Open Access

Roles of α‑synuclein in gastrointestinal microbiome dysbiosis‑related Parkinson's disease progression (Review)

  • Authors:
    • Qingchun Lei
    • Tingting Wu
    • Jin Wu
    • Xiaogang Hu
    • Yingxia Guan
    • Ying Wang
    • Jinyuan Yan
    • Guolin Shi
  • View Affiliations

  • Published online on: August 16, 2021     https://doi.org/10.3892/mmr.2021.12374
  • Article Number: 734
  • Copyright: © Lei et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Parkinson's disease (PD) is the second most common neurodegenerative disease amongst the middle‑aged and elderly populations. Several studies have confirmed that the microbiota‑gut‑brain axis (MGBA) serves a key role in the pathogenesis of PD. Changes to the gastrointestinal microbiome (GM) cause misfolding and abnormal aggregation of α‑synuclein (α‑syn) in the intestine. Abnormal α‑syn is not eliminated via physiological mechanisms and is transported into the central nervous system (CNS) via the vagus nerve. The abnormal levels of α‑syn aggregate in the substantia nigra pars compacta, not only leading to the formation of eosinophilic Lewis Bodies in the cytoplasm and mitochondrial dysfunction in dopaminergic (DA) neurons, but also leading to the stimulation of an inflammatory response in the microglia. These pathological changes result in an increase in oxidative stress (OS), which triggers nerve cell apoptosis, a characteristic of PD. This increase in OS further oxidizes and intensifies abnormal aggregation of α‑syn, eventually forming a positive feedback loop. The present review discusses the abnormal accumulation of α‑syn in the intestine caused by the GM changes and the increased levels of α‑syn transport to the CNS via the MGBA, resulting in the loss of DA neurons and an increase in the inflammatory response of microglial cells in the brain of patients with PD. In addition, relevant clinical therapeutic strategies for improving the GM and reducing α‑syn accumulation to relieve the symptoms and progression of PD are described.

Introduction

Parkinson's disease (PD) is the second most common age-related neurodegenerative disorder after Alzheimer's disease (AD). With the increase in the proportion of the aging population, the incidence of PD is increasing (1). The prevalence of PD is 0.3% in the general population, and as high as 1% in the elderly over the age of 60 years, and 3% in individuals >80 years in industrialized countries (2). A major characteristic of PD is the accumulation of the misfolded α-synuclein (α-syn) protein in cerebral nerve cells, eventually leading to the loss of dopaminergic (DA) neurons in the substantia nigra pars compacta (SNpc) or the death of nerve tissue (3). This results in large areas of dead brain tissue, and the promotion of the formation of eosinophilic inclusion bodies, such as Lewis bodies (LBs) and Lewy neuritis (LN), in the cytoplasm (4). With the death of neurons in the brain, the clinical manifestations of PD comprise a static tremor, muscle rigidity, bradykinesia, abnormal posture gait and a series of non-motor symptoms, such as olfactory disorders, anxiety and depression, cognitive decline, sleep disorders, autonomic dysfunction, fatigue and pain (5,6). Several studies have demonstrated that mitochondrial dysfunction and oxidative stress (OS) serve a key role in the pathogenesis of PD, as they cause a loss of DA neurons (710). Several genes and signaling pathways are involved in the initiation and development of PD. For example, the familial autosomal recessive genes PTEN induced kinase 1 (PINK1)/parkin and the autosomal dominant mutations of Leucine-rich repeat kinase 2 (LRRK2) regulate mitochondrial dysfunction and PD pathogenesis (11,12). Furthermore, it has been reported that a mutated LRRK2 gene can induce PD pathogenesis, which is dependent on the PINK1/parkin pathway via independent mechanisms (13). In addition, several genes, such as GTP cyclohydrolase 1 (GCH1) (14), coiled-coil-helix-coiled-coil-helix domain containing 2 (CHCHD2/PARK22) (15) and VPS35 retromer complex component (VPS35/PARK17) (16) are involved in the induction of mitochondrial dysfunction of patients with PD. Environmental factors are also a pathogenic factor of PD. For example, the occupational and environmental exposure to pesticides and cytokine pathways (17), the influence of genetic polymorphisms on pesticides (18) and the dysregulation of the microRNA (miRNA/miR) network caused by pesticide exposure (19), all serve an important role in the pathophysiology of PD via neurodegeneration of SNpc DA neurons, mitochondrial dysfunction or oxidative damage (20). Some of the related genes, such as miRNAs, may serve as novel non-invasive early biomarkers for the prediction and prognosis of PD (21). It has also been demonstrated that exposure to polystyrene microplastics can induce intestinal injury and neurodegeneration through increased production of reactive oxygen species (ROS) in Caenorhabditis elegans (C. elegans) (22). Recent reports have reported that coronavirus 2019 (COVID-19) colonizes in the gut and the central nervous system (CNS), where it triggers neuroinflammation and neurodegenerative processes, suggesting that patients infected with COVID-19 may be susceptible to certain neurodegenerative disorders, such as PD (23,24). However, the molecular mechanism underlying the development and progression of PD requires further investigation.

Gastrointestinal dysfunction, an important non-motor symptom of PD, not only has a high incidence, but also appears several years prior to the characteristic motor symptoms (25). Currently, the influence of intestinal microbiota on PD has been studied extensively by scientists, and has been termed the microbiota-gut-brain axis (MGBA) (26). Numerous studies have demonstrated that abnormal intestinal microbiotas are not only closely associated with gastrointestinal dysfunction in patients with PD but may also be an important mechanism underlying the pathological process of PD (27,28). Recently, it has been reported that dysfunction of gastrointestinal microbiomes (GM) occurs earlier than/or at the same time as PD, and the pathology of PD is closely associated with the changes of the GM (29). Nielsen et al (30) demonstrated that individuals infected with Helicobacter pylori are more likely to induce the development of PD. Devos et al (31) reported that most patients with PD have colitis, which can enhance the peripheral and brain inflammatory response, and promote the pathogenesis of PD.

α-syn consists of 140 amino acids and the gene encoding it, synuclein α (SNCA), is comprised of five exons and is located at chromosome 4q21.3-q22 (32). It is widely expressed in the CNS, mainly in the presynaptic terminals, and is involved in the regulation of neurotransmission and synaptic homeostasis (33). The α-syn family of proteins contains three members, synapsin-I, synapsin-II and synapsin-III (34). According to the C-terminal splicing structure, α-syn can be divided into α and β subtypes (34). Currently, five synapsin proteins (synapsin-Iα, synapsin-Iβ, synapsin-IIα, synapsin-IIβ and synapsin-IIIα) have been detected (35). A pathological characteristic of PD is the accumulation of the misfolded α-syn protein involved in slow and progressive degeneration of DA neurons in the SNpc (36). α-syn exhibits characteristics of prion-like protein during PD pathogenesis; the misfolded α-syn is an ‘infectious’ protein spreading pathology into the CNS via the vagus nerve (VN) by forming a template that seeds misfolding for nearby α-syn proteins, turning the endogenous physiological protein into a pathogenic protein (37). Previous studies have demonstrated that neurons can absorb α-syn in vitro and in vivo, and α-syn can also transmit between two neurons to neighboring neurons via endocytosis (38,39). However, further studies are required to determine the specific mechanisms. In addition, studies have reported that an abnormal GM can enhance the levels of inflammation via the induction of immune responses, leading to the misfolding of α-syn in patients with PD (40,41). The dysfunction of the enteric nervous system (ENS), and the accumulation of anti-α-syn immune response proteins were detected in the ENS ganglia in mice with α-syn mutations (either A53T or A30P from insertions of an entire human SNCA gene) when they were 3 months old (42). Braak et al (43) demonstrated that PD is initiated by unknown pathogens that traverse the gastric epithelial lining and lead to the formation of misfolded α-syn in nerve cells of the submucosal plexus. The pathological formation of α-syn is retrogradely propagated along the axonal and transneuronal axis along with the VN to reach the CNS (37). Braak et al (43) also reported that the accumulation of misfolded α-syn in the peripheral nervous system (PNS) occurred earlier than that in the CNS of patients with PD. Several studies have speculated that the initiation of PD originates from the PNS, and is retrogradely transported towards the CNS via the PNS (44,45). These results suggest that the abnormal GM can affect the development of PD by inducing α-syn misfolding, abnormal aggregation and transmission from VN to CNS.

Several studies have reported that GM serves an important role in the development of PD by regulating misfolded and abnormal aggregation of α-syn and the MGBA (46,47). The present review focuses on the association between GM, α-syn and MGBA in patients with PD, with an emphasis on the mechanisms of the GM and its role in the pathogenesis of PD (Fig. 1). In addition, adaptable novel potential treatment strategies are discussed. The present review offers an insight into the role of α-syn and MGBA in the pathological progression of PD, and highlights the potential of α-syn and MGBA as relevant drug targets, as well as discussing potential therapeutic candidates.

Figure 1.

Schematic representation of dysregulation of the MGBA pathways implicated in the pathogenesis of PD. The aberrant GM or their products, such as LPS and the inflammatory factors involved in intestinal mucosal inflammation, reactive oxygen/nitrogen species and disruption of intestinal barrier integrity, induce misfolding of α-syn, resulting in abnormal aggregation and the formation of truncated fragments in the ENS of patients with PD. The misfolded, abnormal aggregation and truncated fragments of α-syn are transported from the ENS to the CNS via projections of the VN, as well as autonomic enteric fibers at a speed of 5–10 mm/day. Once abnormal α-syn from the ENS reaches and deposits into the CNS, it subsequently spreads in the CNS through the brainstem, midbrain, basal forebrain and finally reaches the cortical areas via a mechanism similar to that of prion-like protein. Due to an increase in α-syn aggregation in the SNpc of CNS, eosinophilic LBs are formed in the cytoplasm and mitochondrial dysfunction occurs in the DA neurons, resulting in their degradation, which stimulates an inflammatory response of the microglia. These pathological changes result in the production of OS in the brain, which triggers cell apoptosis, and thus is involved in the initiation and progression of PD. The increase in OS further oxidizes and intensifies abnormal aggregation of α-syn in the brain, eventually forming a positive feedback loop. MGBA, microbiota-gut-brain axis; PD, Parkinson's disease; GM, gut microbiota; LPS, lipopolysaccharide; α-syn, α-synuclein; ENS, enteric nervous system; CNS, central nervous system; VN, vagus nerve; SNpc, substantia nigra pars compacta; LB, Lewis Body; DA, dopaminergic; OS, oxidative stress.

An aberrant GM affects α-syn accumulation in the intestines of patients with PD

Numerous studies have confirmed that the GM regulates the autonomic nervous system and CNS via the immune system, neuroendocrine system, direct neural conduction and the interaction of the MGBA (4850). Changes in GM are associated with the pathological process of PD (51).

Changes of the GM in patients with PD

Changes to the GM have been widely reported amongst patients with PD. Scheperjans et al (48) reported that ~77.6% of patients with PD exhibited a reduction of Prevotellaceae (family), and a notable increase in several pathogenic Gram-negative bacteria, such as Enterobacteriaceae (family), Verrucomicrobiaceae (family) and Escherichia coli (species) from their feces. Gerhardt and Mohajeri (52) demonstrated that there was a significant decrease of Prevotellaceae (family), Prevotella (genus) and Prevotella Copri (species) in the intestines of patients with PD. Lubomski et al (53) reported a notable alteration in the microbial population of Firmicutes in the GM of patients with PD compared with healthy individuals. The quantity of Lactobacillaceae (family) and Lactobacillus (genus) increased, whereas the abundance of Lachnospiraceae (family), such as Ruminococcus (genus), Blautia (genus), Dorea (genus), Roseburia (genus) and Faecalibacterium (genus) decreased. According to these results, the researchers hypothesized that the types of pro-inflammatory bacteria in the intestines of patients with PD significantly increased, whereas the types of probiotic bacteria decreased. Keshavarzian et al (54) further confirmed this hypothesis, demonstrating that the proportion of pro-inflammatory bacteria in the intestines of patients with PD increased by studying the mucosal-associated bacteria and fecal microbiota. The results demonstrated that the diversity of fecal microbiota communities was not significantly altered between PD and healthy control (HC) individuals, but the α-diversity at the phylum level and the richness of the genus level were significantly different. Specifically, at the phylum level, the abundance of Bacteroidetes, Proteobacteria and Verrucomicrobia in the fecal microbiota of the PD group significantly increased, whereas the presence of Firmicutes in the fecal microbiota of HC was higher in the HC group. Similar results were obtained by assessing the fecal samples of the PD group compared with the HC group at the genus level, in which the abundance of pro-inflammatory bacteria, such as the Akkermansia, Oscillospira and Bacteroides were increased, and the abundance of butyrate-producing bacteria that produce anti-inflammatory short chain fatty acids (SCFAs), such as Blautia, Coprococcus and Roseburia were significantly reduced in the fecal samples of the PD group. The abundance of Coprobacillaceae (family), Dorea (genus) and anti-inflammatory Faecalibacterium (genus) were rich in the mucosal-associated bacterial populations of the HC group, whereas the Oxalobac-teraceae (family) and Ralstonia (genus) were richer in the PD group (55). Li et al (56) reported that the composition of GM was slightly altered between the healthy and Han Chinese patients with PD using next generation sequencing to analyze the feces. Consistent with other studies, the abundance of Bacteroides (genus) and Prevotellaceae (family) significantly increased in the healthy Han Chinese individuals, whereas the abundance of Ruminococcaceae (family), Verrucomicrobiaceae (family), Porphyromonaceae (family), Hydrogenoanaerobacterium (family) and Lachnospiraceae NK4A (family) increased significantly in the Han Chinese patients with PD. In addition, these studies failed to demonstrate an increase in the abundance of Bifidobacterium (family) and Enterobacteriaceae (family) in the feces of Han Chinese patients with PD, an inconsistent result compared with the study of Caucasian patients with PD (5759). In summary, numerous changes to the GM are involved in the pathological process of PD (Table I).

Table I.

Studies evaluating the changes in the gut microbiota in patients with PD.

Table I.

Studies evaluating the changes in the gut microbiota in patients with PD.

Changes in the GM in patients with PD↓/↑(Refs.)
Changes at the phylum level
  Bacteroidetes; Proteobacteria; Verrucomicrobia(54)
Changes at the family level
  Prevotellaceae(48,52)
  Enterobacteriaceae, Verrucomicrobiaceae(48,56)
  Lactobacillaceae(53)
  Lachnospiraceae(53)
  Coprobacillaceae, Oxalobac-teraceae(55)
  Ruminococcaceae, Porphyromonaceae(56)
  Hydrogenoanaerobacterium, Lachnospiraceae NK4A(56)
Changes at the genus level
  Prevotella(52)
  Lactobacillus(53)
  Ruminococcus, Blautia, Dorea, Roseburia, Faecalibacterium(53,55)
  Akkermansia, Oscillospira, Bacteroides, Ralstonia(55)
  Coprococcus(55)
Changed in species level
  Escherichia coli(48)
  Prevotella copri(52)

[i] ↓, significantly reduced in patients with PD; ↑, significantly increased in patients with PD; PD, Parkinson's disease.

Due to changes in the gut microbiota composition that occur during the course of PD, several studies have investigated fecal microbiota transplant (FMT), a novel therapeutic method, which involves the transfer of gut microbiota from a healthy individual to another via oral administration of fecal material in rodents, or via medication that alters the GM or gastrointestinal endoscopy in humans, as a method to improve the symptoms of constipation in patients with PD (60,61). Tan et al (62) reported that the FMT procedure is an effective treatment in 65.6% of patients with PD, and that several patients exhibited an increase in spontaneous bowel movements by 1–2 times per week in the process of FMT treatment. In a PD mouse model, FMT is necessary for the neuroprotective effects of osteocalcin (63). In addition, FMT alleviates dyskinesia and neurodegeneration of striatal DA, reduces neuroinflammation and activates microglia and astrocytes in the brain of PD mice. Furthermore, FMT can also increase the levels of 5-hydroxytryptophan, decrease fecal SCFAs and improve gut microbial dysbiosis in the intestinal tracts of PD mice (64).

Influence of GM alterations on α-syn

The barrier functions of the gut epithelium serve an important role during host-microbiome interactions, and the disruption of this barrier can lead to intestinal inflammation, production of reactive oxygen/nitrogen species in the gut, and a shift in microbial composition towards pro-inflammatory bacteria (65). Several studies have demonstrated that one of the potential mechanisms by which α-syn enters into the mucosal neuronal tissue is the generation of OS and the disruption of intestinal barrier integrity via aberrant changes to the GM in patients with PD (66,67). Given the effect of aberrant GM on the gastrointestinal barrier, the resultant translocation of bacteria or their products, such as lipopolysaccharide (LPS) can increase OS and intestinal inflammation, which in turn increases the mucosal intestinal permeability, also known as leaky gut, and increases the ability of α-syn to communicate with the ENS (53). Immunohistochemical staining of postmortem analyses of intestines in patients with PD exhibited an inevitable association between impaired intestinal barrier integrity, the increase in the intestinal bacterial flora, the high levels of expression of inflammatory genes and the abnormal accumulation of α-syn in the ENS (6870). Other studies have reported that the volatile SCFA, particularly butyrate, serves a vital role in maintaining intestinal barrier integrity, and a lack of SCFA can increase intestinal permeability, which has been confirmed in studies of patients with inflammatory bowel disease (71,72). Further studies on the GM have reported that Prevotellaceae is involved in the formation of intestinal mucins and the production of SCFA through fiber fermentation in the sigmoid (73,74). Thus, the decrease of Prevotellaceae in the intestines of patients with PD can lead to a decrease in intestinal mucus and an increase in intestinal permeability, and this serves as a prerequisite for entry of α-syn into the ENS via the intestinal barrier, and to maintain excessive α-syn expression or even promote its misfolding (75). The treatment of a colitis mouse model with butyrate can reduce the expression of TNF-α in the intestines and reduce cell shedding (76).

An aberrant GM or their products are involved in the misfolding, abnormal aggregation and presence of truncated fragments of α-syn in the ENS of patients with PD (66,69). In particular, the LPS generated by inflammatory bacteria can increase the nitration and oligomerization of α-syn by upregulating the expression of inducible nitric oxide synthase (iNOS), suggesting that LPS not only increases the inflammatory response to induce gut leakiness and the communication of α-syn, but can also accelerate the neurodegenerative process via the influence of α-syn (55). The monocyte/macrophage-related signaling pathway is involved in the aforementioned biological processes (77). The LPS generated from the gram-negative bacteria can downregulate the expression of tight junction proteins in the intestinal epithelial cells, such as occludin, and upregulate the expression of TNF-α, which activates the macrophages and promotes the expression of α-syn in a mouse model (78,79). Bhattacharyya et al (80) demonstrated that LPS binds to the α-helical intermediates of α-syn to form a lipid-protein complex that acts as a scaffold for growth of α-syn fibers, and rapid nucleation based on CD spectra analysis. LPS also accelerates abnormal aggregation of α-syn by increasing the production and reducing the half-life and lag time of α-syn (81). Wang et al (82) reported that inflammatory activators, such as LPS, aluminum potassium sulfate crystals, nigericin and vitamin K3 (menadione) can activate the inflammasome, including caspase-1, by inducing the cleavage of procaspase-1 to the active caspase-1, which is directly involved in cleaving and inducing the aggregation of α-syn. Furthermore, the 10 or 30-residues of α-syn N-terminal truncations alter the conformation of fibril, thus contributing to a reduction in its stability and induce their compatibility with normal α-syn. The 20-residues of α-syn C-terminal truncations result in it exhibiting unique prion-like properties (83). The prion conformation of α-syn interferes with the lysosomal and proteasomal degradation processes, and ultimately promotes aberrant accumulation (84).

α-syn accumulation contributes to neurodegeneration in the CNS of patients with PD

The primary pathological characteristic of PD is the abnormal aggregation of α-syn in the SNpc (85). Due to its prion-like protein activation mechanism, α-syn can misfold to cause self-abnormal aggregation, and can also transmit signals amongst nervous cells, and spread throughout the nervous system (86). Several studies have indicated that the ENS may be the channel by which α-syn is transferred from the PNS to the CNS (41,87).

Abnormal α-syn in the intestines of patients with PD enters the CNS via the MGBA

Misfolded α-syn can transfer from affected to unaffected cells and serve as a template for pathophysiological aggregation of α-syn in neuronal cells (8891). It has been speculated that α-syn may be transmitted to the dorsal nucleus of the VN via VN fibers at a speed of 5–10 mm/day, and it accumulated in the neurons to trigger cell apoptosis after α-syn was injected into the intestinal wall of rats (92). Kim et al (93) injected 25 µg misfolded α-syn, preformed fibrils (PFF), into the pyloric stomach and upper duodenum (UD) of mice, and detected the distribution of PFF in the brain tissues after 1, 3, 7 and 10 months. The results demonstrated that PFF aggregated in the stomach and spread into the VN 1 month after injection. In the following 3–10 months, the distribution of PFF was detected consecutively in the medulla oblongata, locus coeruleus in the pons, SNc in the ventral mesencephalon brain (VMB), prefrontal cortex and olfactory bulb. The number of DA neurons in the SNpc of mice significantly decreased, and this was accompanied by a decrease in forelimb strength, grip strength, the hindlimb strength, and muscle and motor coordination ability of mice based on a rotarod test 7 months after injection. However, PFF was not detected in the aforementioned areas of the brain in the mice whose VN fibers near UD and neck were cut by truncal vagotomy 7 months after injection with the same dosage of PFF. The death of neuronal cells and the production of LBs were not detected. Braak et al (94) demonstrated that several misfolded α-syn proteins had accumulated in a portion of the ENS that regulated intestinal function and connected the gut to the brain of patients with PD. They speculated that the ENS may be the α-syn accumulation starting point, and α-syn was transported from the VN to the VMB, where the SNpc killed DA cells selectively, resulting in PD.

Abnormal α-syn causes the death of DA neurons in patients with PD via OS

The abnormal α-syn in DA cells transported from MGBA often involves mitochondrial dysfunction and the generation of OS, which serve important roles in the development of PD (95,96). It has been reported that overexpression of α-syn in GT1-7 cells increases mitochondrial volume and abnormal vacuolated cristae (97). Hu et al (98) demonstrated that in the α-syn-overexpressing SH-SY5Y cells, α-syn can interact with ATP-dependent Clp protease (ClpP) to inhibit the activity of ClpP, leading to mitochondrial oxidative damage and neurotoxicity. Chong et al (99) performed experiments using an N27 immortalized rat mesencephalic dopaminergic neuronal model of PD cell lines that stably expressed wild-type human α-syn, and hypothesized that the expression levels of caspase-3 and 9 increased as the OS increased, and the expression of Akt gradually decreased, resulting in cells undergoing mitochondrial dependent apoptosis. This suggests that α-syn promotes apoptosis by increasing OS. Dryanovski et al (100) and Tapias et al (101) reported that the increase in levels of α-syn in DA cells promotes the generation of OS. Musgrove et al (102) detected large quantities of nitrated α-syn in the brain tissues of patients with PD after death using an oxidized modified antibody. The study of SH-SY5Y human neuroblastoma cells stably expressing the C-terminal half of Venus YFP-tagged α-syn or the N-terminal half of Venus YFP-tagged α-syn demonstrated that the cells treated with 25 µm paraquat exhibited a stronger fluorescent signal due to the increase in OS compared with the control groups. However, after co-culturing with 25 µM paraquat and oxidized modified inhibitor, the intensity of the fluorescent signal in the experimental group decreased compared with the control group. Therefore, it was hypothesized that the increased OS was the result of increased spreading of α-syn from the dorsal motor nucleus of the VN to the rostral brain regions, resulting in enhanced cell-to-cell α-syn transmission in patients with PD. Martin et al (103) and Stichel et al (104) reported that LBs were formed in transgenic mice that expressed the human mutational α-syn. Furthermore, the activity of neuronal mitochondrial complexes I and IV significantly reduced. Conversely, inhibiting the activity of mitochondrial complex I promoted the accumulation of α-syn in mice brains, suggesting that increased OS resulted in increased levels of oxidatively modified forms of α-syn, resulting in increased α-syn aggregation and thus, cell-to-cell transmission (105). Thus, an increase in OS and the aggregation of α-syn forms a positive feedback loop throughout the progression of PD (95).

Abnormal α-syn induces neuroinflammation in patients with PD via activation of microglia

In the brain, microglia act as innate immune cells and serve a vital role in the pathogenesis of PD (106). During the early stages of PD, microglias are activated and are positive for human leukocyte antigen-DR (HLA-DR) (107). Microglias maintain continuous activation through the development of the disease (108). The activated microglia release several pro-inflammatory factors that target the blood vessel endothelial cells of the blood-brain barrier (BBB) and promote the expression of adhesion molecules on their surface (109). The adhesion molecules induce T cells and monocytes to enter the brain through the BBB, which further release more pro-inflammatory factors, resulting in neuroinflammation and neuronal apoptosis (110). α-syn is involved in the activity of microglial cells in patients with PD (111). Studies on brain tissue samples of patients with PD revealed high expression of α-syn in neurons, and the microglia were HLA-DR positive (112,113). The microglial cells treated with different (from low to high) concentrations of α-syn can increase the pro-inflammatory effect of microglial cells, and the mRNA expression levels of TNF-α, interleukin (IL)-1β, cyclooxygenase-2 and iNOS levels also increase, which causes apoptosis of nerve cells (114). In the brain, toll-like receptor 4 (TLR4) is primarily expressed by microglial cells, and a small amount of TLR4 is expressed by astrocytes, oligodendrocytes and neurons (115,116). TLR4 can induce the activity of microglial cells, and subsequently induce the secretion of inflammatory chemokines and cytokines (117,118). Fellner et al (119) demonstrated that α-syn can active microglial cells by targeting TLR4, and inducing the production of nuclear factor-κB (NF-κB) and the secretion of cytokines by treating TLR4-deficient and wild-type mice with different forms of α-syn (full length soluble, fibrillized and C-terminally truncated). Further research demonstrated that the C-terminally truncated α-syn was the most effective inducer of TLR4-dependent microglial activity (120). Choi et al (121) indicated that α-syn also reduces cell autophagy by decreasing lysosomal and proteasomal degradation, which further results in increased expression of TLR4 and further strengthened TLR4-dependent microglial activity. Thus, α-syn is considered an inducer of microglial activation in patients with PD. Notably, several studies have reported that NF-κB, matrix metalloproteinase and protease activated receptor 1 are also involved in microglial activity via monomers, polymers and nitration of α-syn (122,123), which further induces the production of excessive ROS in microglia, and thus, the death of DA neurons (124).

Activated microglia cells inhibit the activity of nuclear factor E2-related factor 2 (Nrf2), an antioxidant transcription factor associated with the anti-inflammatory capacity of microglia, via oxidative modification (125). The expression levels of TNF-α, IL-6, IL-1β and iNOS are significantly upregulated in Nrf2-deficient microglia (126). Shavali et al (127) reported that activated microglia further promote nitrification of α-syn through nitric oxide, thereby inducing the continuous spread of α-syn to the adjacent neurons and resulting in neuronal cell death. It has been demonstrated that TNF-α produced by α-syn-activated microglia can result in impaired function of mitochondrial complex I and the generation of OS; in turn, the OS promotes further aggregation of α-syn in neuronal cells (128). Subsequently, the aggregated α-syn increases OS further, forming a positive feedback loop between neuroinflammation, OS and the aggregation and spread of α-syn in patients with PD (129).

Novel treatments based on the MGBA and α-syn for PD

The current treatments available for patients with PD not only have no effect on its relentless progression but may also induce several side effects. There are no effective therapeutic strategies that target the MGBA to slow or halt the neurodegenerative process, or reduce motor and non-motor symptoms. Nutrition based probiotics or prebiotics-based interventions inhibit neuroinflammation and ameliorate the diffusion of α-syn in the MGBA, which offer a novel therapeutic strategy for the treatment of PD and overcome the disadvantages of current therapies (130). An overview of the studies summarizing the novel treatments based on MGBA and α-syn for the management of PD is presented in Fig. 2.

Current clinical treatments for PD

The death of cerebral DA neurons in patients with PD leads to a decrease in secretion of dopamine and other neurotransmitters, resulting in the motor symptoms. Currently, pharmacological therapeutic strategies primarily compensate for the loss of dopamine and other neurotransmitters to alleviate motor symptoms temporarily. The drugs often include dopamine receptor activators and/or the dopamine precursor L-3, 4-dihydrophenylalanine (levodopa). However, levodopa not only relieves the neurodegenerative processes, but also causes several effects, including nausea, emesis, abnormal involuntary movement of head, face, tongue, upper limbs, depression and dysuria. Conversely, levodopa administered orally provides a favorable gastrointestinal environment required for the drug absorption (131). However, most patients with PD also exhibit gastrointestinal disorders, resulting in poor drug absorption (132). Several drug trials on healthy volunteers had demonstrated that levodopa can delay the time of gastric emptying, and long-term use of levodopa can exacerbate the gastrointestinal dysfunction in patients with PD (133). Further studies have reported that long-term use of levodopa may lead to fluctuation of symptoms and resistance to levodopa (134,135). Tyrosine decarboxylase (TDC), which is involved in the transformation of levodopa to dopamine and prevents the uptake of levodopa in the small intestine of patients with PD, is specially coded in the genome of Lactobacillus and Enterococcus in the intestine (136). The content of TDC is negatively correlated with the dosage of levodopa in the feces of patients, which may be used as a biomarker to evaluate the therapeutic effect of levodopa (137). The combination of a TDC inhibitor and levodopa can reduce the dosage and the dependence on levodopa for more effective treatment of PD (138).

Improving GM dysbiosis using probiotics and prebiotics in patients with PD

The abundance of Prevotellaceae was shown to be significantly reduced, the levels of SCFAs were decreased in the feces, whereas the content of pro-inflammatory bacteria, such as Proteus, Pseudomonas and Enterobacteriaceae were increased in patients with PD (6). A study of a nude mouse was demonstrated to exhibit motor symptoms and loss of DA neurons following transplant with intestinal microflora from PD mice (139). However, the transplantation of probiotics can significantly maintain intestinal microbial homeostasis, increase the levels of SCFAs in feces, decrease the activation of microglia and increase the activity of DA neurons in the brains of PD mice (64). Currently, the most common probiotics are Lactobacillus, Enterococcus, Bifidobacterium and yeast (140). Several studies have reported that probiotics can enhance intestinal epithelial integrity and stimulate the homeostasis of the intestinal mucosal immune system (40,141). In addition, probiotics can regulate the immune response, inhibit the growth of pro-inflammatory bacteria and increase the levels of SCFAs in the intestines of patients with PD by regulating the MGBA, reducing the production and aggregation of prion-like α-syn (142,143). Goya et al (144) reported that the probiotic, Bacillus subtilis can reduce α-syn aggregation and clear previously aggregated α-syn via biofilm formation in a C. elegans model of PD. Currently, the most effective means of transplanting probiotics is FMT, which contributes to the reconstruction of a healthy intestinal microflora by transplanting the healthy intestinal microflora into the gastrointestinal tract to effectively improve the symptoms of patients with PD (145). This treatment method has achieved positive effects in clinical trials of PD, autism and chronic fatigue syndrome (146). The supplementation of probiotics or FMT not only improves gastrointestinal function, but also enhances the intestinal absorption capacity of levodopa (147). Although most of the animal model studies demonstrated that FMT is an effective treatment for PD (148,149), further studies are required to determine the molecular mechanism by which FMT exerts its beneficial effects before it can be recommended for use in humans. Optimistically, FMT is gradually being clinically used as a medicinal therapy in humans for the treatment of PD (149). This may effectively reconstitute the major commensal bacterial populations and re-establish the diversity and composition of the GM in patients with PD. Thus, this practice may become commonplace in the near future.

Prebiotics, such as Galacto-Oligosaccharides and Fructo-Oligosaccharide, are non-digestible oligosaccharides, primarily synthesized from lactose or fructose (150). Prebiotics selectively activate certain probiotics in the gut, such as Bifidobacteria (151), promote its metabolism and produce SCFAs to maintain the integrity of the intestinal epithelia and regulate the mucosal immune response (152). Savignac et al (153) demonstrated that prebiotics can increase the expression levels of brain derived neurotrophic factor, which serves a protective role in neurons in the dentate gyrus of the hippocampus of rats. Taken together, prebiotics can selectively reduce intestinal permeability and inflammation via activation of the metabolism of probiotics. Thus, it is hypothesized that the combination of FMT, prebiotics and traditional drugs may both improve intestinal dysfunction and protect neurons, as well as effectively reducing the dosage of levodopa or other drugs to reduce drug dependence (154).

Inhibiting the expression and abnormal aggregation of α-syn in patients with PD

Given that α-syn serves a key role in the development of PD, an increasing number of studies have demonstrated that the inhibition of α-syn expression and abnormal aggregation is an important method for the treatment of PD (155). Preclinical gene therapy using RNA interference (RNAi) targeting α-syn mRNA via gene-silencing is being assessed, and the results have indicated that it can effectively inhibit the expression of α-syn in fibroblasts (156). In addition, miRNAs can downregulate the translation of target mRNAs as post-transcriptional regulators, through binding to the 3′-untranslated region (UTR) complementary sequences (157), which can be used as another means of gene therapy for the treatment of PD. Junn et al (158) reported that miR-7 can bind to the 3′-UTR of α-syn mRNA complementary sequences to inhibit the expression of α-syn, effectively relieving its inhibitory effect on the proteasome, and thus promoting the degradation of α-syn. Masliah et al (159) demonstrated that PD transgenic mice immunized with full length human α-syn protein exhibit decreased accumulation of the misfolded α-syn in neuronal cell bodies and synapses, resulting in reduced PD symptoms, and the mice that produced high affinity antibodies exhibited a greater capacity to reduce α-syn expression compared with the mice that produced lower affinity and/or lower titers of anti-α-syn antibodies. Ghochikyan et al (160) also indicated that a PD mouse model immunized with full length human α-syn protein can produce high affinity antibodies, which can reduce the accumulation of the aggregated forms of α-syn in neuronal cells to relieve neurodegeneration. Other studies indicated that monoclonal antibodies that target α-syn can inhibit the expression and aggregation of α-syn in a PD mouse model, as well as in patients (161,162). Upregulation of α-syn in the extracellular matrix can be recognized by specific antibodies and cleared by the macrophages through phagocytosis (163,164). In addition, α-syn specific antibodies can reduce the activation of CD4+ T cells by clearing cytokines, such as IL-2 and TNF-β, which are released by the activated CD4+ T cell-mediated neurodegeneration (165). A study demonstrated that monoclonal antibodies against α-syn can decrease LB/LN formation to reduce the loss of primary cultured neuronal cells by preventing the uptake of α-syn-PFF (166). Due to the α-syn form, the oligomers and fibrils, as well as the post-translational modifications of α-syn in patients with PD, such as acetylation, phosphorylation and truncation, several studies have demonstrated that D10, a single chain antibody of α-syn, had the highest affinity with α-syn; the single chain antibody D5 can effectively target the oligomer of α-syn; the single chain antibody syn-O1, -O2 and -O4 can effectively reduce neuroinflammation and the loss of neurons by specifically targeting the oligomer of α-syn in the CA3 area of the hippocampus; the single chain antibody syn-10h targets α-syn trimers, and antibodies syn-F1 and F2 prevent neuronal loss and facilitated amelioration in behavioral defects by specifically recognizing and clearing α-syn fibers (167,168). In addition, a specific antibody, LS4-2G12, which recognizes phospho-α-syn (ser129), is highly sensitive to the aggregation in the tissues of patients with PD, and protects neurons from damage mediated by the immune system (169,170). TLR4 inhibitors can be used as a potential drug for the treatment of PD (171). The inhibitor of TLR4 can effectively reduce intestinal inflammation and permeability in patients with PD, which can be used as a competitive inhibitor of mutant α-syn in the brain of patients with PD to inhibit the activation of microglia (172). Taken together, the use of small molecules or antibodies to reduce the toxicity of α-syn may serve as an effective method to inhibit the pathogenesis and development of PD (173).

The failure to eliminate α-syn is also an important cause of PD, and autophagy serves a vital role in the elimination of abnormally aggregated α-syn (174). mTOR complexes are widely involved in regulating apoptosis and autophagy (175). MSDC-0160, a mitochondrial pyruvate carrier inhibitor, enhances autophagy by inhibiting the activation of mTOR in cells (176). It has been reported that MSDC-0160 can effectively protect DA neurons in an MPTP-induced mouse model, and decreases α-syn-induced neural toxicity by increasing autophagy in a C. elegans model of PD (177). In nerve cells, the expression of Beclin 1 can reduce apoptosis and enhance autophagy (178). Spencer et al (179) overexpressed Beclin 1 in a PD mouse model using lentivirus to activate autophagy and reduce the accumulation of α-syn; however, the study only assessed this method in a mouse model. The levels of activated C-Abl, a tyrosine kinase, in the brain of patients with PD increased due to α-syn aggregation (180). Nilotinib, an antitumor drug, can promote intracellular autophagy by inhibiting the PI3k/Akt/mTOR signaling pathway to inhibit C-Abl (181). A recent non-placebo-controlled study reported that the lower dose of Nilotinib is effective and safe for the treatment of PD by effectively inhibiting α-syn aggregation (182). Taken together, these findings suggest that enhancing autophagy and inducing the degradation of α-syn are potential therapeutic strategies for the treatment of PD.

Conclusion

The intestinal microbiota contributes to the pathogenesis of PD, which has changed the previous view that the etiology of PD is concentrated on the brain. Misfolded and abnormally aggregated α-syn in the intestines of patients with PD is transported from the intestines to the brain via the MGBA. The spreading and abnormal aggregation of α-syn in the brain results in the formation of LBs in DA neurons, the activation of microglia, the production of inflammatory factors, an increase in OS, and ultimately apoptosis. MGBA not only serves an important role in the stability of the digestive system, but also serves a key role in the pathogenesis of PD. MGBA abnormalities may be one of the causes of PD. Currently, several novel clinical therapeutic strategies that target the MGBA to slow or halt the neurodegenerative process or reduce the motor and non-motor symptoms, such as food-based therapies, vagotomy, inhibiting the expression and abnormal aggregation of α-syn in patients with PD using RNAi or other gene modification measures, as well as FMT have been suggested. In addition, several genes serve prominent roles between MGBA and the progression of PD, such as the aforementioned PINK1/parkin pathway, LRRK2, TLRs and the PI3k/Akt/mTOR pathway. These genes/pathways may serve as potential therapeutic targets for PD clinical therapy via modulation of the MGBA in the near future. Thus, an improved understanding of the molecular mechanisms that are involved between the gut microbiota, MGBA and the brain in patients with PD may assist in the development of novel therapeutic strategies for the treatment of PD.

Although several changes in the taxa of the GM have been found in patients with PD, several questions regarding the association between the MGBA and PD remain unanswered. For example, similar changes to the taxa of the GM have been found in other diseases, such as decreased levels of prevotella in type I diabetes and an increased abundance of Lactobacillus in type II diabetes (183,184). Thus, identifying specific PD-associated bacterial taxa as a method to diagnose PD may not be viable. In the future, studies should concentrate on confirming the changes to the bacterial taxa in the intestinal tract that may serve as biomarkers for diagnosis of PD. Furthermore, although several studies have attempted to understand the association between the changes in the GM and the pathogenesis of PD, specific molecular mechanisms regarding the associations between the constituents of the intestinal bacterial taxa and PD remain to be determined. Notably, whether the changes in the bacterial taxa are the cause or consequence of PD should be addressed.

Despite recognizing the pathogeny of PD and the role of abnormal aggregation of α-syn, Horsager et al (185) reported that PD can be separated into two hypotypes based on the initial position of the pathological α-syn aggregation. The initial site of the build-up of α-syn aggregates has not been determined, thus, further studies are required to determine the starting position of α-syn polymers and the associated mechanisms. In addition, in the gastrointestinal tract, why α-syn tends to oligomerization or polymer and its functional roles remain unclear. Barbut et al (186) reported that monomeric α-syn did not exhibit the ability of antimicrobial infection, but aggregated α-syn exhibited a property of antimicrobial peptides (AMPs), which is concentration-dependent aggregation in the gastrointestinal tract of patients with PD. Briefly, the ‘intrinsically disordered’ α-syn assumes no specific structure in the aqueous solvent. However, α-syn exhibits a net of cationic charge by the N-terminal (~60 amino acid residue) in the presence of lipid membranes, and interacts with the phospholipids of lipid membranes that contain an abundance of negatively charged headgroups, such as the phosphatidyl serine. Subsequently, the α-syn is pulled toward the lipid membranes electrostatically, which increases the concentration of α-syn on the lipid membrane. The bound α-syn molecules are closer to one another and begin to aggregate, exhibiting the characteristic of concentration-dependent aggregation, which has been described for numerous AMPs (187,188). Despite the understanding of the process of α-syn aggregation, the downstream targeted molecules of aggregated α-syn in the lipid membrane remain unclear. Goya et al (144) demonstrated that a probiotic strain, B. subtilis PXN21, can effectively clear aggregated α-syn. However, whether probiotic can remove α-syn that has bound to the lipid membrane remains unknown. Abbott et al (189) reported that the unbalance of lipid metabolism is an important pathogenesis of PD. Whether the unbalance of lipid metabolism can increase an affinity between α-syn and lipid membrane requires further investigation. If such means are exposed, it can unravel a new wave of therapies that target PD from the foundational pathophysiology, rather than just suppressing the symptoms.

Acknowledgements

Not applicable.

Funding

The present review was funded by the Innovative Team of Functional Neurosurgery of Kunming Medical University (grant no. CXTD201703).

Availability of data and materials

Not applicable.

Authors' contributions

QL, TW, JY and GS conceived and designed the present review. QL, TW, JW, XH, YG, YW, JY and GS performed the literature review and acquired the relevant data. QL, TW, JY and GS drafted the initial manuscript. Data authentication is not applicable. All authors have read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Wu S, Lei L, Song Y, Liu M, Lu S, Lou D, Shi Y, Wang Z and He D: Mutation of hop-1 and pink-1 attenuates vulnerability of neurotoxicity in C. elegans: The role of mitochondria-associated membrane proteins in Parkinsonism. Exp Neurol. 309:67–78. 2018. View Article : Google Scholar : PubMed/NCBI

2 

Balestrino R and Schapira AHV: Parkinson disease. Eur J Neurol. 27:27–42. 2020. View Article : Google Scholar : PubMed/NCBI

3 

Mahoney-Sanchez L, Bouchaoui H, Ayton S, Devos D, Duce JA and Devedjian JC: Ferroptosis and its potential role in the physiopathology of Parkinson's disease. Prog Neurobiol. 196:1018902021. View Article : Google Scholar : PubMed/NCBI

4 

Samii A, Nutt JG and Ransom BR: Parkinson's disease. Lancet. 363:1783–1793. 2004. View Article : Google Scholar : PubMed/NCBI

5 

Khoo TK, Yarnall AJ, Duncan GW, Coleman S, O'Brien JT, Brooks DJ, Barker RA and Burn DJ: The spectrum of nonmotor symptoms in early Parkinson disease. Neurology. 80:276–281. 2013. View Article : Google Scholar : PubMed/NCBI

6 

Lopiano L, Modugno N, Marano P, Sensi M, Meco G, Cannas A, Gusmaroli G, Tamma F, Mancini F, Quatrale R, et al: Motor outcomes in patients with advanced Parkinson's disease treated with levodopa/carbidopa intestinal gel in Italy: An interim analysis from the GREENFIELD observational study. Neurol Sci. 37:1785–1792. 2016. View Article : Google Scholar : PubMed/NCBI

7 

Kalinderi K, Bostantjopoulou S and Fidani L: The genetic background of Parkinson's disease: Current progress and future prospects. Acta Neurol Scand. 134:314–326. 2016. View Article : Google Scholar : PubMed/NCBI

8 

Malpartida AB, Williamson M, Narendra DP, Wade-Martins R and Ryan BJ: Mitochondrial dysfunction and mitophagy in Parkinson's disease: From mechanism to therapy. Trends Biochem Sci. 46:329–343. 2021. View Article : Google Scholar : PubMed/NCBI

9 

Tysnes OB and Storstein A: Epidemiology of Parkinson's disease. J Neural Transm (Vienna). 124:901–905. 2017. View Article : Google Scholar : PubMed/NCBI

10 

Ascherio A and Schwarzschild MA: The epidemiology of Parkinson's disease: Risk factors and prevention. Lancet Neurol. 15:1257–1272. 2016. View Article : Google Scholar : PubMed/NCBI

11 

Ahier A, Dai CY, Kirmes I, Cummins N, Hung GCC, Götz J and Zuryn S: PINK1 and parkin shape the organism-wide distribution of a deleterious mitochondrial genome. Cell Rep. 35:1092032021. View Article : Google Scholar : PubMed/NCBI

12 

Tolosa E, Vila M, Klein C and Rascol O: LRRK2 in Parkinson disease: Challenges of clinical trials. Nat Rev Neurol. 16:97–107. 2020. View Article : Google Scholar : PubMed/NCBI

13 

Wauters F, Cornelissen T, Imberechts D, Martin S, Koentjoro B, Sue C, Vangheluwe P and Vandenberghe W: LRRK2 mutations impair depolarization-induced mitophagy through inhibition of mitochondrial accumulation of RAB10. Autophagy. 16:203–222. 2020. View Article : Google Scholar : PubMed/NCBI

14 

Terbeek J, Martin S, Imberechts D, Kinnart I, Vangheluwe P, Nicholl D and Vandenberghe W: Increased superoxide in GCH1 mutant fibroblasts points to a dopamine-independent toxicity mechanism. Parkinsonism Relat Disord. 82:10–12. 2021. View Article : Google Scholar : PubMed/NCBI

15 

Imai Y, Meng H, Shiba-Fukushima K and Hattori N: Twin CHCH proteins, CHCHD2, and CHCHD10: Key molecules of Parkinson's disease, amyotrophic lateral sclerosis, and frontotemporal dementia. Int J Mol Sci. 20:9082019. View Article : Google Scholar : PubMed/NCBI

16 

Sassone J, Reale C, Dati G, Regoni M, Pellecchia MT and Garavaglia B: The role of VPS35 in the pathobiology of Parkinson's disease. Cell Mol Neurobiol. 41:199–227. 2021. View Article : Google Scholar : PubMed/NCBI

17 

Gangemi S, Gofita E, Costa C, Teodoro M, Briguglio G, Nikitovic D, Tzanakakis G, Tsatsakis AM, Wilks MF, Spandidos DA and Fenga C: Occupational and environmental exposure to pesticides and cytokine pathways in chronic diseases (Review). Int J Mol Med. 38:1012–1020. 2016. View Article : Google Scholar : PubMed/NCBI

18 

Teodoro M, Briguglio G, Fenga C and Costa C: Genetic polymorphisms as determinants of pesticide toxicity: Recent advances. Toxicol Rep. 6:564–570. 2019. View Article : Google Scholar : PubMed/NCBI

19 

Costa C, Teodoro M, Rugolo CA, Alibrando C, Giambo F, Briguglio G and Fenga C: MicroRNAs alteration as early biomarkers for cancer and neurodegenerative diseases: New challenges in pesticides exposure. Toxicol Rep. 7:759–767. 2020. View Article : Google Scholar : PubMed/NCBI

20 

Srivastav S, Fatima M and Mondal AC: Important medicinal herbs in Parkinson's disease pharmacotherapy. Biomed Pharmacother. 92:856–863. 2017. View Article : Google Scholar : PubMed/NCBI

21 

Titze-de-Almeida SS, Soto-Sanchez C, Fernandez E, Koprich JB, Brotchie JM and Titze-de-Almeida R: The promise and challenges of developing miRNA-Based therapeutics for Parkinson's disease. Cells. 9:8412020. View Article : Google Scholar : PubMed/NCBI

22 

Yu Y, Chen H, Hua X, Dang Y, Han Y, Yu Z, Chen X, Ding P and Li H: Polystyrene microplastics (PS-MPs) toxicity induced oxidative stress and intestinal injury in nematode caenorhabditis elegans. Sci Total Environ. 726:1386792020. View Article : Google Scholar : PubMed/NCBI

23 

Schirinzi T, Landi D and Liguori C: COVID-19: Dealing with a potential risk factor for chronic neurological disorders. J Neurol. 268:1171–1178. 2021. View Article : Google Scholar : PubMed/NCBI

24 

Ribeiro DE, Oliveira-Giacomelli A, Glaser T, Arnaud-Sampaio VF, Andrejew R, Dieckmann L, Baranova J, Lameu C, Ratajczak MZ and Ulrich H: Hyperactivation of P2X7 receptors as a culprit of COVID-19 neuropathology. Mol Psychiatry. 26:1044–1059. 2021. View Article : Google Scholar : PubMed/NCBI

25 

Fasano A, Visanji NP, Liu LW, Lang AE and Pfeiffer RF: Gastrointestinal dysfunction in Parkinson's disease. Lancet Neurol. 14:625–639. 2015. View Article : Google Scholar : PubMed/NCBI

26 

Mayer EA, Tillisch K and Gupta A: Gut/brain axis and the microbiota. J Clin Invest. 125:926–938. 2015. View Article : Google Scholar : PubMed/NCBI

27 

Ghaisas S, Maher J and Kanthasamy A: Gut microbiome in health and disease: Linking the microbiome-gut-brain axis and environmental factors in the pathogenesis of systemic and neurodegenerative diseases. Pharmacol Ther. 158:52–62. 2016. View Article : Google Scholar : PubMed/NCBI

28 

Felice VD, Quigley EM, Sullivan AM, O'Keeffe GW and O'Mahony SM: Microbiota-gut-brain signalling in Parkinson's disease: Implications for non-motor symptoms. Parkinsonism Relat Disord. 27:1–8. 2016. View Article : Google Scholar : PubMed/NCBI

29 

Pfeiffer R: Beyond here be dragons: SIBO in Parkinson's disease. Mov Disord. 28:1764–1765. 2013. View Article : Google Scholar : PubMed/NCBI

30 

Nielsen HH, Qiu J, Friis S, Wermuth L and Ritz B: Treatment for helicobacter pylori infection and risk of Parkinson's disease in denmark. Eur J Neurol. 19:864–869. 2012. View Article : Google Scholar : PubMed/NCBI

31 

Devos D, Lebouvier T, Lardeux B, Biraud M, Rouaud T, Pouclet H, Coron E, Bruley des Varannes S, Naveilhan P, Nguyen JM, et al: Colonic inflammation in Parkinson's disease. Neurobiol Dis. 50:42–48. 2013. View Article : Google Scholar : PubMed/NCBI

32 

Mehra S, Sahay S and Maji SK: α-synuclein misfolding and aggregation: Implications in Parkinson's disease pathogenesis. Biochim Biophys Acta Proteins Proteom. 1867:890–908. 2019. View Article : Google Scholar : PubMed/NCBI

33 

Li X, Li YH, Han JY, Yu S and Chen B: cDNA cloning, prokaryotic expression and purification of rat alpha-synuclein. Neurosci Bull. 22:29–33. 2006.PubMed/NCBI

34 

Pogorelov VM, Kao HT, Augustine GJ and Wetsel WC: Postsynaptic mechanisms render Syn I/II/III mice highly responsive to psychostimulants. Int J Neuropsychopharmacol. 22:453–465. 2019. View Article : Google Scholar : PubMed/NCBI

35 

Thiel G: Synapsin I, synapsin II, and synaptophysin: Marker proteins of synaptic vesicles. Brain Pathol. 3:87–95. 1993. View Article : Google Scholar : PubMed/NCBI

36 

Zhang L, Zhang C, Zhu Y, Cai Q, Chan P, Uéda K, Yu S and Yang H: Semi-quantitative analysis of alpha-synuclein in subcellular pools of rat brain neurons: An immunogold electron microscopic study using a C-terminal specific monoclonal antibody. Brain Res. 1244:40–52. 2008. View Article : Google Scholar : PubMed/NCBI

37 

Braak H, Rub U, Gai WP and Del Tredici K; Idiopathic Parkinson's disease, : Possible routes by which vulnerable neuronal types may be subject to neuroinvasion by an unknown pathogen. J Neural Transm (Vienna). 110:517–536. 2003. View Article : Google Scholar : PubMed/NCBI

38 

Luk KC, Kehm VM, Zhang B, O'Brien P, Trojanowski JQ and Lee VM: Intracerebral inoculation of pathological alpha-synuclein initiates a rapidly progressive neurodegenerative alpha-synucleinopathy in mice. J Exp Med. 209:975–986. 2012. View Article : Google Scholar : PubMed/NCBI

39 

Zheng H, Shi C, Luo H, Fan L, Yang Z, Hu X, Zhang Z, Zhang S, Hu Z, Fan Y, et al: alpha-synuclein in Parkinson's disease: Does a prion-like mechanism of propagation from periphery to the brain play a role? Neuroscientist. 27:367–387. 2021. View Article : Google Scholar : PubMed/NCBI

40 

Perez-Pardo P, Kliest T, Dodiya HB, Broersen LM, Garssen J, Keshavarzian A and Kraneveld AD: The gut-brain axis in Parkinson's disease: Possibilities for food-based therapies. Eur J Pharmacol. 817:86–95. 2017. View Article : Google Scholar : PubMed/NCBI

41 

Dogra N, Mani RJ and Katare DP: The gut-brain axis: Two ways signaling in Parkinson's disease. Cell Mol Neurobiol. 2:10072021.

42 

Kuo YM, Li Z, Jiao Y, Gaborit N, Pani AK, Orrison BM, Bruneau BG, Giasson BI, Smeyne RJ, Gershon MD and Nussbaum RL: Extensive enteric nervous system abnormalities in mice transgenic for artificial chromosomes containing Parkinson disease-associated alpha-synuclein gene mutations precede central nervous system changes. Hum Mol Genet. 19:1633–1650. 2010. View Article : Google Scholar : PubMed/NCBI

43 

Braak H, Del Tredici K, Rüb U, de Vos RA, Steur EN and Braak E: Staging of brain pathology related to sporadic Parkinson's disease. Neurobiol Aging. 24:197–211. 2003. View Article : Google Scholar : PubMed/NCBI

44 

Borghammer P and Van Den Berge N: Brain-first versus gut-first Parkinson's disease: A hypothesis. J Parkinsons Dis. 9:S281–S295. 2019. View Article : Google Scholar : PubMed/NCBI

45 

Comi C, Magistrelli L, Oggioni GD, Carecchio M, Fleetwood T, Cantello R, Mancini F and Antonini A: Peripheral nervous system involvement in Parkinson's disease: Evidence and controversies. Parkinsonism Relat Disord. 20:1329–1334. 2014. View Article : Google Scholar : PubMed/NCBI

46 

Nair AT, Ramachandran V, Joghee NM, Antony S and Ramalingam G: Gut microbiota dysfunction as reliable non-invasive early diagnostic biomarkers in the pathophysiology of Parkinson's disease: A critical review. J Neurogastroenterol Motil. 24:30–42. 2018. View Article : Google Scholar : PubMed/NCBI

47 

Rani L and Mondal AC: Unravelling the role of gut microbiota in Parkinson's disease progression: Pathogenic and therapeutic implications. Neurosci Res. 168:100–112. 2021. View Article : Google Scholar : PubMed/NCBI

48 

Scheperjans F, Aho V, Pereira PA, Koskinen K, Paulin L, Pekkonen E, Haapaniemi E, Kaakkola S, Eerola-Rautio J, Pohja M, et al: Gut microbiota are related to Parkinson's disease and clinical phenotype. Mov Disord. 30:350–358. 2015. View Article : Google Scholar : PubMed/NCBI

49 

Mulak A and Boaz B: Brain-gut-microbiota axis in Parkinson's disease. World J Gastroenterol. 21:10609–10620. 2015. View Article : Google Scholar : PubMed/NCBI

50 

Cirstea MS, Yu AC, Golz E, Sundvick K, Kliger D, Radisavljevic N, Foulger LH, Mackenzie M, Huan T, Finlay BB and Appel-Cresswell S: Microbiota composition and metabolism are associated with gut function in Parkinson's disease. Mov Disord. 35:1208–1217. 2020. View Article : Google Scholar : PubMed/NCBI

51 

Lee HS, Lobbestael E, Vermeire S, Sabino J and Cleynen I: Inflammatory bowel disease and Parkinson's disease: Common pathophysiological links. Gut. 70:408–417. 2021.PubMed/NCBI

52 

Gerhardt S and Mohajeri MH: Changes of colonic bacterial composition in Parkinson's disease and other neurodegenerative diseases. Nutrients. 10:7082018. View Article : Google Scholar : PubMed/NCBI

53 

Lubomski M, Tan AH, Lim SY, Holmes AJ, Davis RL and Sue CM: Parkinson's disease and the gastrointestinal microbiome. J Neurol. 267:2507–2523. 2020. View Article : Google Scholar : PubMed/NCBI

54 

Keshavarzian A, Green SJ, Engen PA, Voigt RM, Naqib A, Forsyth CB, Mutlu E and Shannon KM: Colonic bacterial composition in Parkinson's disease. Mov Disord. 30:1351–1360. 2015. View Article : Google Scholar : PubMed/NCBI

55 

Baizabal-Carvallo JF and Alonso-Juarez M: The link between gut dysbiosis and neuroinflammation in Parkinson's disease. Neuroscience. 432:160–173. 2020. View Article : Google Scholar : PubMed/NCBI

56 

Li F, Wang P, Chen Z, Sui X, Xie X and Zhang J: Alteration of the fecal microbiota in North-Eastern Han Chinese population with sporadic Parkinson's disease. Neurosci Lett. 707:1342972019. View Article : Google Scholar : PubMed/NCBI

57 

Hill-Burns EM, Debelius JW, Morton JT, Wissemann WT, Lewis MR, Wallen ZD, Peddada SD, Factor SA, Molho E, Zabetian CP, et al: Parkinson's disease and Parkinson's disease medications have distinct signatures of the gut microbiome. Mov Disord. 32:739–749. 2017. View Article : Google Scholar : PubMed/NCBI

58 

Elfil M, Kamel S, Kandil M, Koo BB and Schaefer SM: Implications of the gut microbiome in Parkinson's disease. Movement disorders: Mov Disord. 35:921–933. 2020. View Article : Google Scholar : PubMed/NCBI

59 

Sorrentino ZA, Xia Y, Gorion KM, Hass E and Giasson BI: Carboxy-terminal truncations of mouse α-synuclein alter aggregation and prion-like seeding. FEBS Lett. 594:1271–1283. 2020. View Article : Google Scholar : PubMed/NCBI

60 

Cryan JF, O'Riordan KJ, Cowan CSM, Sandhu KV, Bastiaanssen TFS, Boehme M, Codagnone MG, Cussotto S, Fulling C, Golubeva AV, et al: The microbiota-gut-brain axis. Physiol Rev. 99:1877–2013. 2019. View Article : Google Scholar : PubMed/NCBI

61 

Koszewicz M, Jaroch J, Brzecka A, Ejma M, Budrewicz S, Mikhaleva LM, Muresanu C, Schield P, Somasundaram SG, Kirkland CE, et al: Dysbiosis is one of the risk factor for stroke and cognitive impairment and potential target for treatment. Pharmacol Res. 164:1052772021. View Article : Google Scholar : PubMed/NCBI

62 

Tan AH, Lim SY, Chong KK, Manap AM, Hor JW, Lim JL, Low SC, Chong CW, Mahadeva S and Lang AE: Probiotics for constipation in Parkinson disease: A randomized placebo-controlled study. Neurology. 96:e772–e782. 2021.PubMed/NCBI

63 

Hou YF, Shan C, Zhuang SY, Zhuang QQ, Ghosh A, Zhu KC, Kong XK, Wang SM, Gong YL, Yang YY, et al: Gut microbiota-derived propionate mediates the neuroprotective effect of osteocalcin in a mouse model of Parkinson's disease. Microbiome. 9:342021. View Article : Google Scholar : PubMed/NCBI

64 

Sun MF, Zhu YL, Zhou ZL, Jia XB, Xu YD, Yang Q, Cui C and Shen YQ: Neuroprotective effects of fecal microbiota transplantation on MPTP-induced Parkinson's disease mice: Gut microbiota, glial reaction and TLR4/TNF-α signaling pathway. Brain Behav Immun. 70:48–60. 2018. View Article : Google Scholar : PubMed/NCBI

65 

Kayisoglu O, Weiss F, Niklas C, Pierotti I, Pompaiah M, Wallaschek N, Germer CT, Wiegering A and Bartfeld S: Location-specific cell identity rather than exposure to GI microbiota defines many innate immune signalling cascades in the gut epithelium. Gut. 70:687–697. 2021. View Article : Google Scholar : PubMed/NCBI

66 

Kelly LP, Carvey PM, Keshavarzian A, Shannon KM, Shaikh M, Bakay RA and Kordower JH: Progression of intestinal permeability changes and alpha-synuclein expression in a mouse model of Parkinson's disease. Mov Disord. 29:999–1009. 2014. View Article : Google Scholar : PubMed/NCBI

67 

Uemura N, Yagi H, Uemura MT, Hatanaka Y, Yamakado H and Takahashi R: Inoculation of α-synuclein preformed fibrils into the mouse gastrointestinal tract induces lewy body-like aggregates in the brainstem via the vagus nerve. Mol Neurodegener. 13:212018. View Article : Google Scholar : PubMed/NCBI

68 

Forsyth CB, Shannon KM, Kordower JH, Voigt RM, Shaikh M, Jaglin JA, Estes JD, Dodiya HB and Keshavarzian A: Increased intestinal permeability correlates with sigmoid mucosa alpha-synuclein staining and endotoxin exposure markers in early Parkinson's disease. PLoS One. 6:e280322011. View Article : Google Scholar : PubMed/NCBI

69 

George S and Brundin P: Immunotherapy in Parkinson's disease: Micromanaging alpha-synuclein aggregation. J Parkinsons Dis. 5:413–424. 2015. View Article : Google Scholar : PubMed/NCBI

70 

Wang Y, Liu Y, Sidhu A, Ma Z, McClain C and Feng W: Lactobacillus rhamnosus GG culture supernatant ameliorates acute alcohol-induced intestinal permeability and liver injury. Am J Physiol Gastrointest Liver Physiol. 303:G32–G41. 2012. View Article : Google Scholar : PubMed/NCBI

71 

Bischoff SC, Barbara G, Buurman W, Ockhuizen T, Schulzke JD, Serino M, Tilg H, Watson A and Wells JM: Intestinal permeability-a new target for disease prevention and therapy. BMC Gastroenterol. 14:1892014. View Article : Google Scholar : PubMed/NCBI

72 

Ploger S, Stumpff F, Penner GB, Schulzke JD, Gäbel G, Martens H, Shen Z, Günzel D and Aschenbach JR: Microbial butyrate and its role for barrier function in the gastrointestinal tract. Ann N Y Acad Sci. 1258:52–59. 2012. View Article : Google Scholar : PubMed/NCBI

73 

Unger MM, Spiegel J, Dillmann KU, Grundmann D, Philippeit H, Burmann J, Faßbender K, Schwiertz A and Schäfer KH: Short chain fatty acids and gut microbiota differ between patients with Parkinson's disease and age-matched controls. Parkinsonism Relat Disord. 32:66–72. 2016. View Article : Google Scholar : PubMed/NCBI

74 

Zhou X, Zhang B, Zhao X, Lin Y, Wang J, Wang X, Hu N and Wang S: Chlorogenic acid supplementation ameliorates hyperuricemia, relieves renal inflammation, and modulates intestinal homeostasis. Food Funct. 12:5637–5649. 2021. View Article : Google Scholar : PubMed/NCBI

75 

Chen ZJ, Liang CY, Yang LQ, Ren SM, Xia YM, Cui L, Li XF and Gao BL: Association of Parkinson's disease with microbes and microbiological therapy. Front Cell Infect Microbiol. 11:6193542021. View Article : Google Scholar : PubMed/NCBI

76 

Watson AJ and Hughes KR: TNF-α-induced intestinal epithelial cell shedding: Implications for intestinal barrier function. Ann N Y Acad Sci. 1258:1–8. 2012. View Article : Google Scholar : PubMed/NCBI

77 

Resnikoff H, Metzger JM, Lopez M, Bondarenko V, Mejia A, Simmons HA and Emborg ME: Colonic inflammation affects myenteric alpha-synuclein in nonhuman primates. J Inflamm Res. 12:113–126. 2019. View Article : Google Scholar : PubMed/NCBI

78 

Choi JG, Kim N, Ju IG, Eo H, Lim SM, Jang SE, Kim DH and Oh MS: Oral administration of proteus mirabilis damages dopaminergic neurons and motor functions in mice. Sci Rep. 8:12752018. View Article : Google Scholar : PubMed/NCBI

79 

Brown GC: The endotoxin hypothesis of neurodegeneration. J Neuroinflammation. 16:1802019. View Article : Google Scholar : PubMed/NCBI

80 

Bhattacharyya D, Mohite GM, Krishnamoorthy J, Gayen N, Mehra S, Navalkar A, Kotler SA, Ratha BN, Ghosh A, Kumar R, et al: Lipopolysaccharide from gut microbiota modulates α-synuclein aggregation and alters its biological function. ACS Chem Neurosci. 10:2229–2236. 2019. View Article : Google Scholar : PubMed/NCBI

81 

Huang C, Zhu L, Li H, Shi FG, Wang GQ, Wei YZ, Liu J and Zhang F: Adulthood exposure to lipopolysaccharide exacerbates the neurotoxic and inflammatory effects of rotenone in the substantia nigra. Front Mol Neurosci. 10:1312017. View Article : Google Scholar : PubMed/NCBI

82 

Wang W, Nguyen LT, Burlak C, Chegini F, Guo F, Chataway T, Ju S, Fisher OS, Miller DW, Datta D, et al: Caspase-1 causes truncation and aggregation of the Parkinson's disease-associated protein alpha-synuclein. Proc Natl Acad Sci USA. 113:9587–9592. 2016. View Article : Google Scholar : PubMed/NCBI

83 

Terada M, Suzuki G, Nonaka T, Kametani F, Tamaoka A and Hasegawa M: The effect of truncation on prion-like properties of α-synuclein. J Biol Chem. 293:13910–13920. 2018. View Article : Google Scholar : PubMed/NCBI

84 

Woerman AL, Kazmi SA, Patel S, Freyman Y, Oehler A, Aoyagi A, Mordes DA, Halliday GM, Middleton LT, Gentleman SM, et al: MSA prions exhibit remarkable stability and resistance to inactivation. Acta Neuropathol. 135:49–63. 2018. View Article : Google Scholar : PubMed/NCBI

85 

Davie CA: A review of Parkinson's disease. Br Med Bull. 86:109–127. 2008. View Article : Google Scholar : PubMed/NCBI

86 

Prusiner SB: Cell biology. A unifying role for prions in neurodegenerative diseases. Science. 336:1511–1513. 2012. View Article : Google Scholar : PubMed/NCBI

87 

Bhattacharyya D and Bhunia A: Gut-brain axis in Parkinson's disease etiology: The role of lipopolysaccharide. Chem Phys Lipids. 235:1050292021. View Article : Google Scholar : PubMed/NCBI

88 

Lee SJ, Desplats P, Sigurdson C, Tsigelny I and Masliah E: Cell-to-cell transmission of non-prion protein aggregates. Nat Rev Neurol. 6:702–706. 2010. View Article : Google Scholar : PubMed/NCBI

89 

Luk KC, Kehm V, Carroll J, Zhang B, O'Brien P, Trojanowski JQ and Lee VM: Pathological α-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science. 338:949–953. 2012. View Article : Google Scholar : PubMed/NCBI

90 

Pan-Montojo F, Schwarz M, Winkler C, Arnhold M, O'Sullivan GA, Pal A, Said J, Marsico G, Verbavatz JM, Rodrigo-Angulo M, et al: Environmental toxins trigger PD-like progression via increased alpha-synuclein release from enteric neurons in mice. Sci Rep. 2:8982012. View Article : Google Scholar : PubMed/NCBI

91 

Mezias C, Rey N, Brundin P and Raj A: Neural connectivity predicts spreading of alpha-synuclein pathology in fibril-injected mouse models: Involvement of retrograde and anterograde axonal propagation. Neurobiol Dis. 134:1046232020. View Article : Google Scholar : PubMed/NCBI

92 

Holmqvist S, Chutna O, Bousset L, Aldrin-Kirk P, Li W, Björklund T, Wang ZY, Roybon L, Melki R and Li JY: Direct evidence of Parkinson pathology spread from the gastrointestinal tract to the brain in rats. Acta Neuropathol. 128:805–820. 2014. View Article : Google Scholar : PubMed/NCBI

93 

Kim S, Kwon SH, Kam TI, Panicker N, Karuppagounder SS, Lee S, Lee JH, Kim WR, Kook M, Foss CA, et al: Transneuronal propagation of pathologic alpha-synuclein from the gut to the brain models Parkinson's disease. Neuron. 103:627–641. 2019. View Article : Google Scholar : PubMed/NCBI

94 

Braak H, Sastre M, Bohl JR, de Vos RA and Del Tredici K: Parkinson's disease: Lesions in dorsal horn layer I, involvement of parasympathetic and sympathetic pre- and postganglionic neurons. Acta Neuropathol. 113:421–429. 2007. View Article : Google Scholar : PubMed/NCBI

95 

Dodiya HB, Forsyth CB, Voigt RM, Engen PA, Patel J, Shaikh M, Green SJ, Naqib A, Roy A, Kordower JH, et al: Chronic stress-induced gut dysfunction exacerbates Parkinson's disease phenotype and pathology in a rotenone-induced mouse model of Parkinson's disease. Neurobiol Dis. 135:1043522020. View Article : Google Scholar : PubMed/NCBI

96 

Santos SF, de Oliveira HL, Yamada ES, Neves BC and Pereira A Jr: The gut and Parkinson's disease-a bidirectional pathway. Front Neurol. 10:5742019. View Article : Google Scholar : PubMed/NCBI

97 

Hsu LJ, Sagara Y, Arroyo A, Rockenstein E, Sisk A, Mallory M, Wong J, Takenouchi T, Hashimoto M and Masliah E: alpha-synuclein promotes mitochondrial deficit and oxidative stress. Am J Pathol. 157:401–410. 2000. View Article : Google Scholar : PubMed/NCBI

98 

Hu D, Sun X, Liao X, Zhang X, Zarabi S, Schimmer A, Hong Y, Ford C, Luo Y and Qi X: Alpha-synuclein suppresses mitochondrial protease ClpP to trigger mitochondrial oxidative damage and neurotoxicity. Acta Neuropathol. 137:939–960. 2019. View Article : Google Scholar : PubMed/NCBI

99 

Chong W, Jimenez J, Mc IM, Saito MA and Kwakye GF: α-Synuclein enhances cadmium uptake and neurotoxicity via oxidative stress and caspase activated cell death mechanisms in a dopaminergic cell model of Parkinson's disease. Neurotox Res. 32:231–246. 2017. View Article : Google Scholar : PubMed/NCBI

100 

Dryanovski DI, Guzman JN, Xie Z, Galteri DJ, Volpicelli-Daley LA, Lee VM, Miller RJ, Schumacker PT and Surmeier DJ: Calcium entry and α-synuclein inclusions elevate dendritic mitochondrial oxidant stress in dopaminergic neurons. J Neurosci. 33:10154–10164. 2013. View Article : Google Scholar : PubMed/NCBI

101 

Tapias V, Hu X, Luk KC, Sanders LH, Lee VM and Greenamyre JT: Synthetic alpha-synuclein fibrils cause mitochondrial impairment and selective dopamine neurodegeneration in part via iNOS-mediated nitric oxide production. Cell Mol Life Sci. 74:2851–2874. 2017. View Article : Google Scholar : PubMed/NCBI

102 

Musgrove RE, Helwig M, Bae EJ, Aboutalebi H, Lee SJ, Ulusoy A and Di Monte DA: Oxidative stress in vagal neurons promotes parkinsonian pathology and intercellular alpha-synuclein transfer. J Clin Invest. 129:3738–3753. 2019. View Article : Google Scholar : PubMed/NCBI

103 

Martin LJ, Pan Y, Price AC, Sterling W, Copeland NG, Jenkins NA, Price DL and Lee MK: Parkinson's disease alpha-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J Neurosci. 26:41–50. 2006. View Article : Google Scholar : PubMed/NCBI

104 

Stichel CC, Zhu XR, Bader V, Linnartz B, Schmidt S and Lübbert H: Mono- and double-mutant mouse models of Parkinson's disease display severe mitochondrial damage. Hum Mol Genet. 16:2377–2393. 2007. View Article : Google Scholar : PubMed/NCBI

105 

Ding H, Xiong Y, Sun J, Chen C, Gao J and Xu H: Asiatic acid prevents oxidative stress and apoptosis by inhibiting the translocation of α-synuclein into mitochondria. Front Neurosci. 12:4312018. View Article : Google Scholar : PubMed/NCBI

106 

Ho MS: Microglia in Parkinson's disease. Adv Exp Med Biol. 1175:335–353. 2019. View Article : Google Scholar : PubMed/NCBI

107 

Harms AS, Delic V, Thome AD, Bryant N, Liu Z, Chandra S, Jurkuvenaite A and West AB: α-synuclein fibrils recruit peripheral immune cells in the rat brain prior to neurodegeneration. Neuropathol Commun. 5:852017. View Article : Google Scholar : PubMed/NCBI

108 

Politis M, Su P and Piccini P: Imaging of microglia in patients with neurodegenerative disorders. Front Pharmacol. 3:962012. View Article : Google Scholar : PubMed/NCBI

109 

Joers V, Tansey MG, Mulas G and Carta AR: Microglial phenotypes in Parkinson's disease and animal models of the disease. Prog Neurobiol. 155:57–75. 2017. View Article : Google Scholar : PubMed/NCBI

110 

Whitton PS: Inflammation as a causative factor in the aetiology of Parkinson's disease. Br J Pharmacol. 150:963–976. 2007. View Article : Google Scholar : PubMed/NCBI

111 

Zhang YN, Fan JK, Gu L, Yang HM, Zhan SQ and Zhang H: Metabotropic glutamate receptor 5 inhibits α-synuclein-induced microglia inflammation to protect from neurotoxicity in Parkinson's disease. J Neuroinflammation. 18:232021. View Article : Google Scholar : PubMed/NCBI

112 

Croisier E, Moran LB, Dexter DT, Pearce RK and Graeber MB: Microglial inflammation in the parkinsonian substantia nigra: Relationship to alpha-synuclein deposition. J Neuroinflammation. 2:142005. View Article : Google Scholar : PubMed/NCBI

113 

Williams GP, Marmion DJ, Schonhoff AM, Jurkuvenaite A, Won WJ, Standaert DG, Kordower JH and Harms AS: T cell infiltration in both human multiple system atrophy and a novel mouse model of the disease. Acta Neuropathol. 139:855–874. 2020. View Article : Google Scholar : PubMed/NCBI

114 

Roodveldt C, Labrador-Garrido A, Gonzalez-Rey E, Fernandez-Montesinos R, Caro M, Lachaud CC, Waudby CA, Delgado M, Dobson CM and Pozo D: Glial innate immunity generated by non-aggregated alpha-synuclein in mouse: Differences between wild-type and Parkinson's disease-linked mutants. PLoS One. 5:e134812010. View Article : Google Scholar : PubMed/NCBI

115 

Vaure C and Liu Y: A comparative review of toll-like receptor 4 expression and functionality in different animal species. Front Immunol. 5:3162014. View Article : Google Scholar : PubMed/NCBI

116 

Uhlén M, Fagerberg L, Hallström BM, Lindskog C, Oksvold P, Mardinoglu A, Sivertsson A, Kampf C, Sjöstedt E, Asplund A, et al: Proteomics. Tissue-based map of the human proteome. Science. 347:12604192015. View Article : Google Scholar : PubMed/NCBI

117 

Wardill HR, Van Sebille YZ, Mander KA, Gibson RJ, Logan RM, Bowen JM and Sonis ST: Toll-like receptor 4 signaling: A common biological mechanism of regimen-related toxicities: An emerging hypothesis for neuropathy and gastrointestinal toxicity. Cancer Treat Rev. 41:122–128. 2015. View Article : Google Scholar : PubMed/NCBI

118 

Rahimifard M, Maqbool F, Moeini-Nodeh S, Niaz K, Abdollahi M, Braidy N, Nabavi SM and Nabavi SF: Targeting the TLR4 signaling pathway by polyphenols: A novel therapeutic strategy for neuroinflammation. Ageing Res Rev. 36:11–19. 2017. View Article : Google Scholar : PubMed/NCBI

119 

Fellner L, Irschick R, Schanda K, Reindl M, Klimaschewski L, Poewe W, Wenning GK and Stefanova N: Toll-like receptor 4 is required for α-synuclein dependent activation of microglia and astroglia. Glia. 61:349–360. 2013. View Article : Google Scholar : PubMed/NCBI

120 

Mariucci G, Pagiotti R, Galli F, Romani L and Conte C: The potential role of toll-like receptor 4 in mediating dopaminergic cell loss and alpha-synuclein expression in the acute MPTP mouse model of Parkinson's disease. J Mol Neurosci. 64:611–618. 2018. View Article : Google Scholar : PubMed/NCBI

121 

Choi I, Zhang Y, Seegobin SP, Pruvost M, Wang Q, Purtell K, Zhang B and Yue Z: Microglia clear neuron-released α-synuclein via selective autophagy and prevent neurodegeneration. Nat Commun. 11:13862020. View Article : Google Scholar : PubMed/NCBI

122 

Wilms H, Rosenstiel P, Romero-Ramos M, Arlt A, Schäfer H, Seegert D, Kahle PJ, Odoy S, Claasen JH, Holzknecht C, et al: Suppression of MAP kinases inhibits microglial activation and attenuates neuronal cell death induced by alpha-synuclein protofibrils. Int J Immunopathol Pharmacol. 22:897–909. 2009. View Article : Google Scholar : PubMed/NCBI

123 

Lee EJ, Woo MS, Moon PG, Baek MC, Choi IY, Kim WK, Junn E and Kim HS: Alpha-synuclein activates microglia by inducing the expressions of matrix metalloproteinases and the subsequent activation of protease-activated receptor-1. J Immunol. 185:615–623. 2010. View Article : Google Scholar : PubMed/NCBI

124 

Zhang W, Dallas S, Zhang D, Guo JP, Pang H, Wilson B, Miller DS, Chen B, Zhang W, McGeer PL, et al: Microglial PHOX and Mac-1 are essential to the enhanced dopaminergic neurodegeneration elicited by A30P and A53T mutant alpha-synuclein. Glia. 55:1178–1188. 2007. View Article : Google Scholar : PubMed/NCBI

125 

Lastres-Becker I, Ulusoy A, Innamorato NG, Sahin G, Rábano A, Kirik D and Cuadrado A: α-Synuclein expression and Nrf2 deficiency cooperate to aggravate protein aggregation, neuronal death and inflammation in early-stage Parkinson's disease. Hum Mol Genet. 21:3173–3192. 2012. View Article : Google Scholar : PubMed/NCBI

126 

Sanchez-Guajardo V, Tentillier N and Romero-Ramos M: The relation between α-synuclein and microglia in Parkinson's disease: Recent developments. Neuroscience. 302:47–58. 2015. View Article : Google Scholar : PubMed/NCBI

127 

Shavali S, Combs CK and Ebadi M: Reactive macrophages increase oxidative stress and alpha-synuclein nitration during death of dopaminergic neuronal cells in co-culture: Relevance to Parkinson's disease. Neurochem Res. 31:85–94. 2006. View Article : Google Scholar : PubMed/NCBI

128 

Stefanova N, Fellner L, Reindl M, Masliah E, Poewe W and Wenning GK: Toll-like receptor 4 promotes alpha-synuclein clearance and survival of nigral dopaminergic neurons. Am J Pathol. 179:954–963. 2011. View Article : Google Scholar : PubMed/NCBI

129 

Campos SS, Alza NP and Salvador GA: Lipid metabolism alterations in the neuronal response to A53T α-synuclein and Fe-induced injury. Arch Biochem Biophys. 655:43–54. 2018. View Article : Google Scholar : PubMed/NCBI

130 

Liao JF, Cheng YF, You ST, Kuo WC, Huang CW, Chiou JJ, Hsu CC, Hsieh-Li HM, Wang S and Tsai YC: Lactobacillus plantarum PS128 alleviates neurodegenerative progression in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced mouse models of Parkinson's disease. Brain Behav Immun. 90:26–46. 2020. View Article : Google Scholar : PubMed/NCBI

131 

Martin WRW, Miles M, Zhong Q, Hartlein J, Racette BA, Norris SA, Ushe M, Maiti B, Criswell S, Davis AA, et al: Is levodopa response a valid indicator of parkinson's disease? Mov Disord. 36:948–954. 2021. View Article : Google Scholar : PubMed/NCBI

132 

Tomlinson CL, Stowe R, Patel S, Rick C, Gray R and Clarke CE: Systematic review of levodopa dose equivalency reporting in Parkinson's disease. Mov Disord. 25:2649–2653. 2010. View Article : Google Scholar : PubMed/NCBI

133 

Epprecht L, Schreglmann SR, Goetze O, Woitalla D, Baumann CR and Waldvogel D: Unchanged gastric emptying and visceral perception in early Parkinson's disease after a high caloric test meal. J Neurol. 262:1946–1953. 2015. View Article : Google Scholar : PubMed/NCBI

134 

Schrag A and Quinn N: Dyskinesias and motor fluctuations in Parkinson's disease. A community-based study. Brain. 123((Pt 11)): 2297–2305. 2000. View Article : Google Scholar : PubMed/NCBI

135 

Nonnekes J, Timmer MH, de Vries NM, Rascol O, Helmich RC and Bloem BR: Unmasking levodopa resistance in Parkinson's disease. Mov Disord. 31:1602–1609. 2016. View Article : Google Scholar : PubMed/NCBI

136 

Zhu H, Xu G, Zhang K, Kong X, Han R, Zhou J and Ni Y: Crystal structure of tyrosine decarboxylase and identification of key residues involved in conformational swing and substrate binding. Sci Rep. 6:277792016. View Article : Google Scholar : PubMed/NCBI

137 

Rekdal VM, Bess EN, Bisanz JE, Turnbaugh PJ and Balskus EP: Discovery and inhibition of an interspecies gut bacterial pathway for levodopa metabolism. Science. 364:eaau63232019. View Article : Google Scholar

138 

van Kessel SP, Frye AK, El-Gendy AO, Castejon M, Keshavarzian A, van Dijk G and El Aidy S: Gut bacterial tyrosine decarboxylases restrict levels of levodopa in the treatment of Parkinson's disease. Nat Commun. 10:3102019. View Article : Google Scholar : PubMed/NCBI

139 

Sampson TR, Debelius JW, Thron T, Janssen S, Shastri GG, Ilhan ZE, Challis C, Schretter CE, Rocha S, Gradinaru V, et al: Gut microbiota regulate motor deficits and neuroinflammation in a model of Parkinson's disease. Cell. 167:1469–1480. 2016. View Article : Google Scholar : PubMed/NCBI

140 

Varankovich NV, Nickerson MT and Korber DR: Probiotic-based strategies for therapeutic and prophylactic use against multiple gastrointestinal diseases. Front Microbiol. 6:6852015. View Article : Google Scholar : PubMed/NCBI

141 

Liu Y, Yin F, Huang L, Teng H, Shen T and Qin H: Long-term and continuous administration of Bacillus subtilis during remission effectively maintains the remission of inflammatory bowel disease by protecting intestinal integrity, regulating epithelial proliferation, and reshaping microbial structure and function. Food Funct. 12:2201–2210. 2021. View Article : Google Scholar : PubMed/NCBI

142 

Ait-Belgnaoui A, Durand H, Cartier C, Chaumaz G, Eutamene H, Ferrier L, Houdeau E, Fioramonti J, Bueno L and Theodorou V: Prevention of gut leakiness by a probiotic treatment leads to attenuated HPA response to an acute psychological stress in rats. Psychoneuroendocrinology. 37:1885–1895. 2012. View Article : Google Scholar : PubMed/NCBI

143 

Evrensel A and Ceylan ME: Fecal microbiota transplantation and its usage in neuropsychiatric disorders. Clin Psychopharm Neu. 14:231–237. 2016.PubMed/NCBI

144 

Goya ME, Xue F, Sampedro-Torres-Quevedo C, Arnaouteli S, Riquelme-Dominguez L, Romanowski A, Brydon J, Ball KL, Stanley-Wall NR and Doitsidou M: Probiotic bacillus subtilis protects against α-synuclein aggregation in C. elegans. Cell Rep. 30:367–380. 2020. View Article : Google Scholar : PubMed/NCBI

145 

Metta V, Leta V, Mrudula KR, Prashanth LK, Goyal V, Borgohain R, Chung-Faye G and Chaudhuri KR: Gastrointestinal dysfunction in Parkinson's disease: Molecular pathology and implications of gut microbiome, probiotics, and fecal microbiota transplantation. J Neurol. 21:10072021.

146 

Khoruts A: Faecal microbiota transplantation in 2013: Developing human gut microbiota as a class of therapeutics. Nat Rev Gastroenterol Hepatol. 11:79–80. 2014. View Article : Google Scholar : PubMed/NCBI

147 

Liang S, Wang T, Hu X, Luo J, Li W, Wu X, Duan Y and Jin F: Administration of Lactobacillus helveticus NS8 improves behavioral, cognitive, and biochemical aberrations caused by chronic restraint stress. Neuroscience. 310:561–577. 2015. View Article : Google Scholar : PubMed/NCBI

148 

Vendrik KEW, Ooijevaar RE, de Jong PRC, Laman JD, van Oosten BW, van Hilten JJ, Ducarmon QR, Keller JJ, Kuijper EJ and Contarino MF: Fecal microbiota transplantation in neurological disorders. Front Cell Infect Microbiol. 10:982020. View Article : Google Scholar : PubMed/NCBI

149 

Fang X: Microbial treatment: The potential application for Parkinson's disease. Neurol Sci. 40:51–58. 2019. View Article : Google Scholar : PubMed/NCBI

150 

Gibson GR, Hutkins R, Sanders ME, Prescott SL, Reimer RA, Salminen SJ, Scott K, Stanton C, Swanson KS, Cani PD, et al: Expert consensus document: The international scientific association for probiotics and prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat Rev Gastroenterol Hepatol. 14:491–502. 2017. View Article : Google Scholar : PubMed/NCBI

151 

Gibson GR and Roberfroid MB: Dietary modulation of the human colonic microbiota: Introducing the concept of prebiotics. J Nutr. 125:1401–1412. 1995. View Article : Google Scholar : PubMed/NCBI

152 

Kovacs Z, Benjamins E, Grau K, Ur Rehman A, Ebrahimi M and Czermak P: Recent developments in manufacturing oligosaccharides with prebiotic functions. Adv Biochem Eng Biotechnol. 143:257–295. 2014.PubMed/NCBI

153 

Savignac HM, Corona G, Mills H, Chen L, Spencer JP, Tzortzis G and Burnet PW: Prebiotic feeding elevates central brain derived neurotrophic factor, N-methyl-D-aspartate receptor subunits and D-serine. Neurochem Int. 63:756–764. 2013. View Article : Google Scholar : PubMed/NCBI

154 

Armstrong MJ and Okun MS: Diagnosis and treatment of Parkinson disease: A review. JAMA. 323:548–560. 2020. View Article : Google Scholar : PubMed/NCBI

155 

Emamzadeh FN and Surguchov A: Parkinson's disease: Biomarkers, treatment, and risk factors. Front Neurosci. 12:6122018. View Article : Google Scholar : PubMed/NCBI

156 

Takahashi M, Suzuki M, Fukuoka M, Fujikake N, Watanabe S, Murata M, Wada K, Nagai Y and Hohjoh H: Normalization of overexpressed α-synuclein causing Parkinson's disease by a moderate gene silencing with RNA interference. Mol Ther Nucleic Acids. 4:e2412015. View Article : Google Scholar : PubMed/NCBI

157 

Ding L, Gu H, Xiong X, Ao H, Cao J, Lin W, Yu M, Lin J and Cui Q: MicroRNAs involved in carcinogenesis, prognosis, therapeutic resistance and applications in human triple-negative breast cancer. Cells. 8:14922019. View Article : Google Scholar : PubMed/NCBI

158 

Junn E, Lee KW, Jeong BS, Chan TW, Im JY and Mouradian MM: Repression of alpha-synuclein expression and toxicity by microRNA-7. Proc Natl Acad Sci USA. 106:13052–13057. 2009. View Article : Google Scholar : PubMed/NCBI

159 

Masliah E, Rockenstein E, Adame A, Alford M, Crews L, Hashimoto M, Seubert P, Lee M, Goldstein J, Chilcote T, et al: Effects of alpha-synuclein immunization in a mouse model of Parkinson's disease. Neuron. 46:857–868. 2005. View Article : Google Scholar : PubMed/NCBI

160 

Ghochikyan A, Petrushina I, Davtyan H, Hovakimyan A, Saing T, Davtyan A, Cribbs DH and Agadjanyan MG: Immunogenicity of epitope vaccines targeting different B cell antigenic determinants of human alpha-synuclein: Feasibility study. Neurosci Lett. 560:86–91. 2014. View Article : Google Scholar : PubMed/NCBI

161 

Vaikath NN, Hmila I, Gupta V, Erskine D, Ingelsson M and El-Agnaf OMA: Antibodies against alpha-synuclein: Tools and therapies. J Neurochem. 150:612–625. 2019. View Article : Google Scholar : PubMed/NCBI

162 

Rabenstein M, Agbo DB, Wolf E, Dams J, Nicolai M, Roeder A, Bacher M, Dodel RC and Noelker C: Effect of naturally occurring α-synuclein-antibodies on toxic alpha-synuclein-fragments. Neurosci Lett. 704:181–188. 2019. View Article : Google Scholar : PubMed/NCBI

163 

Wang Z, Gao G, Duan C and Yang H: Progress of immunotherapy of anti-alpha-synuclein in Parkinson's disease. Biomed Pharmacother. 115:1088432019. View Article : Google Scholar : PubMed/NCBI

164 

Lashuel HA, Overk CR, Oueslati A and Masliah E: The many faces of alpha-synuclein: From structure and toxicity to therapeutic target. Nat Rev Neurosci. 14:38–48. 2013. View Article : Google Scholar : PubMed/NCBI

165 

Sulzer D, Alcalay RN, Garretti F, Cote L, Kanter E, Agin-Liebes J, Liong C, McMurtrey C, Hildebrand WH, Mao X, et al: T cells from patients with Parkinson's disease recognize α-synuclein peptides. Nature. 546:656–661. 2017. View Article : Google Scholar : PubMed/NCBI

166 

Tran HT, Chung CH, Iba M, Zhang B, Trojanowski JQ, Luk KC and Lee VM: Alpha-synuclein immunotherapy blocks uptake and templated propagation of misfolded alpha-synuclein and neurodegeneration. Cell Rep. 7:2054–2065. 2014. View Article : Google Scholar : PubMed/NCBI

167 

Zhou C, Emadi S, Sierks MR and Messer A: A human single-chain Fv intrabody blocks aberrant cellular effects of overexpressed alpha-synuclein. Mol Ther. 10:1023–1031. 2004. View Article : Google Scholar : PubMed/NCBI

168 

El-Agnaf O, Overk C, Rockenstein E, Mante M, Florio J, Adame A, Vaikath N, Majbour N, Lee SJ, Kim C, et al: Differential effects of immunotherapy with antibodies targeting alpha-synuclein oligomers and fibrils in a transgenic model of synucleinopathy. Neurobiol Dis. 104:85–96. 2017. View Article : Google Scholar : PubMed/NCBI

169 

Rutherford NJ, Brooks M and Giasson BI: Novel antibodies to phosphorylated α-synuclein serine 129 and NFL serine 473 demonstrate the close molecular homology of these epitopes. Acta Neuropathol Commun. 4:802016. View Article : Google Scholar : PubMed/NCBI

170 

Wang N, Garcia J, Freeman R and Gibbons CH: Phosphorylated alpha-synuclein within cutaneous autonomic nerves of patients with Parkinson's disease: The implications of sample thickness on results. J Histochem Cytochem. 68:669–678. 2020. View Article : Google Scholar : PubMed/NCBI

171 

Ciesielska A, Matyjek M and Kwiatkowska K: TLR4 and CD14 trafficking and its influence on LPS-induced pro-inflammatory signaling. Cell Mol Life Sci. 78:1233–1261. 2021. View Article : Google Scholar : PubMed/NCBI

172 

Perez-Pardo P, Dodiya HB, Engen PA, Forsyth CB, Huschens AM, Shaikh M, Voigt RM, Naqib A, Green SJ, Kordower JH, et al: Role of TLR4 in the gut-brain axis in Parkinson's disease: A translational study from men to mice. Gut. 68:829–843. 2019. View Article : Google Scholar : PubMed/NCBI

173 

Brundin L, Bryleva EY and Rajamani KT: Role of inflammation in suicide: From mechanisms to treatment. Neuropsychopharmacology. 42:271–283. 2017. View Article : Google Scholar : PubMed/NCBI

174 

Decressac M, Mattsson B, Weikop P, Lundblad M, Jakobsson J and Bjorklund A: TFEB-mediated autophagy rescues midbrain dopamine neurons from α-synuclein toxicity. Proc Natl Acad Sci USA. 110:E1817–E1826. 2013. View Article : Google Scholar : PubMed/NCBI

175 

Maiese K, Chong ZZ, Shang YC and Wang S: mTOR: On target for novel therapeutic strategies in the nervous system. Trends Mol Med. 19:51–60. 2013. View Article : Google Scholar : PubMed/NCBI

176 

Sardi SP, Cedarbaum JM and Brundin P: Targeted therapies for Parkinson's Disease: From genetics to the clinic. Mov Disord. 33:684–696. 2018. View Article : Google Scholar : PubMed/NCBI

177 

Ghosh A, Tyson T, George S, Hildebrandt EN, Steiner JA, Madaj Z, Schulz E, Machiela E, McDonald WG, Escobar Galvis ML, et al: Mitochondrial pyruvate carrier regulates autophagy, inflammation, and neurodegeneration in experimental models of Parkinson's disease. Sci Transl Med. 8:368ra1742016. View Article : Google Scholar : PubMed/NCBI

178 

Erlich S, Shohami E and Pinkas-Kramarski R: Neurodegeneration induces upregulation of beclin 1. Autophagy. 2:49–51. 2006. View Article : Google Scholar : PubMed/NCBI

179 

Spencer B, Potkar R, Trejo M, Rockenstein E, Patrick C, Gindi R, Adame A, Wyss-Coray T and Masliah E: Beclin 1 gene transfer activates autophagy and ameliorates the neurodegenerative pathology in alpha-synuclein models of Parkinson's and lewy body diseases. J Neurosci. 29:13578–13588. 2009. View Article : Google Scholar : PubMed/NCBI

180 

Savitt D and Jankovic J: Targeting α-synuclein in Parkinson's disease: Progress towards the development of disease-modifying therapeutics. Drugs. 79:797–810. 2019. View Article : Google Scholar : PubMed/NCBI

181 

Hussain T, Zhao D, Shah SZA, Sabir N, Wang J, Liao Y, Song Y, Dong H, Mangi MH, Ni J, et al: Nilotinib: A tyrosine kinase inhibitor mediates resistance to intracellular mycobacterium via regulating autophagy. Cells. 8:5062019. View Article : Google Scholar : PubMed/NCBI

182 

Pagan F, Hebron M, Valadez EH, Torres-Yaghi Y, Huang X, Mills RR, Wilmarth BM, Howard H, Dunn C, Carlson A, et al: Nilotinib effects in Parkinson's disease and dementia with lewy bodies. J Parkinsons Dis. 6:503–517. 2016. View Article : Google Scholar : PubMed/NCBI

183 

de Groot P, Nikolic T, Pellegrini S, Sordi V, Imangaliyev S, Rampanelli E, Hanssen N, Attaye I, Bakker G, Duinkerken G, et al: Faecal microbiota transplantation halts progression of human new-onset type 1 diabetes in a randomised controlled trial. Gut. 70:92–105. 2021. View Article : Google Scholar : PubMed/NCBI

184 

Gurung M, Li Z, You H, Rodrigues R, Jump DB, Morgun A and Shulzhenko N: Role of gut microbiota in type 2 diabetes pathophysiology. EBioMedicine. 51:1025902020. View Article : Google Scholar : PubMed/NCBI

185 

Horsager J, Andersen KB, Knudsen K, Skjærbæk C, Fedorova TD, Okkels N, Schaeffer E, Bonkat SK, Geday J, Otto M, et al: Brain-first versus body-first Parkinson's disease: A multimodal imaging case-control study. Brain. 143:3077–3088. 2020. View Article : Google Scholar : PubMed/NCBI

186 

Barbut D, Stolzenberg E and Zasloff M: Gastrointestinal immunity and alpha-synuclein. J Parkinsons Dis. 9:S313–S322. 2019. View Article : Google Scholar : PubMed/NCBI

187 

Perni M, Galvagnion C, Maltsev A, Meisl G, Müller MB, Challa PK, Kirkegaard JB, Flagmeier P, Cohen SI, Cascella R, et al: A natural product inhibits the initiation of α-synuclein aggregation and suppresses its toxicity. Proc Natl Acad Sci USA. 114:E1009–E1017. 2017. View Article : Google Scholar : PubMed/NCBI

188 

Das T and Eliezer D: Membrane interactions of intrinsically disordered proteins: The example of alpha-synuclein. Biochim Biophys Acta Proteins Proteom. 1867:879–889. 2019. View Article : Google Scholar : PubMed/NCBI

189 

Abbott SK, Li H, Muñoz SS, Knoch B, Batterham M, Murphy KE, Halliday GM and Garner B: Altered ceramide acyl chain length and ceramide synthase gene expression in Parkinson's disease. Mov Disord. 29:518–526. 2014. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

October-2021
Volume 24 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Lei Q, Wu T, Wu J, Hu X, Guan Y, Wang Y, Yan J and Shi G: Roles of α‑synuclein in gastrointestinal microbiome dysbiosis‑related Parkinson's disease progression (Review). Mol Med Rep 24: 734, 2021
APA
Lei, Q., Wu, T., Wu, J., Hu, X., Guan, Y., Wang, Y. ... Shi, G. (2021). Roles of α‑synuclein in gastrointestinal microbiome dysbiosis‑related Parkinson's disease progression (Review). Molecular Medicine Reports, 24, 734. https://doi.org/10.3892/mmr.2021.12374
MLA
Lei, Q., Wu, T., Wu, J., Hu, X., Guan, Y., Wang, Y., Yan, J., Shi, G."Roles of α‑synuclein in gastrointestinal microbiome dysbiosis‑related Parkinson's disease progression (Review)". Molecular Medicine Reports 24.4 (2021): 734.
Chicago
Lei, Q., Wu, T., Wu, J., Hu, X., Guan, Y., Wang, Y., Yan, J., Shi, G."Roles of α‑synuclein in gastrointestinal microbiome dysbiosis‑related Parkinson's disease progression (Review)". Molecular Medicine Reports 24, no. 4 (2021): 734. https://doi.org/10.3892/mmr.2021.12374