Abstract
The accumulation of α-synuclein induces neuronal loss in midbrain nuclei and leads to the disruption of motor circuits, while the pathology of α-synuclein in cortical regions remains elusive. To better characterize cortical synucleinopathy, here we generate a mouse model with the overexpression of human α-synuclein in the primary motor cortex (M1) of mice. A combination of molecular, in vivo recording, and behavioral approaches reveal that cortical expression of human α-synuclein results in the overexcitation of cortical pyramidal neurons (PNs), which are regulated by the decreased inhibitory inputs from parvalbumin-interneurons (PV-INs) to impair complex motor skill learning. Further mechanistic dissections reveal that human α-synuclein aggregation activates ferroptosis, contributing to PV-IN degeneration and motor circuit dysfunction. Taken together, the current study adds more knowledge to the emerging role and pathogenic mechanism of ferroptosis in neurodegenerative diseases.
Similar content being viewed by others
Introduction
The aggregation of α-synuclein is one of the major pathogenic factors in neurodegenerative disorders, including Parkinson’s disease (PD) and Alzheimer’s disease (AD)1,2. Current knowledge has identified synucleinopathy and neuronal loss in subcortical nuclei such as the substantia nigra compacta (SNc) in PD patients3. Such a model of midbrain dopaminergic circuit in PD pathogenesis, however, cannot fully explain motor syndromes. For example, limb rigidity and tremor were believed to be affected by the primary motor cortex (M1)4, which regulates voluntary movement via local excitatory-inhibitory circuits and top-down projections. In human PD patients, cortical regions such as the cingulate cortex5 also presented a higher abundance of α-synuclein oligomers. Besides PD, cortical α-synuclein deposits are also found in human Lewy body disease6 and AD patients7. Therefore, further understanding of cortical synucleinopathy is critical for illustrating the pathology of neurodegenerative diseases.
Different theories have been proposed to explore the molecular mechanisms of neuronal degeneration under PD. The emerging model of lipid peroxidation and ferroptosis8 has drawn research interest, as the potential role of ferroptosis in PD pathology has been reported9,10. In specific, potentiated ferroptosis has been found in both drug-induced11 and transgenic PD models with α-synuclein overexpression12. The induction of ferroptosis may occur with the activation of neuroinflammation13,14, contributing to neuronal death. The activation of microglial cells also resulted in the elevated release of chemokines release, further enhancing microglia migration15. As a consequence, microglia-mediated phagocytosis of synapse and neuronal soma disrupted neural networks16,17. These existing studies, however, mainly analyzed the ferroptosis in midbrain nuclei during PD pathogenesis18, with insufficient focus on cortical regions.
In the current work, we generated a mouse synucleinopathy model by viral-mediated overexpression of the human α-synuclein (SNCA) gene into the M1. The de novo production of α-synuclein aggregates resulted in deficits of complex motor learning, in association with hyperactivation of cortical pyramidal neurons (PNs). Such overexcitation was found to be caused by the downregulation of parvalbumin-interneurons (PV-INs). Further molecular studies showed that following SNCA overexpression, iron overload and disrupted ferroptosis probably caused cortical dysfunctions. In sum, our work revealed one previously unrecognized molecular pathway by which SNCA disrupts circuitry integrity and cortical function.
Results
SNCA overexpression in mouse M1 leads to cortical hyperexcitability and motor learning impairments
We first introduced SNCA-coding gene into the bilateral M1 of adult male mice using an adeno-associated virus (AAV) vector (Fig. 1a). Three weeks later, these mice presented significantly elevated levels of α-synuclein in M1 extracts (P < 0.001; Fig. 1b, c, Supplementary Fig. 1). In specific, phosphorylated α-synuclein proteins in neurons (NeuN+) can be found across different layers of M1 (Fig. 1d) as well as in axonal fibers from M1 to the dorsal striatum (DS) (Supplementary Fig. 2a). However, the postsynaptic neurons in DS did not exhibit transmission of protein aggregations, nor did other cortical regions (Supplementary Fig. 2a), thus suggesting M1-specificity of SNCA overexpression. In behavioral phenotyping, SNCA mice displayed normal anxiety levels or novel object recognition function (P > 0.05; Supplementary Fig. 2b–d). When examining the motor functions, SNCA mice had unchanged general motor ability (P > 0.05; Fig. 1e), whilst their abilities to acquire complex motor skills were disrupted. Of note, SNCA mice presented similar latency on the accelerating rotarod compared to control groups during the initial trial, whilst their performance did not improve with repeated training sessions (P < 0.0001; Fig. 1f). In a similar manner, SNCA mice had longer duration in accomplishing vertical climbing or horizontal beam walking task (P < 0.05; Fig. 1g, h). These data suggested that SNCA accumulation in the motor cortex resulted in dysregulated motor learning functions.
Cortical neuronal activity is closely associated with motor learning functions, and hyperexcitability of the cortex has been reported in neurodegenerative diseases19,20. To investigate the cortical neural network homeostasis, we employed ex vivo whole-cell patch-clamp recording on acute M1 slices. SNCA mice had potentiated miniature excitatory postsynaptic currents (mEPSCs; Fig. 1i) of layer 5 pyramidal neurons (L5 PNs). In specific, both the amplitude and frequency of mEPSCs were elevated (P < 0.05; Fig. 1j, k). Such observations of higher synaptic transmission parallelled with the potentiated membrane excitability in SNCA group, including more frequent action potential spikes (P < 0.05; Fig. 1l), plus higher values of afterhyperpolarization (AHP), Rheobase and resting membrane potential (RMP) (P < 0.01, Fig. 1m–o).
The disruption of synaptic functions indicated the dysregulation of the cortico-striatal pathway, which was closely associated with complex motor skills21. To further support this hypothesis, we performed in vivo calcium imaging of M1 by co-expressing genetically encoded calcium indicator GCaMP6s along with SNCA (Fig. 2a). Three weeks later, the calcium transient of L5 PNs was recorded on awake, head-fixed mice (Fig. 2b). SNCA mice displayed higher total integrated calcium activities (P < 0.001; Fig. 2c), which can be attributed to the higher frequency of calcium spikes (P < 0.001; Fig. 2d). To investigate the downstream projecting region of M1, we utilized the in vivo fiber photometry to record the calcium activity of cortical PN fibers within the dorsal striatum (DS) while the mouse was trained on the rotarod (Fig. 2e, f). A major calcium peak occurred when the animal fell down from the rod (Fig. 2g). The strength of this calcium episode, however, was remarkably increased in the SNCA group (P < 0.05; Fig. 2g–i). These results paralleled the hyperexcitability of cortical PNs (Fig. 2b–d) and suggested that cortical overexpression of SNCA resulted in motor learning deficits, which was correlated with hyperexcitation of the cortico-striatal pathway.
Disrupted cortical excitatory-inhibitory balance underlies SNCA-induced motor learning deficits
The excitatory-inhibitory (E/I) balance of the cortical neural network is critical for maintaining normal functions. We thus tested the role of cortical E/I balance in motor learning deficits in SNCA-overexpressed mice. The first piece of evidence came from expressing the designer receptors exclusively activated by designer drugs (DREADD) hM4Di along with AAV-SNCA into mouse M1 (Supplementary Fig. 3a). Chronic infusion of specific ligand clozapine-N-oxide (CNO) remarkably improved motor learning functions (P < 0.01; Supplementary Fig. 3b–d), suggesting the participation of cortical hyperexcitation underlying motor learning deficits of SNCA model. To dissect the circuitry mechanism, local PV-INs were further studied as they innervate cortical PNs to exert indispensable functions in motor learning22. Via immunofluorescent staining, we identified remarkably decreased expression of PV-INs across different M1 layers (P < 0.001; Fig. 3a, b), plus the loss of perineuronal nets (PNNs) which coat PV-INs using wisteria floribunda agglutinin (WFA) as the marker (P < 0.001; Fig. 3c). In addition, we found that only PV-INs showed decreased density under SNCA, but not for vasoactive intestinal peptide (VIP) or somatostatin (SST) interneurons (Supplementary Fig. 4), supporting the subtype-specific effect. To recapitulate such cellular pathology, Cre-dependent Caspase 3 was expressed into M1 of PV-IRES-Cre mice (Supplementary Fig. 5a), leading to decreased density of PV-INs (P < 0.001; Supplementary Fig. 5b, c) and motor learning deficits (P < 0.05; Supplementary Fig. 5d–f) as those in SNCA-overexpressed mice. These data supported the correlation between motor dysfunction and PV-IN loss under synuclein overexpression.
Subsequently, the calcium activity of PV-INs was recorded by in vivo 2-photon microscopy (Fig. 3d). SNCA mice displayed a decreased frequency of calcium transients of PV-INs, resulting in lower total calcium activities (Fig. 3e–g). The depressed inhibitory transmission might contribute to the overexcitation of L5 PNs, which receive GABAergic innervations from PV-INs. To provide the direct causal relationship, we then used a chemogenetic approach to selectively activate cortical PV-INs by expressing excitatory receptor hM3Dq into PV-INs along with AAV-SNCA (Fig. 3h). Chronic CNO infusion significantly elevated cortical neuronal activities (P < 0.001; Fig. 3i, j). Furthermore, behavioral phenotyping showed remarkable improvements in the accelerating rotarod test (P < 0.0001; Fig. 3k), as well as the vertical pole climbing (P < 0.0001; Fig. 3l) or horizontal beam walking task (P < 0.0001; Fig. 3m). Taken together, α-synuclein impaired cortical E/I balance via suppressing PV-INs, leading to motor learning deficits.
Microglia activation impairs cortical circuits to disrupt motor learning
Under SNCA overexpression, microglia is the primary cell type for extracellular protein aggregates clearance23, during which both morphological and molecular features of microglia are prominently altered. We thus used ionized calcium-binding adapter molecule 1 (Iba1) as the cellular marker to study the morphology of cortical microglia (Fig. 4a). In SNCA-overexpressing mice, microglia presented elevated cell density and phagocytosis toward SNCA-expressing cells (P = 0.011; Fig. 4a, b), in association with higher Iba1 and CD68 expressions (P = 0.037; Fig. 4c, d, Supplementary Fig. 6). The morphometric study of individual microglial cells further revealed enlarged soma size (P < 0.001; Fig. 4e, f) and highly branched processes (P < 0.001; Fig. 4g), suggesting the activation of cortical microglia upon SNCA overexpression.
To validate the role of microglia activation in cortical dysfunction, we employed PLX5622, one potent inhibitor of colony-stimulating factor 1 receptor (CSFR1), to suppress the microglial population24. By feeding mice with PLX5622-contained food chows (Fig. 4h), we successfully eliminated the majority of cortical microglial cells (P < 0.001; Fig. 4i, j). Of note, the number of PV-INs was also recovered by microglial ablation (P < 0.001; Fig. 4i, k). Further motor learning tasks revealed the remarkable improvement of motor learning deficits of SNCA mice after microglial inhibition (P < 0.01; Fig. 4l–n). These data highlighted the role of activated microglia in synucleinopathy-related motor learning deficits.
Disruption of ferroptosis mediates cortical PV-INs loss and motor learning dysfunctions
Although we have observed the contribution of microglial activation in cortical deficits under SNCA overexpression, the detailed molecular mechanism remained unclear. We thus employed a transcriptomic study of mouse M1 extracts to investigate the differentially regulated genes (DEGs). Bioinformatics analysis identified a total of 157 up-regulated and 115 down-regulated genes in SNCA mice (Supplementary Fig. 7). Using the Kyoto Encyclopedia of Genes and Genomes (KEGG) database, the ferroptosis pathway was found to be prominently affected (Fig. 5a). Previous knowledge recognized ferroptosis to be involved in the molecular pathology of neurodegenerative diseases8,25. The quantification of gene transcript expression level in M1 found upregulation of ferroptosis-related genes such as Phospholipase A2 group VI (Pla2g6), cystine/glutamate antiporter solute carrier family 7 member 11 (Slc7a11), prostaglandin-endoperoxide synthase 2 (Ptgs2) and transferrin receptor (Tfrc) (P < 0.05; Fig. 5b), supporting the activation of ferroptosis in SNCA group.
To substantiate the role of ferroptosis in SNCA-induced cortical pathology, the potent inhibitor of ferroptosis, ferrostatin-1 (Fer-1)26 was infused into SNCA model mice (Fig. 5c). Fourteen-day drug treatment remarkably recovered normal PV population density in M1 (P = 0.003; Fig. 5d, e) and rescued impairments of complex motor skill learning (P < 0.0001; Fig. 5f–h). As Fer-1 mainly targets the excess free oxidative radicals in the presence of iron26, an alternative approach was performed via short hairpin RNA (shRNA)-mediated silencing of Tfrc gene expression selectively in PV-INs to manipulate the intracellular trafficking of iron (Fig. 5i, j). Morphometric studies revealed that PV-IN population was increased after iron transport was blocked (P < 0.001; Fig. 5k–l). Moreover, the skill of complex motor functions was remarkably improved with the deprivation of Tfrc gene in PV-INs (P < 0.001; Fig. 5m–o) as those of Fer-1 treated animals. Due to the unaffected microglial cell population in DS after SNCA overexpression (Supplementary Fig. 8a), and similar microglial density under either Fer-1 or shRNA (Tfrc) treatment (Supplementary Fig. 8b, c), it is likely that SNCA overexpression in M1 only leads to local neuroinflammatory response, and ferroptosis acts at the downstream of neuroinflammation to confer cellular toxicity to PV-INs. Taken together, our molecular, physiological, and behavioral data supported the role of ferroptosis in mediating PV-INs depopulation and cortical dysfunction, under local overexpression of SNCA.
Discussion
The current study revealed the previously unrecognized role of disrupted E/I balance in cortical regions following α-synuclein aggregation. Cortical inhibitory neurons are critical for maintaining normal neural functions, and the dysregulation of INs may result in neurological disorders. Of note, our work suggested potentially suppressed presynaptic inhibitory inputs towards PNs, leading to the overexcitation of cortico-striatal pathway and motor learning deficits. These results thus illustrated the changes of cortical circuits under SNCA overexpression, providing the neural mechanism for motor learning deficits. As further molecular evidence, iron overload and ferroptosis are believed to be involved in cortical E/I imbalance.
The region- and cell-type-specific expression of α-synuclein is one critical issue for illustrating the circuitry mechanism of PD. Current studies mainly focused on dopaminergic neurons whilst leaving the GABAergic cells largely unattended. Some comparative studies in mouse tissues found the co-existence between α-synuclein and vesicular glutamate transporter 1 (vGluT1) but not in cortical GABAergic neurons under normal conditions27,28. The exogenous supply of SNCA fibrils induced intracellular aggregation formation in inhibitory neurons, which did not present the formation of Lewy body structure as those in excitatory neurons28. These results highlighted the potentially differentially regulated pathway for synucleinopathy in inhibitory neurons. In our model, SNCA was expressed in both glutamatergic and GABAergic cells, the latter of which presented cell loss in the first place. It is worth noticing that the downregulation of PV-INs has been found across different neurological disorders, including autism29, PD30 and AD31. Although no direct evidence has been obtained to demonstrate the neuronal loss of PV-INs in our work, and reports from PD patients did not reveal the neuronal degeneration in cortical regions32, a few models have been discussed to explain the vulnerability of GABAergic neurons under proteinopathy, including the relatively higher energy requirement33, or the preference of microglial phagocytosis targeting the extracellular matrix of PV-INs34.
GABAergic INs play indispensable roles in maintaining the structural and functional integrity of cortical circuits. PV-INs are the major subtype of inhibitory cells forming peri-somatic innervations of PNs to exert rapid and powerful inhibition35, providing critical inputs for maintaining the E/I balance of the cortical network36. From functional perspectives, cortical PV-INs have been found to regulate social behaviors37, sensory integration38,39 and spatial working memory40. The disruption of normal PV-IN function resulted in mental disorders such as schizophrenia41. In motor cortex, PV-INs work with other INs to shape task-dependent ramping signals preceding motor movements42 and help to regulate the plasticity of other inhibitory boutons during rotarod motor learning43. Our data illustrated the participation of PV-INs in tuning the calcium activity of PNs, as cell-specific ablation and chemogenetic excitation effectively recapitulated or reversed SNCA-induced hyperexcitability of local L5 PNs, respectively. These results highlighted the importance of cortical E/I balance in executing normal motor learning tasks. Since inhibitory transmission can be regulated by noninvasive neuromodulation approaches such as transcranial magnetic stimulation (TMS)44, strategies might be developed to target cortical inhibitory network for the alleviation of Parkinsonism.
In searching for the molecular mechanism governing the gradual loss of PV-INs upon SNCA overexpression, we proposed that ferroptosis signaling might be related to PV-INs loss. In the current work, we found prominent changes in ferroptosis-related genes, including Pla2g6, Slc7a11, Ptgs2, and Tfrc. To further demonstrate the participation of this ferroptosis pathway, the pharmaceutical inhibitor Fer-1 was employed to rescue PV-IN degeneration and motor learning deficits. Moreover, the involvement of iron homeostasis was supported via depressing Tfrc gene expression, in order to inhibit cellular iron uptake for relieving iron overload. The excess iron content may further disrupt the catabolism of lipid peroxide, leading to the activation of the ferroptosis cascade in cortical neurons, as supported by a recent study using the spinal injury model45. Since ferroptosis in the striatum might be related to intracerebral hemorrhage46 and PD47, the targeting of ferroptosis may provide a potential strategy for PD treatment.
The activation of the ferroptosis pathway is closely related to PD pathogenesis. Amounts of studies have supported the participation of iron overload and the disruption of critical regulators of the ferroptosis pathway during PD pathogenesis. For example, in a homozygous PD mouse model (α-Synuclein A53T mutant), dopaminergic neuronal death is caused by iron overload during the early disease stage48. The iron overload also induced the transcriptomic dynamic shift of PD patients’ brains toward ferroptosis, as supported by the human-induced pluripotent stem cell-derived microglial cell culture assay49. In a drug-induced PD model, iron accumulation and ferroptosis were also reported, in addition to the rescued phenotype by Fer-1 treatment50 similar to our work. Moreover, a recent study using dopaminergic cell culture model further supported the protective role of Ferritin against neurotoxicity51. These in vitro and in vivo data all agreed on the relationship between iron overload in conjunction with ferroptosis and PD susceptibility and proposed the potential value of targeting the ferroptosis pathway for neuroprotection. Although these abovementioned works mainly studied midbrain dopaminergic neurons, here we presented amounts of data showing the critical role of this pathway in cortical neuronal functions. More importantly, by specifically modulating Tfrc gene expression in PV-INs, we reported the necessary role of ferroptosis in cortical INs under synucleinopathy.
In sum, SNCA disrupted E/I balance of the motor cortex, mainly due to the downregulation of PV-INs, which leads to the hyperexcitation of PNs and motor learning deficits. The loss of PV-INs was potentially related to microglial activation and elevated ferroptosis. Our work provides more mechanisms for neuropathology under synucleinopathy and implies potential molecular and circuitry targets for future interventional strategies.
Methods
Experimental animals
C57BL/6J mice (5 weeks old, male) were purchased from Guangdong Medical Laboratory Animal Center. PV-Cre mice (5 weeks old, male) were obtained from Jackson Laboratory and were bred by Shanghai Model Organisms Center, Inc. Unless otherwise specified, all mice were group housed (4–5 individuals per standard cage) with food and water ad libitum. The animal facility was kept with a 12 h light/dark cycle (light 8:00–20:00, dark 20:00–8:00), and the temperature was maintained at 24 ± 2 °C, with the relative humidity at 50 ± 10%. All animal experimental protocols have been approved by the Ethics Committee of Experimental Animals of Jinan University in accordance with Institutional Animal Care and Use Committee guidelines for animal research.
For PLX5622 treatment, 0.12% PLX5622 was prepared in standard AIN-76A rodent diet, which was provided by Dyets Inc. Mice were provided PLX5622-containing chow starting at 7 days after virus infection. For Fer-1 treatment, mice received intraperitoneal injection of Fer-1 (MedChemExpress, US) at 5 mg/kg for 14 consecutive days.
Stereotaxic injection and two-photon imaging
Mice were anesthetized with 1.25% Avertin. After local sterilization and incision of head skins, the primary motor cortex (M1, AP: +0.5 mm anterior to the Bregma; ML: ±1.2 mm lateral; and DV: −0.6 mm depth) was located under a stereotaxic instrument (RWD Life Science, China). After drilling a hole with the high-speed microdrill (WPI, USA), 100 nl of AAV serotype 2/9 (viral titer, 5 × 1012 genome copies per ml, see Supplementary Table 1) was injected using a glass micropipette connected to an ultra-micro injection pump (Nanoliter 2020, WPI, USA). The glass electrode was retained within the brain tissue for 8 min before retraction. After the virus injection, the scalp was sutured, and the mouse was placed back in the cage for 1-week recovery. Analgesics (buprenorphine at 0.1 mg/kg) was used when necessary.
One day before the imaging session, a transcranial imaging window was prepared. In brief, the head skin and the skull over the primary motor cortex were carefully removed by a microdrill. The window was covered by a glass coverslip using Tissue Adhesive (3M, USA). The imaging session was recorded at 2 Hz using a 920 nm excitation laser with a water-immersed objective (20×; ZEISS, Germany) under an LSM780 two-photon microscope (ZEISS, Germany).
Acquired images were analyzed by ImageJ (NIH, US). The fluorescent value was quantified from the designed region of identifiable soma. The △F/F0 was calculated as (F − F0)/F0, where the F0 was the minimum of F values as the basal level. A calcium spike was defined when the △F/F0 was higher than three-fold standard deviations (SDs).
Behavioral assays
Open field
The test apparatus contained a transparent field at 50 × 50 × 50 cm, with an arbitrary defined central area (25 × 25 cm). During the 5-min single test session, the Ethovison XT software was used to track and analyze the mouse movement trajectory. The anxiety level was evaluated using the time in the central area.
Novel object recognition task
The experiment procedure followed our previous works52. In brief, the animal’s preference toward a novel object was evaluated by the exploration time during 10 min test session, which was carried at 15 min or 24 h after the initial learning session (10 min, including 2 identical objects). The Ethovison XT software was used to record and analyze the movement trajectories.
Accelerating rotarod test
The Rotarod system (UGO Basile, Italy) was set at an initial minimum speed of 2 rpm to reach the maximum speed of 80 rpm within 5 min. The whole test was performed in 3 consecutive days, and on each day, there were 3 test trials. Before the formal test, an acclimation session was performed using 2 rpm velocity for 1 min. The time the mouse fell down from the rod was recorded and was averaged across 3 trials.
Vertical climbing test
The experiment apparatus consists of a vertically placed rod (70 cm height, 1 cm diameter). At the beginning of each session, the mouse was placed on the top and was allowed to climb down to the bottom. There were 3 trials on each day, with 5-min interval between each trial. The whole test session lasted for 3 consecutive days.
Horizontal beam walking test
The experiment was performed using a wooden plate with 3 cm width and 70 cm length, with an escape box at one end of the beam. The whole apparatus was placed 100 cm above the ground. The mouse was placed at the end of the beam, and the time it took to reach the escape box was recorded. A total of 3 test trials were performed each day, with 5-min intervals. The whole test session lasted for 3 consecutive days.
Immunofluorescent staining
Mice were treated with 1.25% Avertin for anesthesia, followed by cardiac perfusion using 0.9% saline and 4% paraformaldehyde solution (PFA). The whole brain tissue was carefully removed and was fixed overnight using 4% PFA. The brain tissue was subsequently immersed in a 30% sucrose solution for 1–2 days of dehydration. The whole brain tissue was then cut into 40 μm using a sliding microtome (Lecia, Germany). After PBS washing and blocking, the brain slices were incubated with primary antibody at 4 °C for 48 h, followed by secondary antibody (see Supplementary Table 2 for antibody information). Images were captured with a confocal microscope (ZEISS, Germany).
Western blotting
Tissue lysates were extracted from the primary motor cortex, using the RIPA buffer (Thermo Scientific, US) containing protease and phosphatase inhibitors. After quantification using Pierce Rapid Gold BCA kit (Thermo Scientific, US), 25 μg protein samples were electrophoresed on 8% SDS-polyacrylamide gel and were transferred to the 0.2 μm PVDF membranes under 280 mA for 120 min. The membrane was washed by the TBST and blocked by 5% BSA and was incubated with primary antibody at 4 °C for 8 h. The membrane was then washed and incubated in a secondary antibody for 2 h (Supplementary Table 2). Images were captured by a protein imaging system (Bio-Rad, US) and the integrated gray values were calculated by ImageJ (NIH, US).
RT-qPCR
Total RNA was extracted from the primary motor cortex with Trizol (Invitrogen, US) reagent. Total RNA was reversely transcribed using Prescript RT reagent Kit with gDNA Eraser (Takara, Japan). Real-time PCR was performed using SYBR Green reagent (Takara, Japan) and specific primers (see Supplementary Table 3).
Whole-cell patch-clamp recording
The mouse was deeply anesthetized with isoflurane, and coronal sections (250 μm thick) containing the primary motor cortex were prepared using a vibratome (VT1200S; Leica Microsystems, Germany) in ice-cold artificial cerebrospinal fluid (aCSF: 119 mM NaCl; 2.5 mM CaCl2,2.5 mM KCl, 1.3 mM MgCl2, 1 mM NaH2PO4, 11 mM glucose, 26.2 mM NaHCO3, 1.3 mM MgCl2, and 290 mM mOsm, pH 7.4). The brain slices were recovered for 1 h at room temperature in aCSF. Slices were then transferred to the recording chamber with continuous perfusion of oxygenated aCSF containing 1 μM tetrodotoxin.
The whole-cell recording was performed on cortical L5 PNs using an epifluorescence microscope (Olympus, Japan). A glass pipette (WPI, US) was filled with internal solutions (135 mM CsMeSO3, 8 mM NaCl, 10 mM Na2-phosphocreatine, 0.25 mM EGTA, 2.17 mM Mg–adenosine 5′-triphosphate, 0.34 mM Na3–guanosine 5′-triphosphate, and 10 mM HEPES (pH 7.4) with CsOH). The miniature excitatory postsynaptic currents (mEPSCs) were recorded using 1 μM TTX and 20 μM bicuculline. The miniature inhibitory postsynaptic currents (mIPSCs) were recorded using 1 μM TTX, 20 μM NBQX and 50 μM D-AP5. All recordings were performed using a Multiclamp 700B amplifier (Molecular Devices, US). Traces were low-pass-filtered at 2 kHz and digitized at 10 kHz. The pipette resistance ranged from 4 to 6 MΩ. During the recording, series resistance (Rs) was maintained between 10–20 MΩ. If Rs changed >20%, the data were excluded. Offline data analysis was performed using Clampfit 10.0 software (Molecular Devices, US).
Fiber photometry
In brief, at 2 weeks after AAV2/9- GCaMP6s virus injection, an optical fiber [OD = 230 mm, numerical aperture = 0.37] was inserted towards the dorsolateral striatum (AP: +0.6 mm, ML: ±2.4 mm, DV: −0.3 mm). The fiber was fixed with a skull penetrating screw and dental acrylic. Mice were then individually housed for 1-week recovery. The fluorescence signals were detected through the optic fiber using a fiber photometry system (Thinker Biotech, China). The 410 nm signal and 470 nm signal were recorded at the same time, while the 410 nm channel was used to correct for false positive signals from animal movement and bleaching. The values of fluorescence change (ΔF/F) were calculated as (F − F0) / F0, where F0 was the averaged baseline fluorescence signal.
Statistics and reproducibility
GraphPad Prism 9.5 (GraphPad, US) was used for statistical analysis and data plotting. All data were plotted as mean ± SEM. Unpaired Student’s t-test was used for comparing the means between two groups in two-tailed manners. For 2 independent variables, 2-way analysis of variance (ANOVA) and Sidak post hoc comparison were adopted. A significant level was defined when P < 0.05. The sample size of each experiment was reported in the relevant figure captions. No replicate was performed in most of the studies, except in those of qPCR, in which 3 replicates were performed for each sample.
Data availability
The RNA sequencing data has been deposited into the SRA database (Accession number: PRJNA1155630). All other datasets used for generating the figures are included in Supplementary Data 1.
References
Goedert, M. NEURODEGENERATION. Alzheimer’s and Parkinson’s diseases: the prion concept in relation to assembled Aβ, tau, and α-synuclein. Science 349, 1255555 (2015).
Morris, H. R., Spillantini, M. G., Sue, C. M. & Williams-Gray, C. H. The pathogenesis of Parkinson’s disease. Lancet 403, 293–304 (2024).
Vázquez-Vélez, G. E. & Zoghbi, H. Y. Parkinson’s disease genetics and pathophysiology. Annu. Rev. Neurosci. 44, 87–108 (2021).
McGregor, M. M. & Nelson, A. B. Circuit mechanisms of Parkinson’s disease. Neuron 101, 1042–1056 (2019).
Roberts, R. F., Wade-Martins, R. & Alegre-Abarrategui, J. Direct visualization of alpha-synuclein oligomers reveals previously undetected pathology in Parkinson’s disease brain. Brain 138, 1642–1657 (2015).
Tanji, K. et al. Proteinase K-resistant α-synuclein is deposited in presynapses in human Lewy body disease and A53T α-synuclein transgenic mice. Acta Neuropathol. 120, 145–154 (2010).
Muntané, G., Dalfó, E., Martinez, A. & Ferrer, I. Phosphorylation of tau and α-synuclein in synaptic-enriched fractions of the frontal cortex in Alzheimer’s disease, and in Parkinson’s disease and related α-synucleinopathies. Neuroscience 152, 913–923 (2008).
Ren, J. X., Sun, X., Yan, X. L., Guo, Z. N. & Yang, Y. Ferroptosis in neurological diseases. Front. Cell. Neurosci. 14, 218 (2020).
Mahoney-Sánchez, L. et al. Ferroptosis and its potential role in the physiopathology of Parkinson’s disease. Prog. Neurobiol. 196, 101890 (2021).
Stockwell, B. R. et al. Ferroptosis: a regulated cell death nexus linking metabolism, redox biology, and disease. Cell 171, 273–285 (2017).
Lin, Z. H. et al. Quercetin protects against MPP+/MPTP-induced dopaminergic neuron death in Parkinson’s disease by inhibiting ferroptosis. Oxid. Med. Cell. Longev. 2022, 7769355 (2022).
Anandhan, A. et al. α-Syn overexpression, NRF2 suppression, and enhanced ferroptosis create a vicious cycle of neuronal loss in Parkinson’s disease. Free Radic. Biol. Med. 192, 130–140 (2022).
Wang, Z. L., Yuan, L., Li, W. & Li, J. Y. Ferroptosis in Parkinson’s disease: glia-neuron crosstalk. Trends Mol. Med. 28, 258–269 (2022).
Ko, C. J., Gao, S. L., Lin, T. K., Chu, P. Y. & Lin, H. Y. Ferroptosis as a major factor and therapeutic target for neuroinflammation in Parkinson’s disease. Biomedicines 9, 1679 (2021).
Li, Y. et al. CXCL12 is involved in α-synuclein-triggered neuroinflammation of Parkinson’s disease. J. Neuroinflammation 16, 263 (2019).
Rajendran, L. & Paolicelli, R. C. Microglia-mediated synapse loss in Alzheimer’s disease. J. Neurosci. 38, 2911–2919 (2018).
Zabel, M. K. et al. Microglial phagocytosis and activation underlying photoreceptor degeneration is regulated by CX3CL1-CX3CR1 signaling in a mouse model of retinitis pigmentosa. Glia 64, 1479–1491 (2016).
Sun, J. et al. Midbrain dopamine oxidation links ubiquitination of glutathione peroxidase 4 to ferroptosis of dopaminergic neurons. J. Clin. Invest. 133, e165228 (2023).
Di Lazzaro, V. et al. Motor cortex hyperexcitability to transcranial magnetic stimulation in Alzheimer’s disease. J. Neurol. Neurosurg. Psychiatry 75, 555–559 (2004).
Ponzo, V. et al. Impaired intracortical transmission in G2019S leucine rich-repeat kinase Parkinson patients. Mov. Disord. 32, 750–756 (2017).
Kreitzer, A. C. & Malenka, R. C. Striatal plasticity and basal ganglia circuit function. Neuron 60, 543–554 (2008).
Chen, S. X., Kim, A. N., Peters, A. J. & Komiyama, T. Subtype-specific plasticity of inhibitory circuits in motor cortex during motor learning. Nat. Neurosci. 18, 1109–1115 (2015).
Bartels, T., De Schepper, S. & Hong, S. Microglia modulate neurodegeneration in Alzheimer’s and Parkinson’s diseases. Science 370, 66–69 (2020).
Dagher, N. N. et al. Colony-stimulating factor 1 receptor inhibition prevents microglial plaque association and improves cognition in 3xTg-AD mice. J. Neuroinflammation 12, 139 (2015).
Hambright, W. S., Fonseca, R. S., Chen, L., Na, R. & Ran, Q. Ablation of ferroptosis regulator glutathione peroxidase 4 in forebrain neurons promotes cognitive impairment and neurodegeneration. Redox Biol. 12, 8–17 (2017).
Miotto, G. et al. Insight into the mechanism of ferroptosis inhibition by ferrostatin-1. Redox Biol. 28, 101328 (2020).
Taguchi, K., Watanabe, Y., Tsujimura, A. & Tanaka, M. Brain region-dependent differential expression of alpha-synuclein. J. Comp. Neurol. 524, 1236–1258 (2016).
Taguchi, K. et al. Differential expression of alpha-synuclein in hippocampal neurons. PLoS ONE 9, e89327 (2014).
Contractor, A., Ethell, I. M. & Portera-Cailliau, C. Cortical interneurons in autism. Nat. Neurosci. 24, 1648–1659 (2021).
Lanoue, A. C., Blatt, G. J. & Soghomonian, J. J. Decreased parvalbumin mRNA expression in dorsolateral prefrontal cortex in Parkinson’s disease. Brain Res. 1531, 37–47 (2013).
Hijazi, S. et al. Early restoration of parvalbumin interneuron activity prevents memory loss and network hyperexcitability in a mouse model of Alzheimer’s disease. Mol. Psychiatry 25, 3380–3398 (2020).
Fu, Y. et al. Adaptive structural changes in the motor cortex and white matter in Parkinson’s disease. Acta Neuropathol. 144, 861–879 (2022).
Ruden, J. B., Dugan, L. L. & Konradi, C. Parvalbumin interneuron vulnerability and brain disorders. Neuropsychopharmacology 46, 279–287 (2021).
Crapser, J. D. et al. Microglia facilitate loss of perineuronal nets in the Alzheimer’s disease brain. EBioMedicine 58, 102919 (2020).
Kvitsiani, D. et al. Distinct behavioural and network correlates of two interneuron types in prefrontal cortex. Nature 498, 363–366 (2013).
Ferguson, B. R. & Gao, W. J. PV interneurons: critical regulators of E/I balance for prefrontal cortex-dependent behavior and psychiatric disorders. Front. Neural Circuits 12, 37 (2018).
Zhang, C. et al. Dynamics of a disinhibitory prefrontal microcircuit in controlling social competition. Neuron 110, 516–531.e516 (2022).
Gothner, T., Gonçalves, P. J., Sahani, M., Linden, J. F. & Hildebrandt, K. J. Sustained activation of PV+ interneurons in core auditory cortex enables robust divisive gain control for complex and naturalistic stimuli. Cereb. Cortex 31, 2364–2381 (2021).
Khan, A. G. et al. Distinct learning-induced changes in stimulus selectivity and interactions of GABAergic interneuron classes in visual cortex. Nat. Neurosci. 21, 851–859 (2018).
Yang, S. S., Mack, N. R., Shu, Y. & Gao, W. J. Prefrontal GABAergic interneurons gate long-range afferents to regulate prefrontal cortex-associated complex behaviors. Front. Neural Circuits 15, 716408 (2021).
Mukherjee, A., Carvalho, F., Eliez, S. & Caroni, P. Long-lasting rescue of network and cognitive dysfunction in a genetic schizophrenia model. Cell 178, 1387–1402.e1314 (2019).
Zhang, C. L. et al. Inhibitory control of synaptic signals preceding locomotion in mouse frontal cortex. Cell Rep. 37, 110035 (2021).
Donato, F., Rompani, S. B. & Caroni, P. Parvalbumin-expressing basket-cell network plasticity induced by experience regulates adult learning. Nature 504, 272–276 (2013).
Casula, E. P. et al. Low-frequency rTMS inhibitory effects in the primary motor cortex: insights from TMS-evoked potentials. Neuroimage 98, 225–232 (2014).
Feng, Z. et al. Iron overload in the motor cortex induces neuronal ferroptosis following spinal cord injury. Redox Biol. 43, 101984 (2021).
Duan, C. et al. Activation of the PPARγ prevents ferroptosis-induced neuronal loss in response to intracerebral hemorrhage through synergistic actions with the Nrf2. Front. Pharmacol. 13, 869300 (2022).
Han, K. et al. Nrf2 knockout altered brain iron deposition and mitigated age-related motor dysfunction in aging mice. Free Radic. Biol. Med. 162, 592–602 (2021).
Zhang, P. et al. Ferroptosis was more initial in cell death caused by iron overload and its underlying mechanism in Parkinson’s disease. Free Radic. Biol. Med. 152, 227–234 (2020).
Ryan, S. K. et al. Microglia ferroptosis is regulated by SEC24B and contributes to neurodegeneration. Nat. Neurosci. 26, 12–26 (2023).
Zuo, Y. et al. Ferritinophagy-mediated ferroptosis involved in paraquat-induced neurotoxicity of dopaminergic neurons: implication for neurotoxicity in PD. Oxid. Med. Cell. Longev. 2021, 9961628 (2021).
Zhang, N. et al. Ferritin confers protection against iron-mediated neurotoxicity and ferroptosis through iron chelating mechanisms in MPP+-induced MES23.5 dopaminergic cells. Free Radic. Biol. Med. 193, 751–763 (2022).
Zhang, W. et al. Treadmill exercise relieves chronic restraint stress-induced cognitive impairments in mice via activating protein phosphatase 2A. Neurosci. Bull. 37, 1487–1492 (2021).
Acknowledgements
This work was funded by the STI2030-Major Projects (2022ZD0207600) to L.Z., National Natural Science Foundation of China (32070955 to L.Z., 81971067 to K.F.S.), Science and Technology Program of Guangzhou, China (202007030012 to K.F.S. and L.Z.), Science and Technology Program of Guangdong (2018B030334001 to K.F.S. and L.Z.), Guangdong Natural Science Foundation (2019A1515011772 to L.Z.).
Author information
Authors and Affiliations
Contributions
B.Z., K.C., and L.Z. designed the whole experiment. B.Z. performed all experiments except electrophysiological studies. Y.D., X.L., Z.X. and W.Z. assisted in the execution and data analysis of behavioral, molecular, and histological assays, respectively. X.H. performed the patch-clamp recordings. This work was supervised by K.-F.S. and L.Z. The manuscript was prepared by B.Z. and L.Z., with inputs from all authors.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Communications Biology thanks Bernard Schneider and the other anonymous reviewer(s) for their contribution to the peer review of this work. Primary handling editors: Christian Wozny, Anam Akhtar and Benjamin Bessieres.
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/.
About this article
Cite this article
Zhang, B., Chen, K., Dai, Y. et al. Human α-synuclein aggregation activates ferroptosis leading to parvalbumin interneuron degeneration and motor learning impairment. Commun Biol 7, 1227 (2024). https://doi.org/10.1038/s42003-024-06896-x
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s42003-024-06896-x
Comments
By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.