez et al., 2012). Beyond serving as a PD marker, in general, DAT and vesicular monoamine transporter 2 (VMAT2) in the CNS collectively regulate dopamine homeostasis. Free extracellular dopamine quickly oxidizes and forms dopamine radicals such as quinolones, damaging dopamine neurons(Anderson et al., 2011; Miller et al., 1996; Terland et al., 1997). DAT rapidly clears extracellular dopamine and transports it into the intracellular space for repackaging into synaptic vesicles via VMAT2(Cheng et al., 2015; Pramod et al., 2013; Vaughan and Foster, 2013). Thus, the clearance and storage mechanisms protect the neurons from the oxidative stress. Peripheral immune cells such as monocytes and macrophages express functional dopaminergic proteins, indicating that in peripheral immune cells dopaminergic proteins modulate the activity of immune cells(Ambrosi et al., 2017; Cosentino et al., 2002; Gaskill et al., 2012; Gopinath et al., 2021; Gopinath, 2022; Kustrimovic et al., 2019; Mackie et al., 2018; Mackie et al., 2022; Matt and Gaskill, 2020; Qiu et al., 2004; Reguzzoni et al., 2002). For example, following TNF-induced inflammation an immune phenotype shift occurs where the number of Tyrosine hydroxylase (TH) expressing monocytes are increased. This is a reversible immune phenotype. Selective soluble TNF blockade, via XPro-1595, reverses the immuno-phenotype to control levels(Barnum et al., 2014; Gopinath et al., 2021; Karamita et al., 2017; MacPherson et al., 2017). Since TNF is increased in the CNS, CSF, and blood of PD patients, these data have significant clinical implication(Barnum et al., 2014; Gopinath et al., 2021; Karamita et al., 2017; MacPherson et al., 2017). Importantly, DAT in human myeloid cells also exhibits an immunosuppressive role(Mackie et al., 2022). In the presence of the inflammogen LPS, DAT works in reverse transport mode, transporting dopamine out of monocytes leading to attenuation of LPS-induced responses(Goodwin et al., 2009; Khoshbouei et al., 2003; Mackie et al., 2018; Mackie et al., 2022; Saha et al., 2014; Sambo et al., 2017). Subsequent studies provided a paradigm shift on the concept of CNS-to-periphery neuroimmune communication in PD. These studies revealed that in PD patients, the number of peripheral immune cells expressing DAT is increased. We posit that this phenotype change is triggered by the loss of CNS dopamine neurons, representing a compensatory mechanism to attenuate peripheral inflammation(Gopinath et al., 2020a; Gopinath et al., 2022).

Summary and Conclusions

Our understanding of the role of innate and adaptive immune cell function in brain health and how it goes awry during aging is still in its infancy. However, there is consensus that inflammatory mediators derived from brain-resident sentinels in charge of protecting neurons may be produced early in the disease process to produce neurotrophic factors for vulnerable neuronal populations that are sending signals that they are in trouble and becoming dysfunctional. In this manner, early immune activation in PD may play protective roles, whereas chronic cycles of inflammation resulting from chronically activated microglia and infiltrating immune cells from the periphery could be maladaptive and lead to the demise of selectively vulnerable neuronal populations. The challenge for the field is to a) identify the immune and inflammatory pathways that compromise neuronal health and survival to enable design of innovative and more effective strategies to limit their effects on neuronal health; b) use this knowledge to design biomarker-directed trials using immunomodulatory neuroprotective drugs to monitor target engagement and patient responsiveness to treatment.

Acknowledgements:

This work was funded by T32-NS082128 (to A.G.), National Center for Advancing Translational Sciences of the National Institutes of Health under University of Florida Clinical and Translational Science Awards TL1TR001428 and UL1TR001427 (to A.G. and P.M.), R01NS071122-07A1 (to H.K.), NIDA Grant R01DA026947-10 (to H.K.), NIA Grant RF1AG057247-05 (to M.G.T.), NINDS Grant RF1NS128800-01 (to M.G.T.,), the Parkinson’s Foundation Research Center of Excellence Award PF-RCE-1945 (to M.G.T.), the Weston Family Foundation (to M.G.T.), the Michael J. Fox Foundation for Parkinson’s Research (MJFF-18212, 18891 and 16778) and Aligning Science Across Parkinson’s (ASAP-020621) (to M.G.T.), UF-Fixel Institute Norman and Susan Fixel endowment (to M.G.T.), UF-Fixel Institute Developmental Fund, DA043895 (to H.K.), by the University of Florida McKnight Brain Institute (MBI) (to A.G.), by the Bryan Robinson Foundation (to A.G.) and by The Karen Toffler Charitable Trust (to A.G.).

Footnotes

Competing interests: Malú Gámez Tansey is a co-inventor on the XPro1595 patent and is a consultant to and has stock ownership in INmune Bio, which has licensed XPro1595 for neurological indications. Dr. Tansey is a collaborator, advisor, and/or consultant to the Parkinson’s Foundation, the Weston Family Foundation, the Alzheimer’s Association, the Bright Focus Foundation, Cerebral Therapeutics, Jaya Biosciences, NysnoBio, SciNeuro, Longevity Biotech, Biogen/IONIS, Amylyx, Merck, Genentech, Sanofi, iMetabolic Pharma, Novo Nordisk, and iMMvention. The remaining authors declare no competing interests.

References

  1. 2020. Isradipine Versus Placebo in Early Parkinson Disease: A Randomized Trial. Ann Intern Med. 172, 591–598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Abbott NJ, et al. , 2018. The role of brain barriers in fluid movement in the CNS: is there a ‘glymphatic’ system? Acta Neuropathol. 135, 387–407. [DOI] [PubMed] [Google Scholar]
  3. Abbott NJ, et al. , 2006. Astrocyte–endothelial interactions at the blood–brain barrier. Nature reviews neuroscience. 7, 41–53. [DOI] [PubMed] [Google Scholar]
  4. Aggarwal BB, 2003. Signalling pathways of the TNF superfamily: a double-edged sword. Nat Rev Immunol. 3, 745–56. [DOI] [PubMed] [Google Scholar]
  5. Aidi-Knani S, et al. , 2015. Kv4 channel blockade reduces motor and neuropsychiatric symptoms in rodent models of Parkinson’s disease. Behav Pharmacol. 26, 91–100. [DOI] [PubMed] [Google Scholar]
  6. Alvarez-Fischer D, et al. , 2013. Bee venom and its component apamin as neuroprotective agents in a Parkinson disease mouse model. PLoS One. 8, e61700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Ambrosi G, et al. , 2017. Complex Changes in the Innate and Adaptive Immunity Accompany Progressive Degeneration of the Nigrostriatal Pathway Induced by Intrastriatal Injection of 6-Hydroxydopamine in the Rat. Neurotox Res. 32, 71–81. [DOI] [PubMed] [Google Scholar]
  8. Anderson DG, et al. , 2011. Oxidation of 3,4-dihydroxyphenylacetaldehyde, a toxic dopaminergic metabolite, to a semiquinone radical and an ortho-quinone. J Biol Chem. 286, 26978–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Anzalone A, et al. , 2012. Dual control of dopamine synthesis and release by presynaptic and postsynaptic dopamine D2 receptors. J Neurosci. 32, 9023–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Badanjak K, et al. , 2021. The Contribution of Microglia to Neuroinflammation in Parkinson’s Disease. Int J Mol Sci. 22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Balasa R, et al. , 2021. Reviewing the Significance of Blood-Brain Barrier Disruption in Multiple Sclerosis Pathology and Treatment. Int J Mol Sci. 22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bandres-Ciga S, et al. , 2020. Genetics of Parkinson’s disease: An introspection of its journey towards precision medicine. Neurobiology of Disease. 137, 104782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Barcia C, et al. , 2011. IFN-γ signaling, with the synergistic contribution of TNF-α, mediates cell specific microglial and astroglial activation in experimental models of Parkinson’s disease. Cell Death Dis. 2, e142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Barnum CJ, et al. , 2014. Peripheral administration of the selective inhibitor of soluble tumor necrosis factor (TNF) XPro®1595 attenuates nigral cell loss and glial activation in 6-OHDA hemiparkinsonian rats. J Parkinsons Dis. 4, 349–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Barnum CJ, Tansey MG, 2012. Neuroinflammation and non-motor symptoms: the dark passenger of Parkinson’s disease? Curr Neurol Neurosci Rep. 12, 350–8. [DOI] [PubMed] [Google Scholar]
  16. Bazzoni F, Beutler B, 1996. The tumor necrosis factor ligand and receptor families. N Engl J Med. 334, 1717–25. [DOI] [PubMed] [Google Scholar]
  17. Beattie EC, et al. , 2002. Control of synaptic strength by glial TNFalpha. Science. 295, 2282–5. [DOI] [PubMed] [Google Scholar]
  18. Belkouch M, et al. , 2011. The Chemokine CCL2 Increases Nav1.8 Sodium Channel Activity in Primary Sensory Neurons through a Gβγ-Dependent Mechanism. The Journal of Neuroscience. 31, 18381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Beltrán E, et al. , 2014. Intrathecal somatic hypermutation of IgM in multiple sclerosis and neuroinflammation. Brain. 137, 2703–14. [DOI] [PubMed] [Google Scholar]
  20. Béraud D, et al. , 2013. Microglial activation and antioxidant responses induced by the Parkinson’s disease protein α-synuclein. J Neuroimmune Pharmacol. 8, 94–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Besong-Agbo D, et al. , 2013. Naturally occurring α-synuclein autoantibody levels are lower in patients with Parkinson disease. Neurology. 80, 169–175. [DOI] [PubMed] [Google Scholar]
  22. Beynon SB, Walker FR, 2012. Microglial activation in the injured and healthy brain: what are we really talking about? Practical and theoretical issues associated with the measurement of changes in microglial morphology. Neuroscience. 225, 162–71. [DOI] [PubMed] [Google Scholar]
  23. Block ML, Hong JS, 2007. Chronic microglial activation and progressive dopaminergic neurotoxicity. Biochem Soc Trans. 35, 1127–32. [DOI] [PubMed] [Google Scholar]
  24. Blum JA, et al. , 2021. Single-cell transcriptomic analysis of the adult mouse spinal cord reveals molecular diversity of autonomic and skeletal motor neurons. 24, 572–583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Booth HDE, et al. , 2017. The Role of Astrocyte Dysfunction in Parkinson’s Disease Pathogenesis. Trends Neurosci. 40, 358–370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Brand MD, et al. , 2004. Mitochondrial superoxide: production, biological effects, and activation of uncoupling proteins. Free Radical Biology and Medicine. 37, 755–767. [DOI] [PubMed] [Google Scholar]
  27. Bräuninger A, et al. , 2003. Epstein-Barr virus (EBV)-positive lymphoproliferations in post-transplant patients show immunoglobulin V gene mutation patterns suggesting interference of EBV with normal B cell differentiation processes. Eur J Immunol. 33, 1593–602. [DOI] [PubMed] [Google Scholar]
  28. Brigelius-Flohé R, Maiorino M, 2013. Glutathione peroxidases. Biochimica et Biophysica Acta (BBA)-General Subjects. 1830, 3289–3303. [DOI] [PubMed] [Google Scholar]
  29. Brink R, Phan TG, 2018. Self-Reactive B Cells in the Germinal Center Reaction. Annu Rev Immunol. 36, 339–357. [DOI] [PubMed] [Google Scholar]
  30. Brioschi S, Colonna M, 2019. The CNS Immune-Privilege Goes Down the Drain(age). Trends Pharmacol Sci. 40, 1–3. [DOI] [PubMed] [Google Scholar]
  31. Brioschi S, et al. , 2021. Heterogeneity of meningeal B cells reveals a lymphopoietic niche at the CNS borders. Science. 373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Brochard V, et al. , 2009. Infiltration of CD4+ lymphocytes into the brain contributes to neurodegeneration in a mouse model of Parkinson disease. J Clin Invest. 119, 182–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Brooks DJ, 2010. Imaging dopamine transporters in Parkinson’s disease. Biomark Med. 4, 651–60. [DOI] [PubMed] [Google Scholar]
  34. Buchanan J, et al. , 2008. Cognitive and neuroinflammatory consequences of mild repeated stress are exacerbated in aged mice. Psychoneuroendocrinology. 33, 755–765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Cabarrocas J, et al. , 2003. Effective and selective immune surveillance of the brain by MHC class I-restricted cytotoxic T lymphocytes. Eur J Immunol. 33, 1174–82. [DOI] [PubMed] [Google Scholar]
  36. Calabresi P, et al. , 2000a. Electrophysiology of dopamine in normal and denervated striatal neurons. 23, S57–S63. [DOI] [PubMed] [Google Scholar]
  37. Calabresi P, et al. , 2000b. Levodopa-induced dyskinesia: a pathological form of striatal synaptic plasticity? Annals of neurology. 47, S60–8; discussion S68. [PubMed] [Google Scholar]
  38. Campos-Acuña J, et al. , 2019. T-Cell-Driven Inflammation as a Mediator of the Gut-Brain Axis Involved in Parkinson’s Disease. Front Immunol. 10, 239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Carvey P, et al. , 1991. The potential use of a dopamine neuron antibody and a striatal‐derived neurotrophic factor as diagnostic markers in Parkinson’s disease. Neurology. 41, 53–58. [DOI] [PubMed] [Google Scholar]
  40. Chan CS, et al. , 2007. ‘Rejuvenation’ protects neurons in mouse models of Parkinson’s disease. Nature. 447, 1081–6. [DOI] [PubMed] [Google Scholar]
  41. Chen L, et al. , 2014. SK channel blockade reverses cognitive and motor deficits induced by nigrostriatal dopamine lesions in rats. Int J Neuropsychopharmacol. 17, 1295–306. [DOI] [PubMed] [Google Scholar]
  42. Chen X, et al. , 2021. Evidence for Peripheral Immune Activation in Parkinson’s Disease. Frontiers in Aging Neuroscience. 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Chen X, et al. , 2018. Potassium Channels: A Potential Therapeutic Target for Parkinson’s Disease. Neurosci Bull. 34, 341–348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Chen X, et al. , 2020. Deficiency of anti-inflammatory cytokine IL-4 leads to neural hyperexcitability and aggravates cerebral ischemia–reperfusion injury. 10, 1634–1645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Cheng L, et al. , 1998. Entry of monocytes into the brain after injection of Corynebacterium parvum. Exp Neurol. 149, 322–8. [DOI] [PubMed] [Google Scholar]
  46. Cheng MH, et al. , 2015. Insights into the Modulation of Dopamine Transporter Function by Amphetamine, Orphenadrine, and Cocaine Binding. Frontiers in Neurology. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Chisholm SP, et al. , 2012. Interleukin-17A increases neurite outgrowth from adult postganglionic sympathetic neurons. J Neurosci. 32, 1146–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Collins LM, et al. , 2012. Contributions of central and systemic inflammation to the pathophysiology of Parkinson’s disease. Neuropharmacology. 62, 2154–2168. [DOI] [PubMed] [Google Scholar]
  49. Cosentino M, et al. , 2002. Catecholamine production and tyrosine hydroxylase expression in peripheral blood mononuclear cells from multiple sclerosis patients: effect of cell stimulation and possible relevance for activation-induced apoptosis. J Neuroimmunol. 133, 233–40. [DOI] [PubMed] [Google Scholar]
  50. Coulter DA, Eid T, 2012. Astrocytic regulation of glutamate homeostasis in epilepsy. Glia. 60, 1215–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Cugurra A, et al. , 2021. Skull and vertebral bone marrow are myeloid cell reservoirs for the meninges and CNS parenchyma. Science. 373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Dagra A, et al. , 2021. α-Synuclein-induced dysregulation of neuronal activity contributes to murine dopamine neuron vulnerability. npj Parkinson’s Disease. 7, 76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Davalos D, et al. , 2005. ATP mediates rapid microglial response to local brain injury in vivo. Nature neuroscience. 8, 752–758. [DOI] [PubMed] [Google Scholar]
  54. De Biase LM, et al. , 2017. Local Cues Establish and Maintain Region-Specific Phenotypes of Basal Ganglia Microglia. Neuron. 95, 341–356.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. De Laere M, et al. , 2018. To the Brain and Back: Migratory Paths of Dendritic Cells in Multiple Sclerosis. J Neuropathol Exp Neurol. 77, 178–192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. De Virgilio A, et al. , 2016. Parkinson’s disease: Autoimmunity and neuroinflammation. Autoimmun Rev. 15, 1005–11. [DOI] [PubMed] [Google Scholar]
  57. Dhanwani R, et al. , 2022. Transcriptional analysis of peripheral memory T cells reveals Parkinson’s disease-specific gene signatures. NPJ Parkinsons Dis. 8, 30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Di Filippo M, et al. , 2021. Interleukin-17 affects synaptic plasticity and cognition in an experimental model of multiple sclerosis. Cell Rep. 37, 110094. [DOI] [PubMed] [Google Scholar]
  59. Dias V, et al. , 2013. The role of oxidative stress in Parkinson’s disease. J Parkinsons Dis. 3, 461–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Donat CK, et al. , 2017. Microglial Activation in Traumatic Brain Injury. Front Aging Neurosci. 9, 208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Doo AR, et al. , 2010. Neuroprotective effects of bee venom pharmaceutical acupuncture in acute 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced mouse model of Parkinson’s disease. Neurol Res. 32 Suppl 1, 88–91. [DOI] [PubMed] [Google Scholar]
  62. Drieu A, et al. , 2022. Parenchymal border macrophages regulate the flow dynamics of the cerebrospinal fluid. Nature. 611, 585–593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Dupuis L, et al. , 2012. A randomized, double blind, placebo-controlled trial of pioglitazone in combination with riluzole in amyotrophic lateral sclerosis. PloS one. 7, e37885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Eissner G, et al. , 2000. Reverse signaling through transmembrane TNF confers resistance to lipopolysaccharide in human monocytes and macrophages. J Immunol. 164, 6193–8. [DOI] [PubMed] [Google Scholar]
  65. Ekstrand MI, Galter D, 2009. The MitoPark Mouse - an animal model of Parkinson’s disease with impaired respiratory chain function in dopamine neurons. Parkinsonism Relat Disord. 15 Suppl 3, S185–8. [DOI] [PubMed] [Google Scholar]
  66. Elyaman W, Khoury SJ, Th9 cells in the pathogenesis of EAE and multiple sclerosis. Seminars in immunopathology, Vol. 39. Springer, 2017, pp. 79–87. [DOI] [PubMed] [Google Scholar]
  67. Ezcurra ALDL, et al. , 2010. Chronic expression of low levels of tumor necrosis factor-α in the substantia nigra elicits progressive neurodegeneration, delayed motor symptoms and microglia/macrophage activation. Neurobiology of disease. 37, 630–640. [DOI] [PubMed] [Google Scholar]
  68. Ferrari CC, Tarelli R, 2011. Parkinson’s disease and systemic inflammation. Parkinsons Dis. 2011, 436813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Ferrari DP, et al. , 2021. Interferon-γ Involvement in the Neuroinflammation Associated with Parkinson’s Disease and L-DOPA-Induced Dyskinesia. Neurotox Res. 39, 705–719. [DOI] [PubMed] [Google Scholar]
  70. Forbes LH, Miron VE, 2022. Monocytes in central nervous system remyelination. Glia. 70, 797–807. [DOI] [PubMed] [Google Scholar]
  71. Friedman WJ, 2001. Cytokines regulate expression of the type 1 interleukin-1 receptor in rat hippocampal neurons and glia. Exp Neurol. 168, 23–31. [DOI] [PubMed] [Google Scholar]
  72. Frost JL, Schafer DP, 2016. Microglia: Architects of the Developing Nervous System. Trends in Cell Biology. 26, 587–597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Furukawa K, Mattson MP, 1998. The transcription factor NF-kappaB mediates increases in calcium currents and decreases in NMDA- and AMPA/kainate-induced currents induced by tumor necrosis factor-alpha in hippocampal neurons. J Neurochem. 70, 1876–86. [DOI] [PubMed] [Google Scholar]
  74. Gagne JJ, Power MC, 2010. Anti-inflammatory drugs and risk of Parkinson disease: a meta-analysis. Neurology. 74, 995–1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Galiano-Landeira J, et al. , 2020. CD8 T cell nigral infiltration precedes synucleinopathy in early stages of Parkinson’s disease. Brain. 143, 3717–3733. [DOI] [PubMed] [Google Scholar]
  76. Gane JM, et al. , 2016. TNF-α Autocrine Feedback Loops in Human Monocytes: The Pro- and Anti-Inflammatory Roles of the TNF-α Receptors Support the Concept of Selective TNFR1 Blockade In Vivo. J Immunol Res. 2016, 1079851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Gantz SC, et al. , 2018. The Evolving Understanding of Dopamine Neurons in the Substantia Nigra and Ventral Tegmental Area. Annual Review of Physiology. 80, 219–241. [DOI] [PubMed] [Google Scholar]
  78. Gao H-M, Hong J-S, 2008. Why neurodegenerative diseases are progressive: uncontrolled inflammation drives disease progression. Trends in immunology. 29, 357–365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Gao HM, et al. , 2003a. Critical role of microglial NADPH oxidase-derived free radicals in the in vitro MPTP model of Parkinson’s disease. Faseb j. 17, 1954–6. [DOI] [PubMed] [Google Scholar]
  80. Gao HM, et al. , 2003b. Synergistic dopaminergic neurotoxicity of MPTP and inflammogen lipopolysaccharide: relevance to the etiology of Parkinson’s disease. Faseb j. 17, 1957–9. [DOI] [PubMed] [Google Scholar]
  81. Gao YJ, et al. , 2009. JNK-induced MCP-1 production in spinal cord astrocytes contributes to central sensitization and neuropathic pain. J Neurosci. 29, 4096–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Gardai SJ, et al. , 2013. Elevated alpha-synuclein impairs innate immune cell function and provides a potential peripheral biomarker for Parkinson’s disease. PLoS One. 8, e71634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Garretti F, et al. , 2019. Autoimmunity in Parkinson’s Disease: The Role of α-Synuclein-Specific T Cells. Front Immunol. 10, 303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Gaskill PJ, et al. , 2012. Characterization and function of the human macrophage dopaminergic system: implications for CNS disease and drug abuse. J Neuroinflammation. 9, 203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. George S, et al. , 2019. Microglia affect α-synuclein cell-to-cell transfer in a mouse model of Parkinson’s disease. Molecular Neurodegeneration. 14, 34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Goldeck D, et al. , 2016. Peripheral Immune Signatures in Alzheimer Disease. Curr Alzheimer Res. 13, 739–49. [DOI] [PubMed] [Google Scholar]
  87. González-Rodríguez P, et al. , 2021. Disruption of mitochondrial complex I induces progressive parkinsonism. Nature. 599, 650–656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Good CH, et al. , 2011. Impaired nigrostriatal function precedes behavioral deficits in a genetic mitochondrial model of Parkinson’s disease. Faseb j. 25, 1333–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Goodwin JS, et al. , 2009. Amphetamine and Methamphetamine Differentially Affect Dopamine Transporters in Vitro and in Vivo*. Journal of Biological Chemistry. 284, 2978–2989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Gopinath A, et al. , 2021. TNFα increases tyrosine hydroxylase expression in human monocytes. NPJ Parkinsons Dis. 7, 62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Gopinath A, et al. , 2020a. Microglia and other myeloid cells in CNS health and disease. Journal of Pharmacology and Experimental Therapeutics. jpet.120.265058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Gopinath A, et al. , 2020b. Microglia and Other Myeloid Cells in Central Nervous System Health and Disease. J Pharmacol Exp Ther. 375, 154–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Gopinath A, et al. , 2020c. A novel approach to study markers of dopamine signaling in peripheral immune cells. J Immunol Methods. 476, 112686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Gopinath A, et al. , 2022. DAT and TH expression marks human Parkinson’s disease in peripheral immune cells. npj Parkinson’s Disease. 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Gordon PH, et al. , 2007. Efficacy of minocycline in patients with amyotrophic lateral sclerosis: a phase III randomised trial. The Lancet Neurology. 6, 1045–1053. [DOI] [PubMed] [Google Scholar]
  96. Gorecki AM, et al. , 2021. TLR2 and TLR4 in Parkinson’s disease pathogenesis: the environment takes a toll on the gut. Translational Neurodegeneration. 10, 47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Grace AA, Bunney BS, 1984. The control of firing pattern in nigral dopamine neurons: single spike firing. J Neurosci. 4, 2866–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Gregersen R, et al. , 2000. Microglia and macrophages are the major source of tumor necrosis factor in permanent middle cerebral artery occlusion in mice. J Cereb Blood Flow Metab. 20, 53–65. [DOI] [PubMed] [Google Scholar]
  99. Gross CC, et al. , 2020. Generation of a Model to Predict Differentiation and Migration of Lymphocyte Subsets under Homeostatic and CNS Autoinflammatory Conditions. Int J Mol Sci. 21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Grozdanov V, et al. , 2014. Inflammatory dysregulation of blood monocytes in Parkinson’s disease patients. Acta Neuropathol. 128, 651–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Grozdanov V, et al. , 2019. Increased Immune Activation by Pathologic α-Synuclein in Parkinson’s Disease. Ann Neurol. 86, 593–606. [DOI] [PubMed] [Google Scholar]
  102. Guzman JN, et al. , 2018. Systemic isradipine treatment diminishes calcium-dependent mitochondrial oxidant stress. J Clin Invest. 128, 2266–2280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Guzman JN, et al. , 2010. Oxidant stress evoked by pacemaking in dopaminergic neurons is attenuated by DJ-1. Nature. 468, 696–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Haavik J, Toska K, 1998. Tyrosine hydroxylase and Parkinson’s disease. Mol Neurobiol. 16, 285–309. [DOI] [PubMed] [Google Scholar]
  105. Haghdoost-Yazdi H, et al. , 2011. Significant effects of 4-aminopyridine and tetraethylammonium in the treatment of 6-hydroxydopamine-induced Parkinson’s disease. Behav Brain Res. 223, 70–4. [DOI] [PubMed] [Google Scholar]
  106. Hallett PJ, Standaert DG
  107. Hamza TH, et al. , 2010. Common genetic variation in the HLA region is associated with late-onset sporadic Parkinson’s disease. Nature genetics. 42, 781–785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Han M, et al. , 2012. Diagnosis of Parkinson’s disease based on disease-specific autoantibody profiles in human sera. PloS one. 7, e32383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Hanisch UK, 2002. Microglia as a source and target of cytokines. Glia. 40, 140–55. [DOI] [PubMed] [Google Scholar]
  110. Hao L-Y, et al. , 2010. DJ-1 is critical for mitochondrial function and rescues PINK1 loss of function. Proceedings of the National Academy of Sciences. 107, 9747–9752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Harms AS, et al. , 2018. Peripheral monocyte entry is required for alpha-Synuclein induced inflammation and Neurodegeneration in a model of Parkinson disease. Exp Neurol. 300, 179–187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Harris MG, et al. , 2014. Immune privilege of the CNS is not the consequence of limited antigen sampling. Sci Rep. 4, 4422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Hartlehnert M, et al. , 2021. Bcl6 controls meningeal Th17-B cell interaction in murine neuroinflammation. Proc Natl Acad Sci U S A. 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Hellstrom IC, et al. , 2005. Chronic LPS exposure produces changes in intrinsic membrane properties and a sustained IL-β-dependent increase in GABAergic inhibition in hippocampal CA1 pyramidal neurons. Hippocampus. 15, 656–664. [DOI] [PubMed] [Google Scholar]
  115. Hernán MA, et al. , 2006. Nonsteroidal anti-inflammatory drugs and the incidence of Parkinson disease. Neurology. 66, 1097–1099. [DOI] [PubMed] [Google Scholar]
  116. Herz J, et al. , 2021. GABAergic neuronal IL-4R mediates T cell effect on memory. Neuron. 109, 3609–3618.e9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Hickey WF, Kimura H, 1988. Perivascular microglial cells of the CNS are bone marrow-derived and present antigen in vivo. Science. 239, 290–2. [DOI] [PubMed] [Google Scholar]
  118. Hill-Burns EM, et al. , 2011. Evidence for more than one Parkinson’s disease-associated variant within the HLA region. PLoS One. 6, e27109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Hilligan KL, Ronchese F, 2020. Antigen presentation by dendritic cells and their instruction of CD4+ T helper cell responses. Cellular & Molecular Immunology. 17, 587–599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Hristovska I, Pascual O, 2015. Deciphering Resting Microglial Morphology and Process Motility from a Synaptic Prospect. Front Integr Neurosci. 9, 73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Huang Y, et al. , 2011. Neuron-specific effects of interleukin-1β are mediated by a novel isoform of the IL-1 receptor accessory protein. J Neurosci. 31, 18048–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Hughes CD, et al. , 2019. Picomolar concentrations of oligomeric alpha-synuclein sensitizes TLR4 to play an initiating role in Parkinson’s disease pathogenesis. Acta neuropathologica. 137, 103–120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Iliff JJ, et al. , 2012. A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid β. Sci Transl Med. 4, 147ra111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Imaizumi T, et al. , 2000. Expression of tumor necrosis factor-alpha in cultured human endothelial cells stimulated with lipopolysaccharide or interleukin-1alpha. Arterioscler Thromb Vasc Biol. 20, 410–5. [DOI] [PubMed] [Google Scholar]
  125. Imamura K, et al. , 2003. Distribution of major histocompatibility complex class II-positive microglia and cytokine profile of Parkinson’s disease brains. Acta neuropathologica. 106, 518–526. [DOI] [PubMed] [Google Scholar]
  126. Imamura K, et al. , 1990. Generation and characterization of monoclonal antibodies against rat microglia and ontogenic distribution of positive cells. Lab Invest. 63, 853–61. [PubMed] [Google Scholar]
  127. Jain RW, Yong VW, 2021. B cells in central nervous system disease: diversity, locations and pathophysiology. Nature Reviews Immunology. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. James SA, et al. , 2015. Direct in vivo imaging of ferrous iron dyshomeostasis in ageing Caenorhabditis elegans. Chemical science. 6, 2952–2962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Jeon Y-M, et al. , 2020. The Role of Glial Mitochondria in α-Synuclein Toxicity. Frontiers in Cell and Developmental Biology. 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Kalliolias GD, Ivashkiv LB, 2016. TNF biology, pathogenic mechanisms and emerging therapeutic strategies. Nat Rev Rheumatol. 12, 49–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Kamath T, et al. , 2022. Single-cell genomic profiling of human dopamine neurons identifies a population that selectively degenerates in Parkinson’s disease. 25, 588–595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Kambayashi T, et al. , 2003. Memory CD8+ T Cells Provide an Early Source of IFN-γ. The Journal of Immunology. 170, 2399–2408. [DOI] [PubMed] [Google Scholar]
  133. Kang S, et al. , 2012. CaV1.3-selective L-type calcium channel antagonists as potential new therapeutics for Parkinson’s disease. Nat Commun. 3, 1146. [DOI] [PubMed] [Google Scholar]
  134. Kang X, et al. , 2021. Tumor Necrosis Factor Inhibition and Parkinson Disease. A Mendelian Randomization Study. 96, e1672–e1679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Kannarkat GT, et al. , 2013. The role of innate and adaptive immunity in Parkinson’s disease. J Parkinsons Dis. 3, 493–514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Kannarkat GT, et al. , 2015. Common Genetic Variant Association with Altered HLA Expression, Synergy with Pyrethroid Exposure, and Risk for Parkinson’s Disease: An Observational and Case-Control Study. NPJ Parkinsons Dis 1, 15002–. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Kapellos TS, et al. , 2019. Human Monocyte Subsets and Phenotypes in Major Chronic Inflammatory Diseases. Front Immunol. 10, 2035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Karamita M, et al. , 2017. Therapeutic inhibition of soluble brain TNF promotes remyelination by increasing myelin phagocytosis by microglia. JCI Insight. 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Karnati S, et al. , 2013. Mammalian SOD2 is exclusively located in mitochondria and not present in peroxisomes. Histochemistry and cell biology. 140, 105–117. [DOI] [PubMed] [Google Scholar]
  140. Karpenko M, et al. , 2018. Interleukin-1β, interleukin-1 receptor antagonist, interleukin-6, interleukin-10, and tumor necrosis factor-α levels in CSF and serum in relation to the clinical diversity of Parkinson’s disease. Cellular immunology. 327, 77–82. [DOI] [PubMed] [Google Scholar]
  141. Kavelaars A, Heijnen CJ, 2021. Immune regulation of pain: Friend and foe. Sci Transl Med. 13, eabj7152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Kawamata H, Manfredi G, 2010. Import, maturation, and function of SOD1 and its copper chaperone CCS in the mitochondrial intermembrane space. Antioxidants & redox signaling. 13, 1375–1384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Kettenmann H, et al. , 2011. Physiology of microglia. Physiological reviews. 91, 461–553. [DOI] [PubMed] [Google Scholar]
  144. Khoshbouei H, et al. , 2003. Amphetamine-induced dopamine efflux. A voltage-sensitive and intracellular Na+-dependent mechanism. J Biol Chem. 278, 12070–7. [DOI] [PubMed] [Google Scholar]
  145. Kim JI, et al. , 2011. Bee venom reduces neuroinflammation in the MPTP-induced model of Parkinson’s disease. Int J Neurosci. 121, 209–17. [DOI] [PubMed] [Google Scholar]
  146. Kim R, et al. , 2022. Serum Inflammatory Markers and Progression of Nonmotor Symptoms in Early Parkinson’s Disease. Mov Disord. [DOI] [PubMed] [Google Scholar]
  147. Kim W-G, et al. , 2000. Regional difference in susceptibility to lipopolysaccharide-induced neurotoxicity in the rat brain: role of microglia. Journal of Neuroscience. 20, 6309–6316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Kim YS, Joh TH, 2006. Microglia, major player in the brain inflammation: their roles in the pathogenesis of Parkinson’s disease. Exp Mol Med. 38, 333–47. [DOI] [PubMed] [Google Scholar]
  149. Kirischuk S, et al. , 2016. Astrocyte sodium signaling and the regulation of neurotransmission. Glia. 64, 1655–66. [DOI] [PubMed] [Google Scholar]
  150. Kleine TO, Benes L, 2006. Immune surveillance of the human central nervous system (CNS): different migration pathways of immune cells through the blood-brain barrier and blood-cerebrospinal fluid barrier in healthy persons. Cytometry A. 69, 147–51. [DOI] [PubMed] [Google Scholar]
  151. Kortekaas R, et al. , 2005. Blood-brain barrier dysfunction in parkinsonian midbrain in vivo. Ann Neurol. 57, 176–9. [DOI] [PubMed] [Google Scholar]
  152. Kraft AD, Harry GJ, 2011. Features of microglia and neuroinflammation relevant to environmental exposure and neurotoxicity. Int J Environ Res Public Health. 8, 2980–3018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Kraft AD, et al. , 2009. Heterogeneity of microglia and TNF signaling as determinants for neuronal death or survival. Neurotoxicology. 30, 785–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Krishnaswamy G, et al. , 1999. Human endothelium as a source of multifunctional cytokines: molecular regulation and possible role in human disease. J Interferon Cytokine Res. 19, 91–104. [DOI] [PubMed] [Google Scholar]
  155. Kumar A, et al. , 2014. Inducible nitric oxide synthase is key to peroxynitrite-mediated, LPS-induced protein radical formation in murine microglial BV2 cells. Free Radic Biol Med. 73, 51–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Küppers R, 2003. Somatic hypermutation and B cell receptor selection in normal and transformed human B cells. Ann N Y Acad Sci. 987, 173–9. [DOI] [PubMed] [Google Scholar]
  157. Kurkowska-Jastrzębska I, et al. , 1999. The inflammatory reaction following 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine intoxication in mouse. Experimental neurology. 156, 50–61. [DOI] [PubMed] [Google Scholar]
  158. Kustrimovic N, et al. , 2019. Peripheral Immunity, Immunoaging and Neuroinflammation in Parkinson’s Disease. Curr Med Chem. 26, 3719–3753. [DOI] [PubMed] [Google Scholar]
  159. Kuter KZ, et al. , 2020. The role of glia in Parkinson’s disease: Emerging concepts and therapeutic applications. Prog Brain Res. 252, 131–168. [DOI] [PubMed] [Google Scholar]
  160. Laman JD, Weller RO, 2013. Drainage of cells and soluble antigen from the CNS to regional lymph nodes. J Neuroimmune Pharmacol. 8, 840–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Lavisse S, et al. , 2021. Increased microglial activation in patients with Parkinson disease using [(18)F]-DPA714 TSPO PET imaging. Parkinsonism Relat Disord. 82, 29–36. [DOI] [PubMed] [Google Scholar]
  162. Lee E, et al. , 2019. MPTP-driven NLRP3 inflammasome activation in microglia plays a central role in dopaminergic neurodegeneration. Cell Death & Differentiation. 26, 213–228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Lee YC, et al. , 2014. Antihypertensive agents and risk of Parkinson’s disease: a nationwide cohort study. PLoS One. 9, e98961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Lein ES, et al. , 2007. Genome-wide atlas of gene expression in the adult mouse brain. Nature. 445, 168–176. [DOI] [PubMed] [Google Scholar]
  165. Lesage S, Brice A, 2009. Parkinson’s disease: from monogenic forms to genetic susceptibility factors. Human molecular genetics. 18, R48–R59. [DOI] [PubMed] [Google Scholar]
  166. Levin SG, et al. , 2016. Role of BK(Ca) Potassium Channels in the Mechanisms of Modulatory Effects of IL-10 on Hypoxia-Induced Changes in Activity of Hippocampal Neurons. Bull Exp Biol Med. 160, 643–5. [DOI] [PubMed] [Google Scholar]
  167. Lieberman OJ, et al. , 2017. alpha-Synuclein-Dependent Calcium Entry Underlies Differential Sensitivity of Cultured SN and VTA Dopaminergic Neurons to a Parkinsonian Neurotoxin. eNeuro. 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Lier J, et al. , 2021. Beyond Activation: Characterizing Microglial Functional Phenotypes. Cells. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Lin L-FH, et al. , 1993. GDNF: a glial cell line-derived neurotrophic factor for midbrain dopaminergic neurons. Science. 260, 1130–1132. [DOI] [PubMed] [Google Scholar]
  170. Lin M, et al. , 2021. In Parkinson’s patient-derived dopamine neurons, the triplication of α-synuclein locus induces distinctive firing pattern by impeding D2 receptor autoinhibition. Acta Neuropathol Commun. 9, 107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Lindqvist D, et al. , 2012. Non-motor symptoms in patients with Parkinson’s disease - correlations with inflammatory cytokines in serum. PLoS One. 7, e47387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Linnerbauer M, et al. , 2020. Astrocyte Crosstalk in CNS Inflammation. Neuron. 108, 608–622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Liss B, Striessnig J, 2019. The Potential of L-Type Calcium Channels as a Drug Target for Neuroprotective Therapy in Parkinson’s Disease. Annual Review of Pharmacology and Toxicology. 59, 263–289. [DOI] [PubMed] [Google Scholar]
  174. Liu C-Y, et al. , 2019. Pharmacological Targeting of Microglial Activation: New Therapeutic Approach. Frontiers in Cellular Neuroscience. 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Liu Q, et al. , 2014. Interleukin-17 inhibits Adult Hippocampal Neurogenesis. 4, 7554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Louveau A, et al. , 2015. Revisiting the Mechanisms of CNS Immune Privilege. Trends Immunol. 36, 569–577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Lu DR, Robinson WH, 2014. Street-experienced peripheral B cells traffic to the brain. Sci Transl Med. 6, 248fs31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Lu J, et al. , 2019. The potassium channel KCa3.1 represents a valid pharmacological target for microgliosis-induced neuronal impairment in a mouse model of Parkinson’s disease. 16, 273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Lull ME, Block ML, 2010. Microglial activation and chronic neurodegeneration. Neurotherapeutics. 7, 354–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Luo X, et al. , 2021. IL-23/IL-17A/TRPV1 axis produces mechanical pain via macrophage-sensory neuron crosstalk in female mice. 109, 2691–2706.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. MacEwan DJ, 2002. TNF receptor subtype signalling: differences and cellular consequences. Cell Signal. 14, 477–92. [DOI] [PubMed] [Google Scholar]
  182. Mackie P, et al. , 2018. The dopamine transporter: An unrecognized nexus for dysfunctional peripheral immunity and signaling in Parkinson’s Disease. Brain Behav Immun. 70, 21–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Mackie PM, et al. , 2022. Functional characterization of the biogenic amine transporters on human macrophages. JCI Insight. 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. MacPherson KP, et al. , 2017. Peripheral administration of the soluble TNF inhibitor XPro1595 modifies brain immune cell profiles, decreases beta-amyloid plaque load, and rescues impaired long-term potentiation in 5xFAD mice. Neurobiol Dis. 102, 81–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Magnusen AF, et al. , 2021. Genetic Defects and Pro-inflammatory Cytokines in Parkinson’s Disease. Frontiers in Neurology. 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Mandal P, et al. , 2020. TNF Signaling Dictates Myeloid and Non-Myeloid Cell Crosstalk to Execute MCMV-Induced Extrinsic Apoptosis. Viruses. 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Marogianni C, et al. , 2020. Neurodegeneration and Inflammation—An Interesting Interplay in Parkinson’s Disease. International Journal of Molecular Sciences. 21, 8421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Marras C, et al. , 2012. Dihydropyridine calcium channel blockers and the progression of parkinsonism. Ann Neurol. 71, 362–9. [DOI] [PubMed] [Google Scholar]
  189. Matt SM, Gaskill PJ, 2020. Where Is Dopamine and how do Immune Cells See it?: Dopamine-Mediated Immune Cell Function in Health and Disease. J Neuroimmune Pharmacol. 15, 114–164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Mazzitelli JA, et al. , 2022. Cerebrospinal fluid regulates skull bone marrow niches via direct access through dural channels. Nat Neurosci. 25, 555–560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. McCoy MK, et al. , 2006. Blocking soluble tumor necrosis factor signaling with dominant-negative tumor necrosis factor inhibitor attenuates loss of dopaminergic neurons in models of Parkinson’s disease. J Neurosci. 26, 9365–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. McCoy MK, Tansey MG, 2008. TNF signaling inhibition in the CNS: implications for normal brain function and neurodegenerative disease. J Neuroinflammation. 5, 45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. McGeer PL, et al. , 1988. Reactive microglia are positive for HLA‐DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology. 38, 1285–1285. [DOI] [PubMed] [Google Scholar]
  194. McGregor MM, Nelson AB, 2019. Circuit Mechanisms of Parkinson’s Disease. 101, 1042–1056. [DOI] [PubMed] [Google Scholar]
  195. Merah-Mourah F, et al. , 2020. Identification of Novel Human Monocyte Subsets and Evidence for Phenotypic Groups Defined by Interindividual Variations of Expression of Adhesion Molecules. Scientific Reports. 10, 4397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Meredith GE, Rademacher DJ, 2011. MPTP mouse models of Parkinson’s disease: an update. J Parkinsons Dis. 1, 19–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Mikita J, et al. , 2011. Altered M1/M2 activation patterns of monocytes in severe relapsing experimental rat model of multiple sclerosis. Amelioration of clinical status by M2 activated monocyte administration. Mult Scler. 17, 2–15. [DOI] [PubMed] [Google Scholar]
  198. Miller JW, et al. , 1996. Oxidative damage caused by free radicals produced during catecholamine autoxidation: protective effects of O-methylation and melatonin. Free Radic Biol Med. 21, 241–9. [DOI] [PubMed] [Google Scholar]
  199. Mirza B, et al. , 2000. The absence of reactive astrocytosis is indicative of a unique inflammatory process in Parkinson’s disease. Neuroscience. 95, 425–32. [DOI] [PubMed] [Google Scholar]
  200. Mogensen FL, et al. , 2021. The Glymphatic System (En)during Inflammation. Int J Mol Sci. 22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Mogi M, et al. , 1994. Tumor necrosis factor-alpha (TNF-alpha) increases both in the brain and in the cerebrospinal fluid from parkinsonian patients. Neurosci Lett. 165, 208–10. [DOI] [PubMed] [Google Scholar]
  202. Mohammad MG, et al. , 2014. Immune cell trafficking from the brain maintains CNS immune tolerance. J Clin Invest. 124, 1228–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Morais VA, et al. , 2009. Parkinson’s disease mutations in PINK1 result in decreased Complex I activity and deficient synaptic function. EMBO Mol Med. 1, 99–111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Morioka T, et al. , 1993. Characterization of microglial reaction after middle cerebral artery occlusion in rat brain. J Comp Neurol. 327, 123–32. [DOI] [PubMed] [Google Scholar]
  205. Mourre C, et al. , 2017. Changes in SK channel expression in the basal ganglia after partial nigrostriatal dopamine lesions in rats: Functional consequences. Neuropharmacology. 113, 519–532. [DOI] [PubMed] [Google Scholar]
  206. Murray PD, et al. , 2002. Cellular sources and targets of IFN-gamma-mediated protection against viral demyelination and neurological deficits. Eur J Immunol. 32, 606–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Murtomäki K, et al. , 2022. Gastrointestinal Symptoms and Dopamine Transporter Asymmetry in Early Parkinson’s Disease. Mov Disord. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Muzio L, et al. , 2021. Microglia in Neuroinflammation and Neurodegeneration: From Understanding to Therapy. Frontiers in Neuroscience. 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Nagatsu T, et al. , 2000. Changes in cytokines and neurotrophins in Parkinson’s disease. J Neural Transm Suppl. 277–90. [DOI] [PubMed] [Google Scholar]
  210. Nagatsu T, et al. , 2019. Human tyrosine hydroxylase in Parkinson’s disease and in related disorders. J Neural Transm (Vienna). 126, 397–409. [DOI] [PubMed] [Google Scholar]
  211. Nagatsu T, Sawada M, 2005. Inflammatory process in Parkinson’s disease: role for cytokines. Curr Pharm Des. 11, 999–1016. [DOI] [PubMed] [Google Scholar]
  212. Nalls M, et al. , 2011. International Parkinson Disease Genomics Consortium Imputation of sequence variants for identification of genetic risks for Parkinson’s disease: A meta-analysis of genome-wide association studies. Lancet. 377, 641–649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Nenov MN, et al. , 2019. Interleukin-10 Facilitates Glutamatergic Synaptic Transmission and Homeostatic Plasticity in Cultured Hippocampal Neurons. Int J Mol Sci. 20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Newman EA, 2003. New roles for astrocytes: regulation of synaptic transmission. Trends Neurosci. 26, 536–42. [DOI] [PubMed] [Google Scholar]
  215. Nicol GD, et al. , 1997. Tumor necrosis factor enhances the capsaicin sensitivity of rat sensory neurons. J Neurosci. 17, 975–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Noh MC, et al. , 2019. Long-term actions of interleukin-1β on K(+), Na(+) and Ca(2+) channel currents in small, IB(4)-positive dorsal root ganglion neurons; possible relevance to the etiology of neuropathic pain. J Neuroimmunol. 332, 198–211. [DOI] [PubMed] [Google Scholar]
  217. Norden DM, Godbout JP, 2013. Review: microglia of the aged brain: primed to be activated and resistant to regulation. Neuropathol Appl Neurobiol. 39, 19–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Norden DM, et al. , 2016. Sequential activation of microglia and astrocyte cytokine expression precedes increased Iba-1 or GFAP immunoreactivity following systemic immune challenge. Glia. 64, 300–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Nutt JG, et al. , 2004. The dopamine transporter: importance in Parkinson’s disease. Ann Neurol. 55, 766–73. [DOI] [PubMed] [Google Scholar]
  220. O’Leary TP, et al. , 2020. Extensive and spatially variable within-cell-type heterogeneity across the basolateral amygdala. eLife. 9, e59003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. C1 - eLife 2020;9:e59003. [Google Scholar]
  222. Öberg M, et al. , 2021. The role of innate immunity and inflammation in Parkinsońs disease. Scand J Immunol. 93, e13022. [DOI] [PubMed] [Google Scholar]
  223. Obreja O, et al. , 2002. IL-1 beta potentiates heat-activated currents in rat sensory neurons: involvement of IL-1RI, tyrosine kinase, and protein kinase C. Faseb j. 16, 1497–503. [DOI] [PubMed] [Google Scholar]
  224. Oliveira LM, et al. , 2015. Elevated α-synuclein caused by SNCA gene triplication impairs neuronal differentiation and maturation in Parkinson’s patient-derived induced pluripotent stem cells. Cell Death Dis. 6, e1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Otomo K, et al. , 2020. In vivo patch-clamp recordings reveal distinct subthreshold signatures and threshold dynamics of midbrain dopamine neurons. 11, 6286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Ouchi Y, et al. , 2009. Neuroinflammation in the living brain of Parkinson’s disease. Parkinsonism Relat Disord. 15 Suppl 3, S200–4. [DOI] [PubMed] [Google Scholar]
  227. Ouchi Y, et al. , 2005. Microglial activation and dopamine terminal loss in early Parkinson’s disease. Ann Neurol. 57, 168–75. [DOI] [PubMed] [Google Scholar]
  228. Panagiotakopoulou V, et al. , 2020. Interferon-γ signaling synergizes with LRRK2 in neurons and microglia derived from human induced pluripotent stem cells. Nature Communications. 11, 5163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  229. Perry VH, et al. , 1985. Immunohistochemical localization of macrophages and microglia in the adult and developing mouse brain. Neuroscience. 15, 313–26. [DOI] [PubMed] [Google Scholar]
  230. Perry VH, et al. , 2010. Microglia in neurodegenerative disease. Nature Reviews Neurology. 6, 193–201. [DOI] [PubMed] [Google Scholar]
  231. Perussia B, 1991. Lymphokine-activated killer cells, natural killer cells and cytokines. Current opinion in immunology. 3, 49–55. [DOI] [PubMed] [Google Scholar]
  232. Peter I, et al. , 2018. Anti-Tumor Necrosis Factor Therapy and Incidence of Parkinson Disease Among Patients With Inflammatory Bowel Disease. JAMA Neurol. 75, 939–946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Phatnani H, Maniatis T, 2015. Astrocytes in neurodegenerative disease. Cold Spring Harb Perspect Biol. 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Pierre DM, et al. , 1997. Somatic evolution in the immune system: the need for germinal centers for efficient affinity maturation. J Theor Biol. 186, 159–71. [DOI] [PubMed] [Google Scholar]
  235. Pike AF, et al. , 2022a. Dopamine signaling modulates microglial NLRP3 inflammasome activation: implications for Parkinson’s disease. Journal of Neuroinflammation. 19, 50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Pike AF, et al. , 2022b. Dopamine signaling modulates microglial NLRP3 inflammasome activation: implications for Parkinson’s disease. J Neuroinflammation. 19, 50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Pike AF, et al. , 2022c. The potential convergence of NLRP3 inflammasome, potassium, and dopamine mechanisms in Parkinson’s disease. npj Parkinson’s Disease. 8, 32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Pissadaki EK, Bolam JP, 2013. The energy cost of action potential propagation in dopamine neurons: clues to susceptibility in Parkinson’s disease. Frontiers in computational neuroscience. 7, 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  239. Poewe W, et al. , 2017. Parkinson disease. Nature reviews Disease primers. 3, 1–21. [DOI] [PubMed] [Google Scholar]
  240. Poly TN, et al. , 2019. Non-steroidal anti-inflammatory drugs and risk of Parkinson’s disease in the elderly population: a meta-analysis. European journal of clinical pharmacology. 75, 99–108. [DOI] [PubMed] [Google Scholar]
  241. Poulin J-F, et al. , 2016. Disentangling neural cell diversity using single-cell transcriptomics. 19, 1131–1141. [DOI] [PubMed] [Google Scholar]
  242. Pöyhönen S, et al. , 2019. Effects of Neurotrophic Factors in Glial Cells in the Central Nervous System: Expression and Properties in Neurodegeneration and Injury. Frontiers in Physiology. 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Pramod AB, et al. , 2013. SLC6 transporters: structure, function, regulation, disease association and therapeutics. Molecular aspects of medicine. 34, 197–219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Probert L, 2015. TNF and its receptors in the CNS: The essential, the desirable and the deleterious effects. Inflammation in Nervous System Disorders. 302, 2–22. [DOI] [PubMed] [Google Scholar]
  245. Qin XY, et al. , 2016. Aberrations in Peripheral Inflammatory Cytokine Levels in Parkinson Disease: A Systematic Review and Meta-analysis. JAMA Neurol. 73, 1316–1324. [DOI] [PubMed] [Google Scholar]
  246. Qiu YH, et al. , 2004. Expression of tyrosine hydroxylase in lymphocytes and effect of endogenous catecholamines on lymphocyte function. Neuroimmunomodulation. 11, 75–83. [DOI] [PubMed] [Google Scholar]
  247. Ransohoff RM, Brown MA, 2012. Innate immunity in the central nervous system. J Clin Invest. 122, 1164–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Ransohoff RM, et al. , 2003. Three or more routes for leukocyte migration into the central nervous system. Nature Reviews Immunology. 3, 569–581. [DOI] [PubMed] [Google Scholar]
  249. Rees K, et al. , 2011. Non‐steroidal anti‐inflammatory drugs as disease‐modifying agents for Parkinson’s disease: evidence from observational studies. Cochrane Database of Systematic Reviews. [DOI] [PubMed] [Google Scholar]
  250. Reeve A, et al. , 2014. Ageing and Parkinson’s disease: why is advancing age the biggest risk factor? Ageing research reviews. 14, 19–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Reguzzoni M, et al. , 2002. Ultrastructural localization of tyrosine hydroxylase in human peripheral blood mononuclear cells: effect of stimulation with phytohaemagglutinin. Cell Tissue Res. 310, 297–304. [DOI] [PubMed] [Google Scholar]
  252. Ribeiro M, et al. , 2019. Meningeal γδ T cell-derived IL-17 controls synaptic plasticity and short-term memory. Sci Immunol. 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Riedhammer C, Weissert R, 2015. Antigen Presentation, Autoantigens, and Immune Regulation in Multiple Sclerosis and Other Autoimmune Diseases. Frontiers in Immunology. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  254. Rosczyk H, et al. , 2008. Neuroinflammation and cognitive function in aged mice following minor surgery. Experimental gerontology. 43, 840–846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Ruszkiewicz J, Albrecht J, 2015. Changes in the mitochondrial antioxidant systems in neurodegenerative diseases and acute brain disorders. Neurochemistry international. 88, 66–72. [DOI] [PubMed] [Google Scholar]
  256. Sagar D, et al. , 2012. Mechanisms of dendritic cell trafficking across the blood-brain barrier. J Neuroimmune Pharmacol. 7, 74–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Saha K, et al. , 2014. Intracellular methamphetamine prevents the dopamine-induced enhancement of neuronal firing. J Biol Chem. 289, 22246–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  258. Salvador AF, et al. , 2021. Neuromodulation by the immune system: a focus on cytokines. 21, 526–541. [DOI] [PubMed] [Google Scholar]
  259. Sambo DO, et al. , 2017. The sigma-1 receptor modulates methamphetamine dysregulation of dopamine neurotransmission. Nat Commun. 8, 2228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Samii A, et al. , 2009. NSAID use and the risk of Parkinson’s disease. Drugs & aging. 26, 769–779. [DOI] [PubMed] [Google Scholar]
  261. Sanchez-Padilla J, et al. , 2014. Mitochondrial oxidant stress in locus coeruleus is regulated by activity and nitric oxide synthase. Nat Neurosci. 17, 832–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  262. Sarkar S, et al. , 2020. Kv1.3 modulates neuroinflammation and neurodegeneration in Parkinson’s disease. J Clin Invest. 130, 4195–4212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Sarlus H, Heneka MT, 2017. Microglia in Alzheimer’s disease. J Clin Invest. 127, 3240–3249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Sawada M, et al. , 2006. Role of cytokines in inflammatory process in Parkinson’s disease. J Neural Transm Suppl. 373–81. [DOI] [PubMed] [Google Scholar]
  265. Schäfers M, et al. , 2003. Increased Sensitivity of Injured and Adjacent Uninjured Rat Primary Sensory Neurons to Exogenous Tumor Necrosis Factor-α after Spinal Nerve Ligation. The Journal of Neuroscience. 23, 3028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  266. Schapira AH, et al. , 2014. Slowing of neurodegeneration in Parkinson’s disease and Huntington’s disease: future therapeutic perspectives. Lancet. 384, 545–55. [DOI] [PubMed] [Google Scholar]
  267. Schröder M, et al. , 2018. Genetic screen in myeloid cells identifies TNF-α autocrine secretion as a factor increasing MDSC suppressive activity via Nos2 up-regulation. Scientific Reports. 8, 13399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Schwartz M, et al. , 2013. How do immune cells support and shape the brain in health, disease, and aging? Journal of Neuroscience. 33, 17587–17596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Segueni N, et al. , 2016. Innate myeloid cell TNFR1 mediates first line defence against primary Mycobacterium tuberculosis infection. Scientific Reports. 6, 22454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Serafini B, et al. , 2004. Detection of ectopic B-cell follicles with germinal centers in the meninges of patients with secondary progressive multiple sclerosis. Brain Pathol. 14, 164–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Shaerzadeh F, et al. , 2020. Microglia senescence occurs in both substantia nigra and ventral tegmental area. Glia. 68, 2228–2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Shen Y, et al. , 2022. CCR5 closes the temporal window for memory linking. 606, 146–152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Sierra A, et al. , 2007. Microglia derived from aging mice exhibit an altered inflammatory profile. Glia. 55, 412–424. [DOI] [PubMed] [Google Scholar]
  274. Simola N, et al. , 2007. The 6-hydroxydopamine model of Parkinson’s disease. Neurotox Res. 11, 151–67. [DOI] [PubMed] [Google Scholar]
  275. Singhania A, et al. , 2021. The TCR repertoire of α-synuclein-specific T cells in Parkinson’s disease is surprisingly diverse. Scientific Reports. 11, 302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Singleton AB, et al. , 2003. alpha-Synuclein locus triplication causes Parkinson’s disease. Science. 302, 841. [DOI] [PubMed] [Google Scholar]
  277. Smajić S, et al. , 2022. Single-cell sequencing of human midbrain reveals glial activation and a Parkinson-specific neuronal state. Brain. 145, 964–978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Sofroniew MV, 2015a. Astrocyte barriers to neurotoxic inflammation. Nature Reviews Neuroscience. 16, 249–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. Sofroniew MV, 2015b. Astrogliosis. Cold Spring Harbor perspectives in biology. 7, a020420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Sofroniew MV, Vinters HV, 2010. Astrocytes: biology and pathology. Acta neuropathologica. 119, 7–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Sommer C, et al. , 1998. Hyperalgesia in experimental neuropathy is dependent on the TNF receptor 1. Exp Neurol. 151, 138–42. [DOI] [PubMed] [Google Scholar]
  282. Sonninen T-M, et al. , 2020. Metabolic alterations in Parkinson’s disease astrocytes. Scientific Reports. 10, 14474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  283. Stellwagen D, et al. , 2005. Differential regulation of AMPA receptor and GABA receptor trafficking by tumor necrosis factor-alpha. J Neurosci. 25, 3219–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Stellwagen D, Malenka RC, 2006. Synaptic scaling mediated by glial TNF-alpha. Nature. 440, 1054–9. [DOI] [PubMed] [Google Scholar]
  285. Stemkowski PL, et al. , 2021. Are sensory neurons exquisitely sensitive to interleukin 1β? J Neuroimmunol. 354, 577529. [DOI] [PubMed] [Google Scholar]
  286. Stemkowski PL, et al. , 2015. Increased excitability of medium-sized dorsal root ganglion neurons by prolonged interleukin-1β exposure is K(+) channel dependent and reversible. J Physiol. 593, 3739–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  287. Strazielle N, et al. , 2016. T-Lymphocytes Traffic into the Brain across the Blood-CSF Barrier: Evidence Using a Reconstituted Choroid Plexus Epithelium. PLoS One. 11, e0150945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Streit WJ, 1993. Microglial-neuronal interactions. J Chem Neuroanat. 6, 261–6. [DOI] [PubMed] [Google Scholar]
  289. Streit WJ, 2002. Microglia as neuroprotective, immunocompetent cells of the CNS. Glia. 40, 133–9. [DOI] [PubMed] [Google Scholar]
  290. Streit WJ, Graeber MB, 1993. Heterogeneity of microglial and perivascular cell populations: insights gained from the facial nucleus paradigm. Glia. 7, 68–74. [DOI] [PubMed] [Google Scholar]
  291. Streit WJ, et al. , 2004. Dystrophic microglia in the aging human brain. Glia. 45, 208–212. [DOI] [PubMed] [Google Scholar]
  292. Subramaniam M, et al. , 2014. Mutant α-synuclein enhances firing frequencies in dopamine substantia nigra neurons by oxidative impairment of A-type potassium channels. J Neurosci. 34, 13586–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Sulzer D, et al. , 2017. T cells from patients with Parkinson’s disease recognize α-synuclein peptides. Nature. 546, 656–661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Surmeier DJ, et al. , 2012. Physiological phenotype and vulnerability in Parkinson’s disease. Cold Spring Harb Perspect Med. 2, a009290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Surmeier DJ, et al. , 2017. Selective neuronal vulnerability in Parkinson disease. 18, 101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Süβ P, et al. , 2021. Chronic peripheral inflammation: a possible contributor to neurodegenerative diseases. Neural Regen Res. 16, 1711–1714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Tabrez S, et al. , 2012. A synopsis on the role of tyrosine hydroxylase in Parkinson’s disease. CNS Neurol Disord Drug Targets. 11, 395–409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Tan E-K, et al. , 2020. Parkinson disease and the immune system — associations, mechanisms and therapeutics. Nature Reviews Neurology. 16, 303–318. [DOI] [PubMed] [Google Scholar]
  299. Tansey MG, et al. , 2022. Inflammation and immune dysfunction in Parkinson disease. Nat Rev Immunol. 1–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Taylor EL, et al. , 2003. Nitric oxide: a key regulator of myeloid inflammatory cell apoptosis. Cell Death Differ. 10, 418–30. [DOI] [PubMed] [Google Scholar]
  301. Terland O, et al. , 1997. Dopamine oxidation generates an oxidative stress mediated by dopamine semiquinone and unrelated to reactive oxygen species. J Mol Cell Cardiol. 29, 1731–8. [DOI] [PubMed] [Google Scholar]
  302. Tomasoni R, et al. , 2017. Lack of IL-1R8 in neurons causes hyperactivation of IL-1 receptor pathway and induces MECP2-dependent synaptic defects. Elife. 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Trist BG, et al. , 2019. Oxidative stress in the aging substantia nigra and the etiology of Parkinson’s disease. Aging Cell. 18, e13031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  304. Troncoso-Escudero P, et al. , 2018. Outside in: Unraveling the Role of Neuroinflammation in the Progression of Parkinson’s Disease. Front Neurol. 9, 860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. van Langelaar J, et al. , 2020. B and T Cells Driving Multiple Sclerosis: Identity, Mechanisms and Potential Triggers. Frontiers in Immunology. 11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Varvel NH, et al. , 2016. Infiltrating monocytes promote brain inflammation and exacerbate neuronal damage after status epilepticus. Proc Natl Acad Sci U S A. 113, E5665–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Vaughan RA, Foster JD, 2013. Mechanisms of dopamine transporter regulation in normal and disease states. Trends Pharmacol Sci. 34, 489–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Venkateshappa C, et al. , 2012. Increased oxidative damage and decreased antioxidant function in aging human substantia nigra compared to striatum: implications for Parkinson’s disease. Neurochemical research. 37, 358–369. [DOI] [PubMed] [Google Scholar]
  309. Veroni C, et al. , 2010. Activation of TNF receptor 2 in microglia promotes induction of anti-inflammatory pathways. Mol Cell Neurosci. 45, 234–44. [DOI] [PubMed] [Google Scholar]
  310. Vila M, et al. , 2001. The role of glial cells in Parkinson’s disease. Curr Opin Neurol. 14, 483–9. [DOI] [PubMed] [Google Scholar]
  311. Viviani B, et al. , 2003. Interleukin-1beta enhances NMDA receptor-mediated intracellular calcium increase through activation of the Src family of kinases. J Neurosci. 23, 8692–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. Wang P, et al. , 2022a. Global Characterization of Peripheral B Cells in Parkinson’s Disease by Single-Cell RNA and BCR Sequencing. Front Immunol. 13, 814239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  313. Wang Y, et al. , 2022b. Interleukin 33-mediated inhibition of A-type K(+) channels induces sensory neuronal hyperexcitability and nociceptive behaviors in mice. Theranostics. 12, 2232–2247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Welser-Alves JV, Milner R, 2013. Microglia are the major source of TNF-α and TGF-β1 in postnatal glial cultures; regulation by cytokines, lipopolysaccharide, and vitronectin. Neurochem Int. 63, 47–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Weng M, et al. , 2018. The Sources of Reactive Oxygen Species and Its Possible Role in the Pathogenesis of Parkinson’s Disease. Parkinsons Dis. 2018, 9163040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  316. Wijeyekoon RS, et al. , 2018. Monocyte Function in Parkinson’s Disease and the Impact of Autologous Serum on Phagocytosis. Frontiers in Neurology. 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Williams GP, et al. , 2022. Central and Peripheral Inflammation: Connecting the Immune Responses of Parkinson’s Disease. J Parkinsons Dis. 12, S129–s136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  318. Wilms H, et al. , 2003. Activation of microglia by human neuromelanin is NF‐κB‐dependent and involves p38 mitogen‐activated protein kinase: implications for Parkinson’s disease. The FASEB journal. 17, 1–20. [DOI] [PubMed] [Google Scholar]
  319. Wobst HJ, et al. , 2020. The clinical trial landscape in amyotrophic lateral sclerosis-Past, present, and future. Med Res Rev. 40, 1352–1384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  320. Wong KL, et al. , 2012. The three human monocyte subsets: implications for health and disease. Immunologic research. 53, 41–57. [DOI] [PubMed] [Google Scholar]
  321. Wyss-Coray T, Mucke L, 2002. Inflammation in neurodegenerative disease—a double-edged sword. Neuron. 35, 419–432. [DOI] [PubMed] [Google Scholar]
  322. Zhang W, et al. , 2005. Aggregated alpha-synuclein activates microglia: a process leading to disease progression in Parkinson’s disease. Faseb j. 19, 533–42. [DOI] [PubMed] [Google Scholar]
  323. Zhang Y, et al. , 2016. Purification and Characterization of Progenitor and Mature Human Astrocytes Reveals Transcriptional and Functional Differences with Mouse. Neuron. 89, 37–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  324. Zhou Z, et al. , 2009. Interleukin-10 provides direct trophic support to neurons. J Neurochem. 110, 1617–27. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES