Skip to main content

Exercise training upregulates CD55 to suppress complement-mediated synaptic phagocytosis in Parkinson’s disease

Abstract

The primary pathological change in Parkinson’s disease (PD) is the progressive degeneration of dopaminergic neurons in the substantia nigra. Additionally, excessive microglial activation and synaptic loss are also typical features observed in PD samples. Exercise trainings have been proven to improve PD symptoms, delay the disease progression as well as affect excessive microglial synaptic phagocytosis. In this study, we established a mouse model of PD by injecting mouse-derived α-synuclein preformed fibrils (M-α-syn PFFs) into the substantia nigra, and demonstrated that treadmill exercise inhibits microglial activation and synaptic phagocytosis in striatum. Using RNA-Seq and proteomics, we also found that PD involves excessive activation of the complement pathway which is closely related to over-activation of microglia and abnormal synaptic function. More importantly, exercise training can inhibit complement levels and complement-mediated microglial phagocytosis of synapses. It is probably triggered by CD55, as we observed that CD55 in the striatum significantly increased after exercise training and up-regulation of that molecule rescued motor deficits of PD mice, accompanied with reduced microglial synaptic phagocytosis in the striatum. This research elucidated the interplay among microglia, complement, and synapses, and analyzed the effects of exercise training on these factors. Our work also suggested CD55 as a complement-relevant candidate molecule for developing therapeutic strategies of PD.

Graphical Abstract

Introduction

Parkinson’s disease (PD) is the second most common neurodegenerative disorder, with incidence and mortality rates increasing globally each year [1]. Clinically, PD manifests through hallmark symptoms such as resting tremors, muscular rigidity, reduced movement, and abnormal posture and gait [2,3,4]. The primary pathological change in PD is the progressive degeneration of dopaminergic neurons in the substantia nigra pars compacta of the midbrain [5,6,7]. Additionally, excessive microglial activation [8,9,10,11,12,13] and synaptic loss [14, 15] are also typical features observed in PD samples. In the early stages of PD, there is already a loss of synapse in the striatal region [16], accompanied by a reduction of dopamine transporter protein [17]. Loss of synapse accelerated the progress of PD, but no pharmacological treatments which can reverse the decline of synapse has been developed currently [18, 19].

Exercise trainings have been proven to improve symptoms and delay the progression of PD [20]. Previous studies have shown that exercise trainings can enhance the balance, gait, and overall functional capacity [20,21,22], and reduce the incidence of falls in PD patients [23]. Our own group has also validated these findings, showcasing significant enhancements in motor abilities post-rehabilitation training [24]. Through animal experiments, researchers have not only observed related molecular changes, but also explored underlying mechanisms which may help to enhance the efficacy of exercise training and lay the groundwork for identifying new therapeutic targets for PD [25,26,27,28]. After exercise training, the number of microglia in the striatum tends to decrease, with more of those shifting into the resting state [29] compared to the MPTP-PD mice [30]. Concurrently, exercise trainings rescue the down-regulation of PSD95 and synaptophysin in the striatum of MPTP-treated mice [31]. That indicated synaptic function in the striatal region is enhanced. Experiment with electron microscopy also revealed significant neurogenesis of dendrites and axons in the striatal region following exercise [32]. Nevertheless, further research is needed to determine whether microglia associate with synaptic loss in PD and whether exercise training reduces synaptic loss in PD by regulating microglia.

Microglia are known to participate in synaptic phagocytosis in various diseases, and synaptic alterations are early events in many neurodegenerative diseases. In models of Alzheimer’s disease, microglia act as potential and key cellular mediators of synaptic loss [33]. Increased microglia-mediated synaptic loss through phagocytosis exacerbates cognitive impairment [34]. Also, persistent microglial phagocytic activity results in decreased synaptic density and cognitive decline after stroke [35, 36] and in multiple sclerosis [37]. Similarly, excessive microglial synaptic phagocytosis has been observed in various PD animals [38,39,40,41]. Exercise training was reported to effectively restrict microglia-synapse engulfment. For example, long-term voluntary exercise can protect hippocampal synapses in APP/PS1 mice, reduce microglial activity and dendritic spine loss [42, 43]. In a TDP-43-induced model of ALS, treadmill exercise was found to prevent the worsening of motor dysfunction and further inhibit microglial synaptic phagocytosis [44]. However, how exercise training regulates microglial synaptic phagocytosis in PD is still not fully understood.

In this study, we employed treadmill training for PD mice to investigate how exercise regulates microglia-mediated synaptic phagocytosis in PD. We found that exercise rescues motor deficit, protects dopaminergic neurons, inhibits microglial activation and complement-mediated synaptic phagocytosis in PD utilizing immunocytochemistry, RNA-seq and proteomics. Additionally, CD55 probably accounts for the attenuation of complement-mediated synaptic phagocytosis in the striatum after treadmill training. Our work elucidated the interplay among microglia, complement and synapses. Moreover, a novel complement-related mechanism via which physical exercise exert neuroprotection in a mouse model of Parkinson’s disease was explored.

Materials and methods

Animals

C57BL/6 J mice (6 or 8 weeks old, males) were purchased from Shanghai SLAC Experimental Animal Co., Ltd. 5 mice per cage were housed in a SPF environment equipped with a ventilation system. The ambient temperature was maintained at 22 °C ± 2 °C, and relative humidity ranged from 40 to 70%. The mice were kept in a 12-h light/12-h dark cycle (7:00–19:00) with access to ample food and water. All animal experiments were conducted in accordance with the guidelines for the care and use of laboratory animals from the National Institutes of Health and were approved by the Ethics Committee of Tongji University.

Mouse model of Parkinson’s disease

The mouse model of PD was established by injecting mouse-derived alpha-synuclein preformed fibrils (M-α-syn PFFs) (AnaSpec, AS-56082-100) into the substantia nigra. Eight-week-old male mice were divided into Sham and PFFs groups. Animals were assigned a number, then use a random number table to assign the animals into groups. PFFs group received bilateral injections of PFFs into the substantia nigra (P3.0 mm, L/R1.3 mm, V4.5 mm), with a dosage of 2 μg per side at a continuous flow rate of 0.2 μl/min, while Sham group received an equivalent volume of saline. Behavior tests were performed weekly.

Behavior test

Behavior tests included pole test and rotarod test. The rotarod test included three training sessions, with speeds of 10 rpm/min constant, 2–20 rpm/min constant acceleration, and 4–40 rpm/min constant acceleration. Mice were then tested on the rotarod at 4–40 rpm/min constant acceleration for 5 min. Each experiment was conducted in triplicate, and the intervals between experiments exceeded thirty minutes. In the pole test, mice were trained to climb a vertical pole (50 cm height, 0.5 cm diameter) with a textured surface. The time taken for mice to climb from the top to the bottom was recorded in three repetitions, with intervals exceeding fifteen minutes between trials.

Tissue preparation

After behavior tests, brain tissues of mice were immediately collected. Following sodium pentobarbital anesthesia, blood was removed from the body using 1 × PBS, and the tissues were fixed by perfusion with PFA. After sucrose gradient dehydration, the tissues were embedded in OCT and stored at − 80 °C. For another subset of mice, brain tissues were dissected under a stereomicroscope after cervical dislocation, and the substantia nigra and striatum were isolated. The obtained tissues were rapidly frozen in liquid nitrogen and transferred to − 80 °C for storage.

Western blot analysis

Brain tissues were lysed using RIPA lysis buffer (Beyotime,P0013B) supplemented with protease inhibitor mixture (Roche, 4693116001) and phosphatase inhibitor mixture (Roche, 4906845001). BCA assay kit (Beyotime, P0010) was used to measure the total protein concentration. Proteins were separated by SDS-PAGE, and then transferred to a PVDF membrane. Membranes were blocked in 5% skimmed milk for 1 h, incubated overnight with specific primary antibodies at 4 °C, washed three times in 1 × TBST, incubated with secondary antibodies at room temperature for 1 h, and finally detected using the ECL protein blotting substrate kit. ImageJ software was used for quantitative analysis of band density.

Immunofluorescence

Frozen brain section (30 μm thickness) were collected for immunofluorescence. The substantia nigra and striatum brain slices were washed three times for 10 min each in 1 × PBS and then blocked with immunofluorescence blocking solution (Beyotime, P0260) for 1 h. After blockage, the slices were incubated overnight with primary antibodies at 4 °C. After three washes with 1 × PBS, the slices were incubated with secondary antibodies at room temperature for 1 h. Following three washes with 1 × PBS, the slices were coverslipped using mounting medium and imaged using a Leica confocal laser scanning microscope. 63 × oil objective was utilized to collect microglia confocal images. The microglia were scanned from top to bottom with 0.3-µm steps in the z direction with 2048 × 2048-pixel resolution and all collected images were merged to one image. ImageJ software was employed for image processing.

Treadmill exercise protocol

After three weeks of PFFs administration, PFFs group was divided into PFFs and Exer groups. Exer group participated in treadmill exercise training, consisting of 5 days per week, 40 min per day, with a speed ranging up to 15 m/min (6 m/min for 5 min, 9 m/min for 5 min, 12 m/min for 20 min, 15 m/min for 5 min, and 12 m/min for 5 min) for 4 weeks. Behavior tests were conducted after each exercise session, and after completion of the exercise regimen, mice were sacrificed to collect samples for subsequent analysis.

RNA-sequence

RNA-Seq was conducted by GENE DENOVO. Total RNA was extracted using Trizol reagent kit. After reverse transcription and RNA amplification, RNA libraries were constructed. Prior to information analysis, raw data underwent data filtering to reduce analysis interference caused by invalid data. Preliminary filtering was performed on the obtained valid data, removing genes with zero expression in over half of the samples, and data normalization was conducted to ensure greater consistency within samples. PCA analysis was performed on three groups to ensure good clustering within each group. Further, differential expression gene (DEGs) analysis was conducted pairwise among the three groups, and then KEGG and GO functional enrichment analysis using DEGs were conducted. Foldchange ≥ 2 or ≤ − 2, and p-value ≤ 0.05 were criteria for DEGs.

Proteomics

4D-DIA relative quantitative proteomics by GENE DENOVO were applied in our study. Proteins were extracted from the samples, and the total concentration was determined using BCA assay. Subsequently, peptide samples were prepared and subjected to mass spectrometry detection. This process involved comparing the obtained mass spectrometry data to the reference library to identify and quantify the proteins present in each sample. The obtained protein quantification results were statistically analyzed and subjected to bioinformatics analysis. After inter-sample correction using the Limma package [45], differential multiples were calculated using the Limma package, and Gene Set Enrichment Analysis (GSEA) was performed using the ClusterProfiler package [46]. Foldchange ≥ 1.2 or ≤ 0.8, and p-value ≤ 0.05 were criteria for Differential Expression Protein.

Enzyme-linked immunosorbent assay (ELISA)

After blood collection, samples were put on ice for 30 min. Subsequently, centrifugation was performed at 12,000 rpm for 15 min to obtain the serum. Striatal tissues were homogenized using lysis buffer, and the supernatant was collected after centrifugation at 12,000 rpm for 15 min. ELISA kits were used to quantify dopamine (DA) level in the striatum and complement levels (C3 and C1q) in the serum. The experimental procedures were performed according to the manufacturer’s instructions.

Golgi staining

FD Rapid Golgi Stain Kit (FDNeurotech, PK401) was used for Golgi staining according to the manufacturer’s instructions. After cervical dislocation, mouse brain tissues were separated and washed in PBS to remove blood. The brain tissues were immersed in a mixture of Solution A and Solution B (1:1) for 14 days and then transferred to Solution C for dehydration, with the entire process conducted in the dark. After dehydration, brain tissues were sliced into 200 μm sections using a vibratome, and the sections were collected on chrome-alum-gelatin-coated slides. Stained slices were imaged using a Zeiss confocal microscope under bright-field illumination, and images of the striatum were collected for subsequent analysis.

Virus injection

An adeno-associated virus (AAV) expressing CD55 (pAAV-EF1A-mCD55-mCherry:WPRE) (the AAV was constructed and packaged by VectorBuilder) was constructed and injected into the striatum (A0.8 mm, L/R1.2 mm, V3.6 mm). Virus injection included treatment and pre-treatment experiments. In the treatment experiment, eight-week-old male C57BL/6 J mice were divided into Sham, PFFs, and CD55-OE (overexpression) groups. PFFs and CD55-OE groups underwent PD modeling by receiving PFFs in the substantia nigra. CD55-OE group received AAV injection in the striatum, while Sham group and PFFs group received equivalent volumes of saline. Behavior tests, including pole test and rotarod test, were performed weekly, and mice were sacrificed for brain retrieval after three weeks. In the pre-treatment experiment, six-week-old mice were divided into Sham, PFFs, and CD55-OE groups. CD55-OE group received AAV injection in the striatum, while the PFFs and Sham groups received equivalent volumes of negative virus in the striatum. After two weeks, PFFs and CD55-OE groups underwent PD modeling in the substantia nigra, while Sham group received saline. Behavior tests were conducted weekly, and mice were sacrificed for brain retrieval three weeks after PD modeling.

Antibody list

Antibody

Vendor

Catalog

Dilution

RRID

TH

Merck

T1299

WB1:5000

IF1:5000

AB_477560

TH

Abcam

ab76442

IF1:200

AB_1524535

iba1

Abcam

ab5076

IF1:200

AB_2224402

iba1

Wako

019-19741

IF1:200

AB_839504

GAPDH

Proteintech

81640-5-RR

WB1:5000

AB_3086559

β-actin

Proteintech

20536-1-AP

WB1:5000

AB_10700003

Tubulin

Proteintech

11224-1-AP

WB1:5000

AB_2210206

CD68

CST

#97778

IF1:200

AB_2928056

C3

Proteintech

21337-1-AP

WB1:1000

AB_2878843

C3

Invitrogen

PA1-29715

WB1:1000

AB_2066730

PSD95

CST

#36233

WB1:1000

IF1:200

AB_2721262

C1q

Abcam

ab11861

IF1:200

AB_298643

C1q

Abcam

ab182451

IF1:200

AB_2732849

CD55

Proteintech

26580-1-AP

WB1:1000

AB_2880559

Statistical analysis

GraphPad Prism V.9 was used for all statistical analyses. All data are presented as mean ± SD. Normality of data was assessed, and statistical analysis was performed using unpaired two-tailed t-tests or one-way ANOVA. Differences between two groups were determined by unpaired two-tailed Student’s t-test. When comparing more than two groups, one-way ANOVA analysis was used, followed by Tukey’s post hoc test. Significance between groups was evaluated as follows: ns p > 0.05, *p < 0.05, **p < 0.01, ***p < 0.001, and ****p < 0.0001.

Results

Injection of α-syn PFFs into the substantia nigra effectively induces behavioral and pathological changes associated with Parkinson’s disease in wild-type mice

Injection of pre-formed fibrils of α-synuclein (PFFs) has been an alternative method proposed for modeling PD by increasing the level of α-synuclein in the substantia nigra. This model can more closely replicate the physiological processes that may occur in the brain of PD patients [47]. Given that the substantia nigra is the primary brain region affected in PD, injection into the substantia nigra allows for a more direct observation of the impact of α-syn on dopaminergic neurons. Therefore, we chose the substantia nigra as the main injection site. Prior to PFFs injection, we conducted baseline Behavior tests on mice to ensure no significant intergroup differences. The modeling lasted for three weeks, during which we performed behavior tests including pole test and rotarod test every week. The results showed that after PFFs injection three weeks, mice in PFFs group exhibited a decline in behavioral abilities, with reduced latency time on the rotarod and prolonged climbing time compared to Sham group, which showed statistical differences (Fig. 1a, b). To further evaluate the pathological changes, we validated TH protein levels in the two core brain areas of PD, substantia nigra (SN) and striatum. The results showed a significant decrease in TH protein levels in both the substantia nigra and striatum of PFFs group compared to Sham group (Fig. 1 c). The number of TH-positive neurons in the substantia nigra also declined in PFFs group, according to the results of immunofluorescence assay. That reflected the loss and death of nigral dopaminergic neurons in PFFs-treated mice (Fig. 1d). Additionally, we assessed microglia, the central immune cells involved in the development of Parkinson’s disease in various groups. We chose CD68 as an indicator of microglial activation, and detected an upregulation of nigral/striatal microglia in PFFs group, along with increased number of microglia and an elevated proportion of CD68 expression per microglia, suggesting that microglial aggregation and over-activation occur in the substantia nigra and striatum in the PFFs-induced mouse model (Fig. 1e–h). Moreover, PFFs administration significantly suppressed the expression of postsynaptic density protein 95 (PSD95) in the striatum, indicating the synaptic loss under that condition (Fig. 1i). These results showed that PFFs administration into the substantia nigra of wild-type mice for three weeks can significantly induce motor skill decline, loss of dopaminergic neurons in the substantia nigra, activation of microglia, and loss of synapses in the striatum.

Fig. 1
figure 1

Behavioral and pathological changes in the PFF-induced PD mouse model. a Results of rotarod test in Sham and PFFs groups over time and at day21 (Sham group n = 8, PFFs group n = 21). b Results of poles test in Sham and PFFs groups over time and at day21. (Sham group n = 10, PFFs group n = 22) c Western blot showing the expression of TH protein in the substantia nigra (SN) and striatum. (n = 6 per group) d Representative images of immunostaining and relative quantitative analysis of dopamine neurons in the substantia nigra. Scale bar = 1000 μm. (n = 4 per group) e, f Representative images of immunostaining and relative quantitative analysis of active microglia in the substantia nigra. Scale bar = 20 μm. (n = 4 per group) (n = 20 per group from 4 mice each group) g, h Representative images of immunostaining and relative quantitative analysis of active microglia in the striatum. Scale bar = 20 μm. (n = 4 per group) (n = 20 per group from 4 mice each group) i Results of western blot analysis for the striatal expression of PSD95 protein in Sham and PFFs group. (n = 6 per group) All data are presented as mean ± SD. Statistical analysis was performed using unpaired two-tailed t-tests

Exercise training improves behavioral and pathological outcomes in PFFs mice

In previous studies and earlier researches of our team, we found that exercise training could improve motor skill levels and slow down the disease progression in both PD patients and mouse models. To further explore the specific mechanism of exercise training, we conducted treadmill training for PFFs mice. After three weeks of α-syn PFFs administration, mice were forced to participate in a four-week treadmill training, five times per week, 40 min each time (Fig. 2a, b). Behavior tests were conducted weekly. After exercise training, we assessed the mice’s body weight, and the results showed no significant differences among the groups (Fig. 2c). The results of pole test and rotarod test showed a significant decline in motor skills in PFFs group compared to Sham group, while exercise training partially rescued the motor ability of PFFs mice (Fig. 2d, e). Similarly, results of immunofluorescence staining showed that, compared to Sham group, PFFs mice exhibited a significant reduction in TH-positive neurons in the substantia nigra, while that change was recovered in PFFs group, indicating that exercise training could reduce damage to dopaminergic neurons and delay their loss (Fig. 2f). For microglia in the substantia nigra, they were prone to accumulate and shift into an activated state in PFFs group, compared to Sham group. While after exercise training, there was a downregulation of microglia in the substantia nigra region, accompanied by a decreased proportion of cells in the activated state (Fig. 2g, h). Also, we assessed the synaptic protein PSD95 in the striatum among various groups. Western blot data indicated a significant reduction in PSD95 in PFFs group compared to Sham group, while exercise training attenuated the loss of PSD95. Similarly, results of Golgi staining also revealed dendritic spine loss in PFFs group, which could be alleviated by exercise training (Fig. 2i). In order to further explore the relationship between microglia and synapses as well as the effect of exercise training, we conducted immunofluorescence staining and observed the co-localization of CD68 and PSD95 in the striatum. The results showed that activated microglia were more abundant in the striatum of PFFs group compared to Sham group. The co-localization level of CD68 and PSD95 was increased after PFFs injection, which suggested an increased phagocytosis of microglia. After exercise training, the co-localization of CD68 and PSD95 in microglia decreased, indicating that exercise training could suppress microglial synaptic phagocytosis (Fig. 2j). Overall, these results demonstrated that exercise training can effectively inhibit the activation of microglia and reduce the loss of synapse in the striatum.

Fig. 2
figure 2

Exercise training pattern and behavioral, pathological evaluation. a Experimental paradigm of exercise training. b Exercise training pattern. c Analysis of mouse body weight. (n = 7 per group) (d) Results of rotarod test among three groups over time and the last of exercise training (Day49). (Sham group n = 11, PFFs group n = 10, Exer group n = 11) e Results of poles test among three groups over time and at the last day of exercise training (Day49). (Sham group n = 10, PFFs group n = 11, Exer group n = 11) f Representative images of immunostaining and relative quantitative analysis of dopamine neurons in the substantia nigra. Scale bar = 1000 μm. (n = 4 per group) g Representative images of immunostaining and relative quantitative analysis of dopamine neurons and microglia in the substantia nigra. Scale bar = 20 μm. (n = 4 per group) (n = 20 per group from 4 mice each group) i Western blot and Golgi staining results of striatal PSD95 expression and spine density in the striatum. (n = 6 per group) (n = 18 per group from 3 mice each group) j Representative images of immunostaining and relative quantitative analysis of phagocytic function of microglia in the striatum. Scale bar = 20 μm. (n = 4 per group) (n = 20 per group from 4 mice each group) (n = 4 per group) All data are presented as mean ± SD. Statistical analysis was performed using one-way ANOVA

Proteomics and RNA-seq reveal that exercise improves synaptic function and inhibits upregulation of the complement pathway in PD mice

To further investigate how exercise training regulates microglial synaptic phagocytosis in PD, we analyzed the nigral RNA-Seq data and found that Sham group and Exer group clustered closer together in PCA dimensional reduction analysis, suggesting that exercise corrected transcriptional changes induced by PFFs (Fig. 3a). 201 differential genes were identified between Sham and PFFs group, and 179 were achieved between PFFs and Exer group (Fig. 3b). Functional analysis of upregulated genes in PFFs group compared to Sham group revealed significant enrichment in pathways such as Phagosome and Complement and coagulation (Fig. 3c), while downregulated genes were enriched in the neuroactive ligand-receptor interaction pathway (Figure S1a). Compared to PFFs group, Exer group showed downregulation of chemokine-related pathways (Fig. 3d) and upregulation of immune response pathways (Figure S1b). We confirmed these results in the transcriptomic sequencing data. We found that lysosome-related protein CD68 was upregulated in PFFs group but downregulated after exercise (Figure S1c). Additionally, inflammatory indicators of microglia (Cst3 and Cst7) exhibited a similar trend for expression patterns (Figure S1d). Various components of the complement cascade, including C1q, C3, and C4b, were significantly upregulated in PFFs group, while exercise training suppressed this abnormal upregulation (Figure S1e). Further, we conducted proteomics in the substantia nigra and striatum regions using 4D-DIA technology. From results of the nigral proteomics, we detected 96 differentially expressed proteins between Sham and PFFs group, as well as 235 between PFFs and Exer group (Fig. 3e). Results of GESA also pointed to pathways “complement and coagulation cascades”, which was upregulated in PFFs group compared to Sham group, and downregulated after exercise training (Fig. 3f). In agreement with transcriptomic sequencing data, various components of the “complement and coagulation cascades” (such as key protein C3) were upregulated in PFFs group but downregulated after exercise training (Fig. 3g) in Proteomics. The expression trend of proteins involved in the “complement and coagulation cascades” pathway in the striatum was similar to that in the substantia nigra (Fig. 3h, i). Taken together, we believe that changes related to the complement system play a crucial role in the onset and progression of PD, as well as in the improvement of the motor deficit after physical exercise. Moreover, complement may be an important node to connect microglia and synapse loss.

Fig. 3
figure 3

Proteomic and RNA-Seq data analysis revealed the protective effect of exercise and de-regulated complement pathway after exercise. a PCA plot of RNA-Seq data for three groups. b Volcano plot of DEGs. c PFFs group vs Sham group up-regulated genes enrichment pathway. d Exer group vs PFFs group down-regulated genes enrichment pathway. e Differential expression proteins in the substantia nigra of three groups. f GSEA showing the significant changed pathway in the substantia nigra of Sham group vs PFFs group and PFFs group vs Exer group. g Relative expression of proteins associated with “Complement and Coagulation cascades” pathway in the substantia nigra of three groups. h Differential expression proteins in the striatum of three groups. i Heatmap of relative expression of proteins associated with “Complement and Coagulation cascades” pathway in the striatum of three groups

Exercise training suppresses upregulation of complement C3 and reduces synaptic phagocytosis by microglia

Based on RNA-Seq data and proteomic results described above, we first evaluate the alterations of the complement system in the peripheral blood of mice under various conditions. As we expected, serum concentrations of complement C3 and C1q were higher in PFFs group compared to Sham group, while treadmill exercise training markedly reduced those complement levels in the peripheral blood of PFFs-treated mice (Fig. 4a, b). Furthermore, we conducted a correlation analysis and found both serum C3 and C1q levels are negatively related to latency to fall in the rotarod test (Fig. 4c, d). In the substantia nigra or the striatum, two critical brain regions with PD, the loss of TH protein in PFFs group was accompanied by an increase in complement C3, while exercise training reduced the level of complement C3 (Fig. 4e, f). The microglial expression of complement C1q in the striatum exhibited a similar pattern. More C1q expression in microglia were observed after PFFs treatment, and exercise training significantly reversed the change (Fig. 4g). Furthermore, we performed Elisa to detect dopamine level in the striatum, and found that decreased in PFFs group compared to Sham group, and covered after treadmill training (Fig. 4h). Therefore, we figured out that complement pathway was definitely over-activated in PFFs mouse, resulting in the increased microglial phagocytosis of synapse. And exercise training on PD mice may downregulate complement pathway to improve the synapse survival and motor function. In agreement with that hypothesis, the striatal level of CD55, known as a negative regulator of complement system, was markedly downregulated in response to PFFs treatment and rebounded after exercise (Fig. 4i). All of those indicated CD55 might be the key molecule participating in PD progress as well as exercise-induced protective effects, probably via the regulation of complement and microglial synaptic phagocytosis.

Fig. 4
figure 4

Exercise training suppressed complement C3 and C1q levels and inhibited complement-induced phagocytic function of microglia. Level of serum C3 (a) (Sham group n = 6, PFFs group n = 5, Exer group n = 7) and C1q (b) (Sham group n = 5, PFFs group n = 7, Exer group n = 7) concentration measured by Elisa. The correlation of latency of rotarod test and serum C3 (c) and C1q (d) concentration. e Western blot showing the expression of TH protein and complement C3α in the substantia nigra. (n = 6 per group) f Western blot showing the expression of TH protein and complement C3α in the striatum. (n = 6 per group) g Immunofluorescence results and relative quantitative analysis of C1q and microglia co-localization in the striatum. Scale bar = 20 μm. (n = 20 per group from 4 mice each group) h Level of DA concentration measured by Elisa in the striatum. i Western blot showing the expression of CD55 protein in the striatum. (n = 6 per group) All data are presented as mean ± SD. Statistical analysis was performed using one-way ANOVA

Overexpression of CD55 effectively improves PD motor skills and reduces synaptic phagocytosis by microglia

Based on the results above, we constructed AAV virus overexpressing CD55 to further explore its role in affecting the progress of PD. The genetic modulation of CD55 were introduced on mice before or at the same time of PFFs treatment, respectively.

Under the simultaneous treatment condition, we conducted behavioral experiments, including pole and rotarod tests, on three groups of mice. The results showed significant behavioral differences between Sham group and PFFs or CD55-OE group. However, there was no statistical difference in behavior between the CD55-OE and PFFs group (Fig. 5a, b), indicating that simultaneous administration of PFFs to the substantia nigra and CD55-OE-AAV into the striatum did not improve the motor deficits in the mouse model of PD. Results of quantitative western blot analysis revealed that striatal CD55 level was downregulated in PFFs group compared to Sham group, while the CD55-OE-AAV injection elevated that expression (Fig. 5c). This validated the efficacy of our virus infection as well as the potential impact of CD55 on PD. The complement C3α expression level in the striatum exhibited a decreasing trend in response to CD55-OE-AAV treatment. When analyzing the core pathological areas of PD, the substantia nigra and striatum, we found that CD55-OE-AAV could reduce the loss of TH protein (Fig. 5d). Also, we investigated the synapses after CD55-OE-AAV treatment. Western blot results indicated a significant reduction in PSD95 in PFFs group compared to Sham group. CD55-OE-AAV introduction partly attenuated the loss of PSD95, but it didn’t show a statistical significance. While, results of Golgi staining revealed dendritic spine loss in PFFs group, which was rescued by CD55 overexpression (Fig. 5e). This was further confirmed by immunofluorescence staining, where staining for iba1, CD68 and PSD95 in the striatum revealed a significant inhibition in microglial activation and a decrease in microglial phagocytosis of synapses in the CD55-OE group compared to PFFs group (Fig. 5f). Based on the result above, we found that increased CD55 expression can to some extent attenuate PD-related pathological changes.

Fig. 5
figure 5

Behavioral and pathological evaluation of AAV treatment experiment. a The latency time of the rotarod test and pole test in Sham, PFFs, and CD55-OE groups over time and at day21. (Sham group n = 8, PFFs group n = 7, CD55-OE group n = 11) b Time of the pole test in the three groups over time and at day21. (Sham group n = 9, PFFs group n = 8, CD55-OE group n = 11) c Results of western blot for the expression of CD55 and complement C3α in the striatum. (n = 6 per group) d Results of western blot for the expression of TH protein in the substantia nigra and striatum. (n = 6 per group) e Results of western blot for the expression of PSD95 protein in the striatum and spine density count in the striatum. (n = 6 per group) (n = 18 per group from 3 mice each group) f Changes in microglial states and synaptic engulfment function in the striatum among Sham, PFFs, and CD55-OE groups. Scale bar = 10 μm. (n = 15 per group from 3 mice each group) (n = 3per group) All data are presented as mean ± SD. Statistical analysis was performed using one-way ANOVA

Further, we conducted the pre-treatment experiment in which the virus was injected into mice two weeks before PFFs modeling to observe the impact of high CD55 expression on the progression of PD. Behavioral results from the pre-treatment experiment showed a significant difference in pole and rotarod test, with the CD55-OE group exhibiting prolonged retention time and shortened climbing time compared to PFFs group (Fig. 6a, b). Next, we assessed CD55 and complement C3α levels in the striatum. A significant upregulation of CD55 in the CD55-OE group compared to both the Sham and PFFs groups was detected. The CD55-OE group showed a significant downregulation of complement C3α level compared to PFFs group, confirming that the overexpression of CD55 indeed inhibited complement C3α (Fig. 6c). Quantitative analysis of TH protein in the substantia nigra and striatum indicated that PFFs-treated mice exhibited TH protein loss as expected, while the CD55-OE-AAV pretreatment delayed this process (Fig. 6d). Western blot analysis disclosed that striatal PSD95 was markedly suppressed after PPFs injection, while CD55-OE-AAV pre-treatment attenuated the loss of PSD95.Similarly, results of Golgi staining also revealed dendritic spine loss in the stratum in PFFs group, which was reversed by CD55-OE-AAV pre-treatment (Fig. 6e). Meanwhile, immunofluorescence staining for CD68 and iba1 revealed a reduction in microglial activation in the CD55-OE-AAV group compared to PFFs group (Fig. 6f), suggesting that CD55-OE-AAV could restrict microglial phagocytic activity towards synapses in the striatum. These results indicate that overexpressing CD55 in the striatum can improve behavioral and pathological aspects of PFFs-induced PD to some extent, probably via tuning down excessive microglial phagocytosis of synapses.

Fig. 6
figure 6

Behavioral and pathological evaluation of AAV pre-treatment experiment. a The latency time of the rotarod test and pole test in Sham, PFFs, and CD55-OE groups over time and at day21. (Sham group n = 7, PFFs group n = 6, CD55-OE group n = 8) b Time of the pole test in the three groups over time and at day21. (Sham group n = 8, PFFs group n = 6, CD55-OE group n = 7) c Results of western blot for the expression of CD55 and complement C3α in the striatum. (n = 5 per group) d Results of western blot for the expression of TH protein in the substantia nigra and striatum. (n = 5 per group) e Results of western blot for the expression of PSD95 protein in the striatum and spine density count in the striatum. (n = 5 per group) (n = 18 per group from 3 mice each group) f Changes in microglial states and synaptic engulfment function in the striatum among Sham, PFFs, and CD55-OE groups. Scale bar = 10 μm. (n = 15 per group from 3 mice each group) (n = 3per group) All data are presented as mean ± SD. Statistical analysis was performed using one-way ANOVA

Proteomics comprehensively reveals the mechanism of overexpression of CD55 and explores the similarity between exercise and CD55

To further explore the role of CD55 in PD treatment and investigate the commonalities between CD55-OE-AAV and exercise, we performed proteomics of the striatum from the CD55-OE-AAV pre-treatment experiment. PCA revealed that samples from the same group clustered together, indicating similar protein expression patterns within groups. Additionally, the distinct separation between groups suggests that CD55 altered the protein differences caused by PFFs (Fig. 7a). Based on the PCA results, we identified differentially expressed proteins between pairs of groups. PFFs pretreatment upregulated 401 proteins and downregulated 324 proteins, compared to the Sham group. While CD55-OE group upregulated 190 proteins and downregulated 99 proteins compared to PFFs group (Fig. 7b). Furthermore, we found the “complement activation pathway” was upregulated after PFFs pretreatment. While in CD55-OE group that pathway was downregulated compared to the Sham group (Fig. 7c). To comprehensively investigate how CD55-OE-AAV improves PD, we conducted an intersection analysis of differentially expressed proteins across the groups. First, we analyzed the subset of proteins upregulated in PFFs group but downregulated in CD55-OE group, identifying 32 differentially expressed proteins. GO functional enrichment analysis showed that these 32 proteins were mostly enriched in RNA splicing and metabolism pathways, such as “mRNA processing”, “RNA splicing”, and “nuclear-transcribed mRNA catabolic process, nonsense-mediated decay”. Similarly, we analyzed the proteins downregulated in PFFs group but upregulated in CD55-OE group, identifying 63 differentially expressed proteins. These proteins were primarily associated with dopaminergic neurons and dopamine-related pathways, including “dopamine transport” and “dopamine biosynthetic process” (Fig. 7d). Additionally, we aimed to identify similarities in proteomics between exercise intervention and CD55-OE-AAV pre-treatment. We found that compared to Sham group, the “neuropeptide signaling pathway” was significantly upregulated in both Exer group and CD55-OE group (Fig. 7e), while “synaptic transmission, glutamatergic” was significantly downregulated (Fig. 7f). These results suggest that the shared mechanisms by which exercise and AAV improve PD may involve enhancing neuropeptide signaling in the striatum and inhibiting the excessively upregulated glutamatergic system. Also, the relationship between neuropeptide, glutamate and complement still remains to be explored.

Fig. 7
figure 7

Proteomics analysis of CD55-OE-AAV pre-treatment group revealed the protective effects of CD55 and potential shared pathways between CD55-AAV overexpression and exercise intervention. a PCA analysis of Sham, PFFs and CD55-OE groups. b Volcano plot of DE Proteins between pairs of groups. c Results of “complement activation pathway” from GSEA of three groups. d Different expression module of three groups and relative pathway enrichment. e Exercise and CD55-OE-AAV common up-regulated GSEA pathway. f Exercise and CD55-OE-AAV common down-regulated GSEA pathway

Discussion

Exercise training attenuates motor deficits and pathological features of PD. The potential causes might be related to neuronal survival, synaptic plasticity, neuroinflammation, oxidative stress, neurotrophic factors, etc. [48, 49]. However, the understanding of the specific process mediating the exercise-induced protective effect on PD is still rare. In this study, we found PFFs mice actually exhibited over-activated microglia and synapse loss in the striatum, and exercise training definitely reversed that phenomenon. Furthermore, using RNA-Seq and proteomics, we observed excessive activation of the complement pathway, which is closely related to excessive microglial synaptic phagocytosis in PD mice. Treadmill running on PFFs-treated mice markedly inhibited complement levels, and exercise-induced impacts on motor functions, microglia and synapse were mimicked by overexpressing CD55. Therefore, we proposed that complement may be the key element through which exercise suppresses microglial synapse elimination and slows the PD progress.

As mentioned earlier, PD exhibits abnormalities in synaptic structure and function. Here, we verified these changes through various methods. We detected the decreased expression level of PSD95 in the striatum in PFFs group compared to Sham group, and concurrently observed a significant loss of spines in the striatum through Golgi staining. The reduction of spine density in the striatum is regarded as the consequence of dopamine denervation under the condition of PD, which might further disturb the connectivity of the corticostriatal and thalamostriatal systems [50]. The striatal spine pathology in PFFs mice was ameliorated by treadmill exercise or the overexpression of CD55 in our study. That at least partially pointed out the repair of basal ganglia neurocircuitry.

Induction of α-synuclein into the brain are suggested to trigger neuroinflammation. Striatal CD11b + cells presenting a pro-inflammatory state were detected when AAV9 for α-synuclein was injected into the substantia nigra of mice [51]. A recent study unraveled that chronic expose to PD patient-derived α-synuclein fibrils (FPD) promoted the microglial release of IL-1b, IL-6, TNF-a and glutamate. Moreover, treatment of FPD together with Tumor necrosis factor-α (TNFα) and prostaglandin E2 (PGE2) resulted in a microglial metabolic reprogramming closely linked to M1-type polarized cells [52]. Consistent with these results, we also found microglia in the striatum are more active in PFFs mouse, while exercise can reverse the change. Excessive synaptic pruning by microglia has been proposed as one of the potential causes for many neurodegenerative diseases [33, 53]. In order to discern the correlation between microglia and synapse in our study, we conducted immunofluorescence co-staining. As a result, phagocytosis of synapse by microglia increased after PFFs induction and decreased after exercise training.

Our main discovery is that the complement pathway, negatively regulated after exercise, plays a crucial role in mediating microglial phagocytosis of striatal synapse related to the PD status. From the data of RNA-Seq and proteomics, we observed an upregulation of the complement cascade reaction in PD and a downregulation of it after exercise training. Similarly, we detected elevated levels of complement in both the substantia nigra-striatum region and peripheral blood serum in PD samples through WB and ELISA assays. After exercise, decrease in complement expressions were observed in PD mice. Additionally, the performance in behavior tests was negatively associated with serum concentration of C1q and C3, which also suggests the potential of complement in affecting motor ability. In previous studies, C1q or C3 knockout as well as application of endogenous neuronal complement inhibitor SRPX2 were found to suppress synaptic elimination, indicating that the complement cascade is involved in synapse loss [54, 55]. Specially, C1q and C3 may bind to adjacent weaker synapses and eliminate them through the phagocytosis of activated microglia [56]. Upregulation of complement levels has been observed in various neurodegenerative diseases. In brain samples of Alzheimer’s disease (AD) subjects, C3 transcription level was significantly higher than controls, and was also significantly correlated with cognitive decline [57]. Blocking C1 activation using genetic-editing techniques or by antibodies can alleviate Aβ synaptic toxicity [33]. Multiple sclerosis shows synaptic phagocytosis by microglia and severe synaptic loss, associated with increased complement component C3. In a mouse model of depressive disorder, behavioral impairments were rescued by using C1q neutralizing antibody suppress the C1q/C3-CR3 pathway and reduce abnormal microglial phagocytosis of synapse [58]. Survival and loss of dopaminergic neurons in the substantia nigra are also linked to C3, as reported by Bodea and colleagues, complement C3-deficient mice were resistant to LPS-induced dopaminergic denervation [59]. Complement system are supposed to be regulated by physical training, which might contribute to the exercise-related effects on various diseases. For example, exercise can alleviate neuropathic pain, primarily by reducing the level of complement C3 [60]. In AD, voluntary exercise was suggested to effectively reduce the number of hippocampal microglia, synaptic colocalization with microglia, as well as the C1q/C3 complement system [42, 43]. In a TDP-43-induced ALS model, treadmill exercise prevented the worsening of motor dysfunction, enhanced the release of clusterin to block the complement pathway, and further inhibited complement-mediated microglial synaptic phagocytosis [44]. In our study, we found that complement levels increased in PD mice, within peripheral blood and multiple brain regions, while exercise training inhibited the activation of complement and related excessive microglial phagocytosis of striatal synapses.

Furthermore, we validated the alteration of CD55, an important complement deactivator on PD. We found that in the striatum region of PFFs group, CD55 level were significantly decreased compared to Sham group, but they increased after exercise training. Changes of CD55 in the central nervous system have been rarely researched. An ultra-deep sequencing on brain predicted a variant transcription factor binding sites upstream of the CD55 gene, which contributes to AD pathogenesis by affecting the complement system [61]. Also, multifocal motor neuropathy (MMN) immunopathology is associated with CD55 promotor mutation and lower CD55 expression. CD55 can decrease complement activity, protect the integrity of the blood–brain barrier, and mitigate blast-induced neurotrauma [62]. In primary cultured hypoxic neurons, exogenously added CD55 can reduce cell death and apoptosis [63]. Overexpression of CD55 on cultured IMR32 neuronal cells was also found to protect neurons from complement attack and cell death [64]. To our knowledge, the changes in CD55 in PD have not yet been reported. We herein report that in the striatum of PFFs-treated mice, CD55 level was significantly decreased compared to sham-operated mice, but that expression increased after exercise training. In order to figure out the significance of CD55 in PD, we utilized an AAV-mediated overexpression technique on PFFs-treated mice. The results showed that elevation in CD55 expression can to some extent improve behavioral and pathological aspects of PD, probably via reducing excessive microglial phagocytosis of striatal synapses. CD55 might also couple exercise and microglial function during the intervention on symptoms of Parkinson’s disease.

Previous research has described various mechanisms through which exercise training slow PD progress [25, 65,66,67,68,69]. Here, we proposed that exercise may alleviate motor and pathological impairments of PD by reducing complement levels, thereby mitigating microglia and subsequent phagocytosis. Complement-mediated microglial phagocytosis in the striatum is one of the central alterations in PFF-treated mouse model of PD, giving rise to dopaminergic denervation. Elevation in CD55 level by treadmill training or by genetic modification would antagonize complement activation and bring beneficial effects to PD subjects. This study investigated the interactions among microglia, complement system and synapses during the development and exercise intervention of Parkinson’s disease. CD55 targeting on complement can be considered as the critical molecule of exercise therapy in PD, as well as the potential pharmaceutical target.

Data availability

No datasets were generated or analysed during the current study.

Abbreviations

PD:

Parkinson’s disease

PSD95:

Post-synaptic density-95

AD:

Alzheimer’s disease

References

  1. Dorsey ER, Elbaz A, Nichols E, Abd-Allah F, Abdelalim A, Adsuar JC, et al. Global, regional, and national burden of Parkinson’s disease, 1990–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2018;17(11):939–53.

    Article  Google Scholar 

  2. Olanow CW, Tatton WG. Etiology and pathogenesis of Parkinson’s disease. Annu Rev Neurosci. 1999;22:123–44.

    Article  CAS  PubMed  Google Scholar 

  3. Samii A, Nutt JG, Ransom BR. Parkinson’s disease. Lancet. 2004;363(9423):1783–93.

    Article  CAS  PubMed  Google Scholar 

  4. Hayes MT. Parkinson’s disease and parkinsonism. Am J Med. 2019;132(7):802–7.

    Article  PubMed  Google Scholar 

  5. Lees AJ, Hardy J, Revesz T. Parkinson’s disease. Lancet. 2009;373(9680):2055–66.

    Article  CAS  PubMed  Google Scholar 

  6. Kalia LV, Lang AE. Parkinson’s disease. Lancet. 2015;386(9996):896–912.

    Article  CAS  PubMed  Google Scholar 

  7. Schapira AH, Jenner P. Etiology and pathogenesis of Parkinson’s disease. Mov Disord. 2011;26(6):1049–55.

    Article  PubMed  Google Scholar 

  8. Su X, Federoff HJ. Immune responses in Parkinson’s disease: interplay between central and peripheral immune systems. Biomed Res Int. 2014;2014: 275178.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Prinz M, Jung S, Priller J. Microglia biology: one century of evolving concepts. Cell. 2019;179(2):292–311.

    Article  CAS  PubMed  Google Scholar 

  10. Wolf SA, Boddeke HW, Kettenmann H. Microglia in physiology and disease. Annu Rev Physiol. 2017;79:619–43.

    Article  CAS  PubMed  Google Scholar 

  11. Hickman S, Izzy S, Sen P, Morsett L, El Khoury J. Microglia in neurodegeneration. Nat Neurosci. 2018;21(10):1359–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Subhramanyam CS, Wang C, Hu Q, Dheen ST. Microglia-mediated neuroinflammation in neurodegenerative diseases. Semin Cell Dev Biol. 2019;94:112–20.

    Article  CAS  PubMed  Google Scholar 

  13. Colonna M, Butovsky O. Microglia function in the central nervous system during health and neurodegeneration. Annu Rev Immunol. 2017;35:441–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Witzig VS, Komnig D, Falkenburger BH. Changes in striatal medium spiny neuron morphology resulting from dopamine depletion are reversible. Cells. 2020;9(11):2441.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Janezic S, Threlfell S, Dodson PD, Dowie MJ, Taylor TN, Potgieter D, et al. Deficits in dopaminergic transmission precede neuron loss and dysfunction in a new Parkinson model. Proc Natl Acad Sci U S A. 2013;110(42):E4016–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Delva A, Van Weehaeghe D, Koole M, Van Laere K, Vandenberghe W. Loss of presynaptic terminal integrity in the substantia nigra in early Parkinson’s disease. Mov Disord. 2020;35(11):1977–86.

    Article  CAS  PubMed  Google Scholar 

  17. Fazio P, Svenningsson P, Forsberg A, Jönsson EG, Amini N, Nakao R, et al. Quantitative analysis of 1⁸F-(E)-N-(3-Iodoprop-2-Enyl)-2β-carbofluoroethoxy-3β-(4’-Methyl-Phenyl) nortropane binding to the dopamine transporter in Parkinson disease. J Nucl Med. 2015;56(5):714–20.

    Article  CAS  PubMed  Google Scholar 

  18. Bloem BR, Okun MS, Klein C. Parkinson’s disease. Lancet. 2021;397(10291):2284–303.

    Article  CAS  PubMed  Google Scholar 

  19. Hill DR, Huters AD, Towne TB, Reddy RE, Fogle JL, Voight EA, et al. Parkinson’s disease: advances in treatment and the syntheses of various classes of pharmaceutical drug substances. Chem Rev. 2023;123(23):13693–712.

    Article  CAS  PubMed  Google Scholar 

  20. Shen X, Wong-Yu IS, Mak MK. Effects of exercise on falls, balance, and gait ability in Parkinson’s disease: a meta-analysis. Neurorehabil Neural Repair. 2016;30(6):512–27.

    Article  PubMed  Google Scholar 

  21. Mak MK, Wong-Yu IS, Shen X, Chung CL. Long-term effects of exercise and physical therapy in people with Parkinson disease. Nat Rev Neurol. 2017;13(11):689–703.

    Article  PubMed  Google Scholar 

  22. Lorenzo-García P, Cavero-Redondo I, NúñezdeArenas-Arroyo S, Guzmán-Pavón MJ, Priego-Jiménez S, Álvarez-Bueno C. Effects of physical exercise interventions on balance, postural stability and general mobility in Parkinson’s disease: a network meta-analysis. J Rehabil Med. 2024;56:jrm10329.

    Article  PubMed  Google Scholar 

  23. Morris ME, Menz HB, McGinley JL, Watts JJ, Huxham FE, Murphy AT, et al. A randomized controlled trial to reduce falls in people with Parkinson’s disease. Neurorehabil Neural Repair. 2015;29(8):777–85.

    Article  PubMed  Google Scholar 

  24. Zhang T-Y, Hu Y, Nie Z-Y, Jin R-X, Chen F, Guan Q, et al. Effects of Tai Chi and multimodal exercise training on movement and balance function in mild to moderate idiopathic Parkinson disease. Am J Phys Med Rehabil. 2015;94(10S):921.

    Article  PubMed  Google Scholar 

  25. Lau YS, Patki G, Das-Panja K, Le WD, Ahmad SO. Neuroprotective effects and mechanisms of exercise in a chronic mouse model of Parkinson’s disease with moderate neurodegeneration. Eur J Neurosci. 2011;33(7):1264–74.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Petzinger GM, Walsh JP, Akopian G, Hogg E, Abernathy A, Arevalo P, et al. Effects of treadmill exercise on dopaminergic transmission in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-lesioned mouse model of basal ganglia injury. J Neurosci. 2007;27(20):5291–300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Tajiri N, Yasuhara T, Shingo T, Kondo A, Yuan W, Kadota T, et al. Exercise exerts neuroprotective effects on Parkinson’s disease model of rats. Brain Res. 2010;1310:200–7.

    Article  CAS  PubMed  Google Scholar 

  28. Smith BA, Goldberg NRS, Meshul CK. Effects of treadmill exercise on behavioral recovery and neural changes in the substantia nigra and striatum of the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-lesioned mouse. Brain Res. 2011;1386:70–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sung YH, Kim SC, Hong HP, Park CY, Shin MS, Kim CJ, et al. Treadmill exercise ameliorates dopaminergic neuronal loss through suppressing microglial activation in Parkinson’s disease mice. Life Sci. 2012;91(25–26):1309–16.

    Article  CAS  PubMed  Google Scholar 

  30. Sconce MD, Churchill MJ, Greene RE, Meshul CK. Intervention with exercise restores motor deficits but not nigrostriatal loss in a progressive MPTP mouse model of Parkinson’s disease. Neuroscience. 2015;299:156–74.

    Article  CAS  PubMed  Google Scholar 

  31. Tong W, Zhang K, Yao H, Li L, Hu Y, Zhang J, et al. Transcriptional profiling reveals brain region-specific gene networks regulated in exercise in a mouse model of Parkinson’s disease. Front Aging Neurosci. 2022;14: 891644.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Shin MS, Jeong HY, An DI, Lee HY, Sung YH. Treadmill exercise facilitates synaptic plasticity on dopaminergic neurons and fibers in the mouse model with Parkinson’s disease. Neurosci Lett. 2016;621:28–33.

    Article  CAS  PubMed  Google Scholar 

  33. Hong S, Beja-Glasser VF, Nfonoyim BM, Frouin A, Li S, Ramakrishnan S, et al. Complement and microglia mediate early synapse loss in Alzheimer mouse models. Science. 2016;352(6286):712–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Ding X, Wang J, Huang M, Chen Z, Liu J, Zhang Q, et al. Loss of microglial SIRPα promotes synaptic pruning in preclinical models of neurodegeneration. Nat Commun. 2021;12(1):2030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Wake H, Moorhouse AJ, Jinno S, Kohsaka S, Nabekura J. Resting microglia directly monitor the functional state of synapses in vivo and determine the fate of ischemic terminals. J Neurosci. 2009;29(13):3974–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Alawieh AM, Langley EF, Feng W, Spiotta AM, Tomlinson S. Complement-dependent synaptic uptake and cognitive decline after stroke and reperfusion therapy. J Neurosci. 2020;40(20):4042–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Michailidou I, Willems JGP, Kooi E-J, van Eden C, Gold SM, Geurts JJG, et al. Complement C1q–C3–associated synaptic changes in multiple sclerosis hippocampus. Ann Neurol. 2015;77(6):1007–26.

    Article  CAS  PubMed  Google Scholar 

  38. Barcia C, Ros CM, Annese V, Gómez A, Ros-Bernal F, Aguado-Llera D, et al. Erratum: 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. 2012;3(8):e379–e379.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Barcia C, Ros CM, Annese V, Carrillo-de Sauvage MA, Ros-Bernal F, Gómez A, et al. ROCK/Cdc42-mediated microglial motility and gliapse formation lead to phagocytosis of degenerating dopaminergic neurons in vivo. Sci Rep. 2012;2(1):809.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Marinova-Mutafchieva L, Sadeghian M, Broom L, Davis JB, Medhurst AD, Dexter DT. Relationship between microglial activation and dopaminergic neuronal loss in the substantia nigra: a time course study in a 6-hydroxydopamine model of Parkinson’s disease. J Neurochem. 2009;110(3):966–75.

    Article  CAS  PubMed  Google Scholar 

  41. Virgone-Carlotta A, Uhlrich J, Akram MN, Ressnikoff D, Chrétien F, Domenget C, et al. Mapping and kinetics of microglia/neuron cell-to-cell contacts in the 6-OHDA murine model of Parkinson’s disease. Glia. 2013;61(10):1645–58.

    Article  PubMed  Google Scholar 

  42. Wang Y-Y, Zhou Y-N, Jiang L, Wang S, Zhu L, Zhang S-S, et al. Long-term voluntary exercise inhibited AGE/RAGE and microglial activation and reduced the loss of dendritic spines in the hippocampi of APP/PS1 transgenic mice. Exp Neurol. 2023;363: 114371.

    Article  CAS  PubMed  Google Scholar 

  43. Yang J, Yuan S, Jian Y, Lei Y, Hu Z, Yang Q, et al. Aerobic exercise regulates GPR81 signal pathway and mediates complement- microglia axis homeostasis on synaptic protection in the early stage of Alzheimer’s disease. Life Sci. 2023;331: 122042.

    Article  CAS  PubMed  Google Scholar 

  44. Wei J-A, Liu L, Song X, Lin B, Cui J, Luo L, et al. Physical exercise modulates the microglial complement pathway in mice to relieve cortical circuitry deficits induced by mutant human TDP-43. Cell Rep. 2023;42(3): 112240.

    Article  CAS  PubMed  Google Scholar 

  45. Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015;43(7): e47.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, et al. clusterProfiler 4.0: a universal enrichment tool for interpreting omics data. Innovation (Camb). 2021;2(3):100141.

    CAS  PubMed  Google Scholar 

  47. Volpicelli-Daley LA, Kirik D, Stoyka LE, Standaert DG, Harms AS. How can rAAV-α-synuclein and the fibril α-synuclein models advance our understanding of Parkinson’s disease? J Neurochem. 2016;139(Suppl 1):131–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Mahalakshmi B, Maurya N, Lee SD, Bharath Kumar V. Possible neuroprotective mechanisms of physical exercise in neurodegeneration. Int J Mol Sci. 2020;21(16):5895.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Feng YS, Yang SD, Tan ZX, Wang MM, Xing Y, Dong F, et al. The benefits and mechanisms of exercise training for Parkinson’s disease. Life Sci. 2020;245: 117345.

    Article  CAS  PubMed  Google Scholar 

  50. Villalba RM, Smith Y. Differential striatal spine pathology in Parkinson’s disease and cocaine addiction: a key role of dopamine? Neuroscience. 2013;251:2–20.

    Article  CAS  PubMed  Google Scholar 

  51. Basurco L, Abellanas MA, Ayerra L, Conde E, Vinueza-Gavilanes R, Luquin E, et al. Microglia and astrocyte activation is region-dependent in the α-synuclein mouse model of Parkinson’s disease. Glia. 2023;71(3):571–87.

    Article  CAS  PubMed  Google Scholar 

  52. Yildirim-Balatan C, Fenyi A, Besnault P, Gomez L, Sepulveda-Diaz JE, Michel PP, et al. Parkinson’s disease-derived α-synuclein assemblies combined with chronic-type inflammatory cues promote a neurotoxic microglial phenotype. J Neuroinflammation. 2024;21(1):54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Cunningham C. Microglia and neurodegeneration: the role of systemic inflammation. Glia. 2013;61(1):71–90.

    Article  PubMed  Google Scholar 

  54. Stevens B, Allen NJ, Vazquez LE, Howell GR, Christopherson KS, Nouri N, et al. The classical complement cascade mediates CNS synapse elimination. Cell. 2007;131(6):1164–78.

    Article  CAS  PubMed  Google Scholar 

  55. Cong Q, Soteros BM, Wollet M, Kim JH, Sia GM. The endogenous neuronal complement inhibitor SRPX2 protects against complement-mediated synapse elimination during development. Nat Neurosci. 2020;23(9):1067–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Fourgeaud L, Boulanger LM. Synapse remodeling, compliments of the complement system. Cell. 2007;131(6):1034–6.

    Article  CAS  PubMed  Google Scholar 

  57. Litvinchuk A, Wan YW, Swartzlander DB, Chen F, Cole A, Propson NE, et al. Complement C3aR inactivation attenuates tau pathology and reverses an immune network deregulated in tauopathy models and Alzheimer’s disease. Neuron. 2018;100(6):1337–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Han Q-Q, Shen S-Y, Liang L-F, Chen X-R, Yu J. Complement C1q/C3-CR3 signaling pathway mediates abnormal microglial phagocytosis of synapses in a mouse model of depression. Brain Behav Immun. 2024;119:454–64.

    Article  CAS  PubMed  Google Scholar 

  59. Bodea LG, Wang Y, Linnartz-Gerlach B, Kopatz J, Sinkkonen L, Musgrove R, et al. Neurodegeneration by activation of the microglial complement-phagosome pathway. J Neurosci. 2014;34(25):8546–56.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Wang C, He H, Gao T, Sun X, Du L, Yang Y, et al. Analgesic effect of exercise on neuropathic pain via regulating the complement component 3 of reactive astrocytes. Anesth Analg. 2024;139:840.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Helgadottir HT, Lundin P, Wallén Arzt E, Lindström AK, Graff C, Eriksson M. Somatic mutation that affects transcription factor binding upstream of CD55 in the temporal cortex of a late-onset Alzheimer disease patient. Hum Mol Genet. 2019;28(16):2675–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Li Y, Chavko M, Slack JL, Liu B, McCarron RM, Ross JD, et al. Protective effects of decay-accelerating factor on blast-induced neurotrauma in rats. Acta Neuropathol Commun. 2013;1:52.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Wang Y, Li Y, Dalle Lucca SL, Simovic M, Tsokos GC, Dalle Lucca JJ. Decay accelerating factor (CD55) protects neuronal cells from chemical hypoxia-induced injury. J Neuroinflammation. 2010;7:24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. van Beek J, van Meurs M, Hart BA, Brok HP, Neal JW, Chatagner A, et al. Decay-accelerating factor (CD55) is expressed by neurons in response to chronic but not acute autoimmune central nervous system inflammation associated with complement activation. J Immunol. 2005;174(4):2353–65.

    Article  PubMed  Google Scholar 

  65. Rezaee Z, Marandi SM, Alaei H, Esfarjani F. Exercise-induced neuroprotection in the 6-hydroxydopamine Parkinson’s disease model. Neurotox Res. 2020;38(4):850–8.

    Article  CAS  PubMed  Google Scholar 

  66. Hauser DN, Hastings TG. Mitochondrial dysfunction and oxidative stress in Parkinson’s disease and monogenic parkinsonism. Neurobiol Dis. 2013;51:35–42.

    Article  CAS  PubMed  Google Scholar 

  67. Forman HJ, Zhang H, Rinna A. Glutathione: overview of its protective roles, measurement, and biosynthesis. Mol Aspects Med. 2009;30(1–2):1–12.

    Article  CAS  PubMed  Google Scholar 

  68. Cooper AJ, Kristal BS. Multiple roles of glutathione in the central nervous system. Biol Chem. 1997;378(8):793–802.

    CAS  PubMed  Google Scholar 

  69. Oliveira LOD, da Silva PIC, Filho RPR, Progênio RCS, de Oliveira VDPS, Silva RC, et al. Prior exercise protects against oxidative stress and motor deficit in a rat model of Parkinson’s disease. Metab Brain Dis. 2020;35(1):175–81.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all members of Professor Siguang Li’s lab.

Funding

This work was funded by the National Natural Science Foundation of China (No.82172545, 82230084), National Key Clinical Specialty Discipline Construction Program of China (Z155080000004), Shanghai Sailing Program (22YF1442100), Shanghai Research Center of Rehabilitation Medicine (Top Priority Research Center of Shanghai) (2023ZZ02027) and Shanghai Blue Cross Brain Hospital Co., Ltd. and Shanghai Tongji University Education Development Foundation.

Author information

Authors and Affiliations

Authors

Contributions

Lingjing Jin and Weifang Tong made significant contributions to the conception and design of the study. Hongkai Yao was responsible for carrying out the main experimental work. Ruoyu Li prepared PFFs. Xuerui Xiang, Yi Yang, Yunxi Liu carried out the exercise training. Wen Cheng assisted with neuropathological studies. Yunping Song, Yunjiao Zhou and Yijing He analyzed the data. The initial draft of the manuscript was written by Hongkai Yao. LingjingJin and Siguang Li revised the manuscript. All authors provided valuable feedback and comments on earlier versions of the manuscript. All authors carefully reviewed and approved the final version of the manuscript.

Corresponding authors

Correspondence to Siguang Li or Lingjing Jin.

Ethics declarations

Ethics approval and consent to participate

The study about animals was approved by the ethics committee of Tongji University (TJBC00321101).

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

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

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yao, H., Tong, W., Song, Y. et al. Exercise training upregulates CD55 to suppress complement-mediated synaptic phagocytosis in Parkinson’s disease. J Neuroinflammation 21, 246 (2024). https://doi.org/10.1186/s12974-024-03234-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12974-024-03234-0

Keywords