Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3370 2022-11-29 11:17:47 |
2 update references and layout Meta information modification 3370 2022-11-30 03:30:29 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Wang, Y.;  Zhang, Z.;  Li, B.;  He, B.;  Li, L.;  Nice, E.C.;  Zhang, W.;  Xu, J. Gut Microbiota in Neurodegenerative Diseases. Encyclopedia. Available online: https://encyclopedia.pub/entry/37118 (accessed on 27 July 2024).
Wang Y,  Zhang Z,  Li B,  He B,  Li L,  Nice EC, et al. Gut Microbiota in Neurodegenerative Diseases. Encyclopedia. Available at: https://encyclopedia.pub/entry/37118. Accessed July 27, 2024.
Wang, Yu, Zhe Zhang, Bowen Li, Bo He, Lei Li, Edouard C. Nice, Wei Zhang, Jia Xu. "Gut Microbiota in Neurodegenerative Diseases" Encyclopedia, https://encyclopedia.pub/entry/37118 (accessed July 27, 2024).
Wang, Y.,  Zhang, Z.,  Li, B.,  He, B.,  Li, L.,  Nice, E.C.,  Zhang, W., & Xu, J. (2022, November 29). Gut Microbiota in Neurodegenerative Diseases. In Encyclopedia. https://encyclopedia.pub/entry/37118
Wang, Yu, et al. "Gut Microbiota in Neurodegenerative Diseases." Encyclopedia. Web. 29 November, 2022.
Gut Microbiota in Neurodegenerative Diseases
Edit

An imbalance between oxidants and antioxidants in the body can lead to oxidative stress, which is one of the major causes of neurodegenerative diseases. The gut microbiota contains trillions of beneficial bacteria that play an important role in maintaining redox homeostasis. The microbiota–gut–brain axis has emerged as a new field that has revolutionized the study of the pathology, diagnosis, and treatment of neurodegenerative diseases. Indeed, a growing number of studies have found that communication between the brain and the gut microbiota can be accomplished through the endocrine, immune, and nervous systems. Importantly, dysregulation of the gut microbiota has been strongly associated with the development of oxidative stress-mediated neurodegenerative diseases. Therefore, a deeper understanding of the relationship between the gut microbiota and redox homeostasis will help explain the pathogenesis of neurodegenerative diseases from a new perspective and provide a theoretical basis for proposing new therapeutic strategies for neurodegenerative diseases.

neurodegenerative diseases oxidative stress redox homeostasis gut microbiota

1. Introduction

Progressive neuronal necrosis and degeneration are hallmarks of neurodegenerative diseases (NDs), including Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), and multiple sclerosis (MS), which are caused by neurotoxic etiological agents in the brain and surrounding organs. Because of the diverse and complex pathological symptoms, uncertain pathogenesis, restricted clinical examination, difficulty in making an early diagnosis, and lack of treatment options, NDs have imposed significant societal and economic burdens [1][2][3][4]. The current treatment of NDs mainly focuses on relieving symptoms, and no effective cure is currently available. Therefore, the search for novel and effective strategies to treat NDs remains a priority.
As a result of normal metabolic processes, reactive oxygen species (ROS) and reactive nitrogen species (RNS), being free radicals, are neutralized by endogenous antioxidants in cells and tissues to maintain redox homeostasis [5][6]. This redox homeostasis is disrupted under certain pathophysiological conditions, such as injury, inflammation, genetic mutations, and ischemia/reperfusion, causing oxidative stress, which is associated with a variety of progressive NDs [5][7]. The central nervous system (CNS) can form large amounts of free radicals due to its high oxygen demand and metabolism of neurotransmitters [8][9]. Notably, nerve cells are particularly vulnerable to damage because of the abundance of free radicals and relatively weak antioxidant defenses compared with other organs [10][11]. In addition, oxidative stress causes mitochondrial dysfunction, which is unable to meet the high energy requirements for normal biochemical and physiological functions of neuronal cells, thus leading to neuronal cell death [12]. Thus, maintenance of redox homeostasis is essential for neuronal survival and function.

2. Oxidative Stress and NDs

Oxidative stress, the result of an imbalance in the relative abundance of reactive ROS and antioxidants, can create a detrimental state that leads to cellular damage and dysfunction. Increased oxidative stress is able to damage cell membranes, alter protein structure and function, and cause DNA damage [13][14]. Therefore, maintenance of redox homeostasis is essential for cell biological function. As mentioned earlier, the CNS is particularly sensitive to oxidative stress due to its high metabolic rate, relative antioxidant scarcity, and unique structural features [15][16]. In addition, due to the presence of high levels of metal ions and polyunsaturated fatty acids, neuronal cells are more prone to oxidative stress, leading to cell damage and a series of NDs-related events through mitochondrial dysfunction, inflammation, and neuronal death [16][17][18]. In fact, an oxidative stress-induced imbalance in redox homeostasis is still a central component of the pathogenesis of several NDs, such as AD, PD, and MS. The common features among these NDs are ineffective antioxidant defense systems, imbalances of redox homeostasis, mitochondrial dysfunction, neuroinflammation, neuronal loss and degeneration (Figure 1).
Figure 1. The effect of oxidative stress in neurodegenerative diseases. An oxidant/antioxidant imbalance leads to oxidative stress, which causes DNA and protein damage, lipid peroxidation, and apoptosis. Dysfunctional mitochondria and activated neurons secrete inflammatory cytokines that cross the blood–brain barrier, leading to inflammation, α-synuclein, β-aggregation, and neuronal plaque accumulation in neurons, leading to neuron loss and degeneration.
Oxidative stress and disruption of cerebral redox homeostasis frequently occur in human NDs. For instance, in the pathology of AD, amyloid β (Aβ) and tau protein aggregates can interact with metal ions and maintain normal cellular signaling [19][20]. Furthermore, previous studies have shown that the high levels of zinc in the neocortical and hippocampal regions of AD patients suggest the vital role of zinc in the maintenance of redox homeostasis in the affected brain regions [21][22]. Notably, accumulated Aβ-induced oxidative stress can inhibit the activity of complex IV, leading to ATP depletion and mitochondrial dysfunction [23][24]. It has been demonstrated that the abnormal aggregation of α-synuclein (α-syn), mitochondrial dysfunction, and excessive oxidative stress are closely related to dopaminergic neuron death during PD progression [25][26][27]. As the main pathogenic factor of HD, soluble and aggregated mutant Htt (mHtt) protein with cytotoxicity induces apoptosis through oxidative stress, resulting in continuous degeneration of neurons [28][29][30]. Interestingly, in patients with NDs, oxidative stress biomarkers such as malondialdehyde and 8-hydroxyguanosine are elevated, and the gene superoxide dismutase 1 (SOD1), which plays an important role in oxidative stress defense mechanisms, is also frequently mutated [31][32].
Despite the advanced understanding of the mechanisms described above, a wide gap remains between this knowledge and the availability of effective therapies. Given that the imbalance of redox homeostasis is one of the key factors in the pathogenesis of NDs, numerous studies have been conducted on the treatment of NDs using various types of antioxidants (Table 1). Overall, most of the clinical trial results of NDs have shown favorable therapeutic effects, especially the alterations of pathological markers and improvements in neurological function, suggesting that antioxidants have great therapeutic potential for NDs. However, more efforts are required to explore novel therapeutic strategies and approaches to achieve broad therapeutic applicability and functional recovery of the nervous system.
Table 1. Antioxidants with therapeutic effects on neurodegenerative diseases.
Antioxidants Therapeutic
Target
Mechanism Reference
Luteolin PD Increased dopamine absorption [33]
Selenium AD Degradation of Aβ plaques [34][35]
Curcumin PD NRF2 activation [36]
α-Tocopherol AD Aβ plaque degradation [37]
Quercetin AD, PD Hydroxyl radical scavenging [38]
Ginsenosides AD Inhibition of Aβ aggregation [39]
PLGA NPs AD, PD, MS Protection against oxidative stress [40]
Macrophage-derived
exosomes
PD Protection against oxidative stress and inflammation [41]
Coenzyme Q10 AD Reduction of oxidative stress and senile plaques [42]
Ferulic acid AD Inhibition of neuronal oxidative stress [43]

3. Gut Microbiota, Oxidative Stress, and Neurodegeneration

The human gastrointestinal tract is the largest immune organ and harbors complex and dynamic microbiota [44][45]. Gut microbiota stability can be impacted by several variables, including genetics, lifestyle, nutrition, medications, illness, and age, which in turn have a significant impact on the regulation of metabolism, homeostasis, immunological response, and other processes [46][47]. Therefore, an imbalance in the representation of the gut microbiota may contribute to various diseases, from inflammatory bowel disease to obesity and diabetes, as well as several common NDs, such as AD, PD, and MS (Table 2). In addition, growing numbers of studies have demonstrated that the gut microbiota alters the oxidative/antioxidant balance in the CNS and causes neurodegeneration [48][49][50][51].
Table 2. Alterations in the gut microbiota composition in various neurodegenerative diseases.
Neurodegenerative
Disease
Experimental
Subject
Gut Microbiota Reference
AD Fecal samples from AD Firmicutes, Bifidobacterium ↓
Bacteroidetes ↑
[52]
  Symptomatic
Tg2576 mice
Firmicutes, Bacteroidetes, Lactobacillus ↑ [53]
  Fecal samples from
AD patients
Ruminococcacea ↑ Lachnospirace ↓ [54]
  Male patients with AD Bacteroidetes,
Blautia ↑
Firmicutes, Bifidobacterium ↓
[55]
  Amyloid-positive
patients
Escherichia,
Shigella ↑
Eubacterium
rectale ↓
[56]
PD Patients with PD Enterobacteriaceae, Serratia ↑
Blautia, Coprococcus, Lachnospiraceae ↓
[57]
  16S microbiome
datasets
Akkermansia,
Lactobacillus, Bifidobacterium ↑ Faecalibacterium, Lachnospiraceae ↓
[58]
  Patients with PD Butyricicoccus,
Clostridium ↑ Shigella, Lactobacillus ↓
[59]
MS Patients with MS Caproic acid, producers ↑
Butyric acid, producers ↓
[60]
  Patients with MS Patescibacteria ↑ Lachnospiraceae, Ruminococcaceae ↓ [61]

3.1. Gut–Brain Axis under Physiological Conditions

The brain and the gut microbiota are strictly intertwined and communicate in a variety of ways, including the production of bacterial metabolites, neurotransmitters, and cytokines [62] (Figure 2). Notably, the term “microbiota–gut–brain axis” refers to an interaction between the brain and the gut microbiota that involves four major routes of communication [63][64]. The first route of communication involves the vagus nerve, which connects the enteric nerve system and the brain stem. Recent research indicates that the gut microbiota influences host behaviors such as anxiety, feeding, and depression by activating vagal neurons and altering neurotransmitters such as γ-aminobutyric acid (GABA) and oxytocin in the brain [65][66]. The second important mode that directly or indirectly affects brain activity involves serotonin, which is mainly produced by gut enterochromaffin cells and modulates a variety of physiological processes. Interestingly, increased levels of serotonin and serotonin precursors alleviated depression in a mouse model of depression after treatment with the probiotic Bifidobacterium [67]. Thirdly, the gut microbiota plays an essential role in microglial activation and neuroinflammation. For instance, Luck and colleagues demonstrated that germ-free mice carry more immature microglia than conventional mice, and Bifidobacterium spp. can activate microglia through transcriptional activation [68]. In addition, alterations in microglial function were also observed in NDs and other behaviors, suggesting that the gut microbiota mediates effects on NDs through microglia [69]. Notably, the gut microbiota plays a vital role in energy harvest and neuroinflammation, and alterations in the gut–brain vagal pathway may promote obesity. It has been demonstrated that a diet-induced shift in the gut microbiome may disrupt vagal gut–brain communication resulting in microglia activation, increased gut inflammation, and body fat accumulation [70][71]. Finally, the gut microbiota communicates by transferring chemical signals directly to the brain. A previous study indicated that short-chain fatty acids (SCFAs) derived from the fermentation of the gut microbiota had been shown to modulate neuroplasticity in the CNS and improve depressive behavior in mice [50].
Figure 2. Microbiota–gut–brain axis. The brain and gut communicate through neural, metabolic, endocrine, and immunological pathways. The brain influences gut health through the vagus nerve, the hypothalamic–pituitary–adrenal (HPA) axis, and systemic circulation. Signals from the gut, including short-chain fatty acids (SCFAs), neurotransmitters, and amino acids, modulate brain function via neuronal cells, the immune system, and endocrine mechanisms.

3.2. Gut Microbiota-Mediated Oxidative Stress and Neurodegeneration

There are four main symbiotic bacteria that are parasitic in the human gut, namely Actinobacteria, Bacteroidetes, Proteobacteria, and Firmicutes. Among them, Firmicutes accounts for the largest proportion, including Streptococcus, Lactobacillus, and Mycoplasma [69]. Recent studies have found that, in the presence of the microbiota, the intestinal epithelium cells produce physiological levels of oxidative stress that affect the composition and function of the gut microbiota. Such alterations in gut microbiota increase the alterations of biomacromolecules reaching the systemic circulation and CNS by directly affecting the permeability of the intestine [72]. Indeed, the gut microbiota can alter cellular oxidative stress status by regulating mitochondrial activity [73]. In addition, gut Lactobacilli, Bifidobacterium, and Streptococcus can produce nitric oxide (NO) in various ways in the gut [74][75]. It is now generally accepted that NO in nanomolar concentrations is neuroprotective, whereas higher concentrations of NO may result in oxidative stress, which is closely related to axonal degeneration, neuroinflammation, and NDs [69]. In addition, certain pathogenic bacteria such as Salmonella and Escherichia coli are able to produce hydrogen sulfide (H2S) in the gut by degrading sulfur-containing amino acids. Furthermore, increased levels of H2S alter various host metabolic activities, such as increased lactate, decreased oxygen consumption, decreased ATP production, and elevated levels of proinflammatory compounds, which have been linked to neuroinflammation [76][77][78]. The role of gut microbiota in neurodegeneration is shown in Figure 3.
Figure 3. The role of the gut microbiota in neurodegeneration is depicted schematically. Bad mood, increasing age, drugs, dietary changes, and circadian rhythms can disrupt gut microbiota homeostasis. When gut dysbiosis occurs, beneficial bacteria in the gut are transformed into pathogenic bacteria, producing a large number of harmful metabolites and proinflammatory molecules, resulting in increased blood–brain barrier permeability and peripheral inflammatory responses, thereby aggravating oxidative stress in the brain. At the same time, dysbiosis can induce bad mood. Increased levels of ROS in neuronal mitochondria, endoplasmic reticulum, and peroxisomes, increased protein and lipid oxidation, and accumulation of neurotoxic proteins lead to neurodegeneration.

3.2.1. Alzheimer’s Disease

AD is the most common ND worldwide with an insidious onset and progressive development [4][79]. It is characterized by progressive impairment of cognition and episodic memory, culminating in the development of dementia [80]. Specific histopathological hallmarks in the brain associated with AD include Aβ plaques, neurofibrillary tangles (NFTs), hyperphosphorylation of tau proteins (tau tangles), and neuronal loss [81][82]. Oxidative stress has been suggested to play an essential role in AD etiology prior to plaque formation, leading to mitochondrial dysfunction in neurons and synapses, as well as Aβ protein production [83][84]. Previous studies have shown that oxidative stress plays a pivotal role in the development of AD. For instance, it has been demonstrated that aggregated Aβ protein stimulates microglia to produce ROS through positive feedback on Aβ plaque deposition [79]. In addition, tau protein aggregation in neurons leads to reduced NADH-ubiquitin reductase activity, leading to oxidative stress and mitochondrial dysfunction [85]. Interestingly, ROS can affect the activity of stress kinases, such as the phosphorylation-c-Jun N-terminal kinase 1 (p-JNK) pathway, which is associated with neuronal cell death due to tau hyperphosphorylation and accumulation of Aβ [86]. There is ample evidence that the oxidation of nucleic acid species in the AD brain is dominated by the mitochondrial genome, and lipid peroxidation results in the production of certain cytotoxic agents, such as 4-hydroxyalkenals [87][88][89].
Recent studies have shown that the gut microbiota plays a significant role in the pathogenesis of AD [90]. Dysregulation of the gut microbiota leads to oxidative stress, inflammation, disruption of the blood–brain barrier, activation of the immune system, neurofibrillary tangles, and Aβ plaques followed by neurodegeneration [91][92]. There are numerous bacteria in the human gut that play a vital role in the etiology of AD, including Staphylococcus aureus, Escherichia coli, Salmonella, Mycobacterium, Klebsiella pneumoniae, and Streptococcus, which promote the production and aggregation of the Aβ protein in the enteric nervous system [93][94]. Interestingly, in the APPSWE/PS1ΔE9 transgenic mouse model of AD chronically treated with broad-spectrum combination antibiotics, the gut microbiome of transgenic mice shifted toward proinflammatory bacteria, with a decrease in amyloid plaque deposition and neuroinflammation [54][95]. Additionally, microbial amyloid protein, produced by coccus-shaped bacteria, is able to activate the innate immune system and triggers responses by Toll-like receptors (TLRs) and cluster of differentiation 14 (CD14), resulting in inadequate recognition of misfolded Aβ and decreased Aβ clearance, followed by the production of cytokines leading to intestinal disturbances [96]. Notably, age-related reductions in gut microbial diversity are also implicated in AD. It has been demonstrated that with growing age, there is an increase in Proteobacteria and a decrease in Bifidobacterium spp., which results in interference in lipid metabolism and a failure to maintain hippocampal plasticity as well as memory functions [97][98].
Another possible connecting link between the gut microbiota and microbiota-mediated cerebral amyloid accumulation involves a cross-seeding mechanism of microbial amyloid (i.e., promoting misfolded aggregation of amyloid from one protein to another) in a manner similar to the reproduction of prions [96][99][100]. Notably, distinct amyloid conformations interact with cellular targets to produce various toxicities, which may explain the different AD phenotypes [101]. Given the multiple roles of gut microbiota dysbiosis in the pathogenesis of AD, modulation of AD through dietary and gut microbiota interventions may be potential therapeutic strategies, which will be discussed in detail later.

3.2.2. Parkinson’s Disease

PD, the second most common ND after AD, is a long-term neurological disorder that causes both motor and non-motor symptoms [102][103][104]. Motor symptoms include resting tremors, akinesia, muscular rigidity, postural instability, and gait abnormalities [105][106][107]. Non-motor symptoms include anxiety, depression, autonomic dysfunction, cognitive decline, and sleep disturbances [108][109]. The hallmarks of PD are loss of dopaminergic (DA) neurons and abnormal accumulation of α-syn within the cytoplasm of nerve cells called Levy bodies [110][111][112][113]. Notably, oxidative stress, mitochondrial dysfunction, dopamine metabolism, abnormal protein aggregation, and the gut microbiota are associated with the pathogenesis of PD [114][115][116]. As one of the main pathogenic factors of PD, oxidative stress has been linked to α-syn protein aggregation and degeneration in DA neurons [86][110][117]. For instance, analysis of the postmortem brain tissue in PD showed that oxidative stress degenerates DA neurons, reduces the levels of glutathione (GSH), increases the levels of oxidative stress markers, and stimulates lipid, DNA, and RNA oxidation [118][119]. Additionally, Tong and colleagues have demonstrated that oxidative stress in DA neurons can activate the p38 mitogen-activated protein kinase pathway, ultimately leading to neuronal apoptosis [120]. Interestingly, in the 6-hydroxydopamine-induced PD model in mice, Antrodia camphorata polysaccharide reduced ROS by increasing the expression and activity of antioxidant enzymes, ultimately attenuating the damage of DA neurons in the substantia nigra and improving motor performance [121].
Notably, PD patients often present with gastrointestinal dysfunction, which suggests that the imbalance of the gut microbiota is one of the causes of triggering or aggravating PD [122][123]. Indeed, gut inflammation, early accumulation of α-syn, increased intestinal permeability, and constipation problems are common in PD patients, again demonstrating the critical role of the gut microbiota in PD [69][124]. It has been demonstrated that the disruption of gut microbiota leads to oxidative stress through overstimulation of the immune system, which in turn activates intestinal neurons and intestinal glia cells, leading to increased misfolding and accumulation of α-syn in the CNS [125][126]. In addition, coming to the role of the gut microbiota, toxins and microbial products produced by certain pathogenic bacteria are able to cause mitochondrial dysfunction in intestinal cells and the CNS, which is directly associated with PD pathogenesis [127]. In line with this, it is proposed that the pathogenic bacterium E. coli can produce an amyloid protein called curli, which promotes the accumulation of α-syn protein in the brain and causes motor defects in mice [128]. Conversely, when treated with gut-restricted amyloid inhibitors, these mice showed significant improvements in constipation and motor function, suggesting the role of the gut microbiota in the etiology of PD symptoms [129]. Interestingly, when the gut microbiota from PD patients was transplanted into a germ-free α-syn overexpressed mouse model, a similar pattern of physical injury to PD patients was observed, suggesting the vital role of the gut microbiota in PD. As discussed above, decreased production of hydrogen (H2) by the gut microbiota has been proposed as one of the essential factors in PD [130]. According to a recent study, 50% H2 saturated water was successful in preventing nigrostriatal degeneration in PD rats and reducing the oxidative stress markers in the 1-methyl-4-phenyl-1,2,3,4-tetrahydropyridine (MPTP) mouse model [131]. Taken together, these observations suggest a critical role for gut microbiota in PD, and intervention through gut microbiota is expected to provide promising strategies for the prevention and treatment of PD.

3.2.3. Multiple Sclerosis

MS is an immune-mediated chronic inflammatory and central nervous system demyelinating disease with a complex and unclear pathogenesis [132]. It has been established that the pathogenesis of MS involves both genetic and environmental factors. The most extensively accepted hypothesis is that autoreactive B and T cells cause axonal and myelin damage, as well as neurodegeneration [133][134]. The major neuropathological hallmarks of MS pathology are inflammation and degeneration of both white matter and gray matter [135]. However, the development of MS may be influenced by a combination of internal and external factors, ultimately leading to immune dysregulation.
Growing evidence suggests that the imbalance of redox homeostasis plays a vital role in the pathogenesis of MS. For instance, it has been demonstrated that the excessive generation of ROS, mitochondrial dysfunction, and impairment of antioxidant defense systems play important roles in the pathogenesis of MS [136]. Notably, ROS has been shown to be a mediator of axonal injury and demyelination in both MS patients and animal models of MS. In addition, oxidative stress mediates mitochondrial dysfunction in MS patients and leads to CNS energy failure in MS-susceptible individuals [137][138]. Furthermore, recent studies have demonstrated that the gut microbiota has a significant impact on MS and can be influenced by external factors [139]. For instance, Cosorich et al. have demonstrated that T helper 17 (TH17) cells, key players in MS, originate in the gut and that increased TH17 cell frequency is associated with specific alterations of the gut microbiota in MS patients [140]. Interestingly, transplantation of the MS microbiota in a mouse model resulted in an increased incidence of autoimmune encephalomyelitis, leading to an exacerbation of MS symptoms [141][142]. Additionally, diets have been shown to affect the balance of the gut microbiota and indirectly influence the development of MS [143]. Moreover, dietary studies in MS patients suggest that dietary interventions supplemented with vitamin D in a low-calorie diet have a positive effect on alleviating chronic inflammatory symptoms in MS [144]. Recently, intermittent fasting was introduced into the treatment of MS due to its availability of abundant gut microbiota as well as the secretion of glutathione and leptin [145]. All these studies have shown that modification of the gut microbiota can be considered a promising therapeutic strategy for MS.

References

  1. Chen-Plotkin, A.S. Unbiased approaches to biomarker discovery in neurodegenerative diseases. Neuron 2014, 84, 594–607.
  2. Kwon, M.J.; Kim, S.; Han, M.H.; Lee, S.B. Epigenetic Changes in Neurodegenerative Diseases. Mol. Cells 2016, 39, 783–789.
  3. Muddapu, V.R.; Dharshini, S.A.P.; Chakravarthy, V.S.; Gromiha, M.M. Neurodegenerative Diseases—Is Metabolic Deficiency the Root Cause? Front. Neurosci. 2020, 14, 213.
  4. Wang, Y.; Chen, R.; Yang, Z.; Wen, Q.; Cao, X.; Zhao, N.; Yan, J. Protective Effects of Polysaccharides in Neurodegenerative Diseases. Front. Aging Neurosci. 2022, 14, 917629.
  5. Ashok, A.; Andrabi, S.S.; Mansoor, S.; Kuang, Y.; Kwon, B.K.; Labhasetwar, V. Antioxidant Therapy in Oxidative Stress-Induced Neurodegenerative Diseases: Role of Nanoparticle-Based Drug Delivery Systems in Clinical Translation. Antioxidants 2022, 11, 408.
  6. Kurutas, E.B. The importance of antioxidants which play the role in cellular response against oxidative/nitrosative stress: Current state. Nutr. J. 2016, 15, 71.
  7. Cunha-Oliveira, T.; Montezinho, L.; Mendes, C.; Firuzi, O.; Saso, L.; Oliveira, P.J.; Silva, F.S.G. Oxidative Stress in Amyotrophic Lateral Sclerosis: Pathophysiology and Opportunities for Pharmacological Intervention. Oxid. Med. Cell. Longev. 2020, 2020, 5021694.
  8. Magistretti, P.J.; Allaman, I. A cellular perspective on brain energy metabolism and functional imaging. Neuron 2015, 86, 883–901.
  9. Siraki, A.G.; O’Brien, P.J. Prooxidant activity of free radicals derived from phenol-containing neurotransmitters. Toxicology 2002, 177, 81–90.
  10. Cobley, J.N.; Fiorello, M.L.; Bailey, D.M. 13 reasons why the brain is susceptible to oxidative stress. Redox Biol. 2018, 15, 490–503.
  11. Salim, S. Oxidative Stress and the Central Nervous System. J. Pharmacol. Exp. Ther. 2017, 360, 201–205.
  12. Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D.; et al. Oxidative stress, aging, and diseases. Clin. Interv. Aging 2018, 13, 757–772.
  13. Gandhi, S.; Abramov, A.Y. Mechanism of oxidative stress in neurodegeneration. Oxid. Med. Cell. Longev. 2012, 2012, 428010.
  14. Li, B.; Jiang, J.; Assaraf, Y.G.; Xiao, H.; Chen, Z.S.; Huang, C. Surmounting cancer drug resistance: New insights from the perspective of N(6)-methyladenosine RNA modification. Drug Resist. Updates 2020, 53, 100720.
  15. Lin, M.T.; Beal, M.F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443, 787–795.
  16. Wang, X.; Michaelis, E.K. Selective neuronal vulnerability to oxidative stress in the brain. Front. Aging Neurosci. 2010, 2, 12.
  17. Selivanov, V.A.; Votyakova, T.V.; Pivtoraiko, V.N.; Zeak, J.; Sukhomlin, T.; Trucco, M.; Roca, J.; Cascante, M. Reactive oxygen species production by forward and reverse electron fluxes in the mitochondrial respiratory chain. PLoS Comput. Biol. 2011, 7, e1001115.
  18. DeTure, M.A.; Dickson, D.W. The neuropathological diagnosis of Alzheimer’s disease. Mol. Neurodegener. 2019, 14, 32.
  19. Mattson, M.P. Pathways towards and away from Alzheimer’s disease. Nature 2004, 430, 631–639.
  20. Selkoe, D.J. Alzheimer’s disease results from the cerebral accumulation and cytotoxicity of amyloid beta-protein. J. Alzheimer’s Dis. JAD 2001, 3, 75–80.
  21. Cuajungco, M.P.; Fagét, K.Y. Zinc takes the center stage: Its paradoxical role in Alzheimer’s disease. Brain Res. Brain Res. Rev. 2003, 41, 44–56.
  22. Huang, X.; Moir, R.D.; Tanzi, R.E.; Bush, A.I.; Rogers, J.T. Redox-active metals, oxidative stress, and Alzheimer’s disease pathology. Ann. N. Y. Acad. Sci. 2004, 1012, 153–163.
  23. Praticò, D. Oxidative stress hypothesis in Alzheimer’s disease: A reappraisal. Trends Pharmacol. Sci. 2008, 29, 609–615.
  24. Zhao, Y.; Zhao, B. Oxidative stress and the pathogenesis of Alzheimer’s disease. Oxid. Med. Cell. Longev. 2013, 2013, 316523.
  25. Luk, K.C.; Kehm, V.; Carroll, J.; Zhang, B.; O’Brien, P.; Trojanowski, J.Q.; Lee, V.M. Pathological α-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science 2012, 338, 949–953.
  26. Mandemakers, W.; Morais, V.A.; De Strooper, B. A cell biological perspective on mitochondrial dysfunction in Parkinson disease and other neurodegenerative diseases. J. Cell Sci. 2007, 120, 1707–1716.
  27. Ren, X.; Butterfield, D.A. Fidelity of the PINK1 knockout rat to oxidative stress and other characteristics of Parkinson disease. Free Radic. Biol. Med. 2021, 163, 88–101.
  28. Reijonen, S.; Putkonen, N.; Nørremølle, A.; Lindholm, D.; Korhonen, L. Inhibition of endoplasmic reticulum stress counteracts neuronal cell death and protein aggregation caused by N-terminal mutant huntingtin proteins. Exp. Cell Res. 2008, 314, 950–960.
  29. Ayala-Peña, S. Role of oxidative DNA damage in mitochondrial dysfunction and Huntington’s disease pathogenesis. Free Radic. Biol. Med. 2013, 62, 102–110.
  30. Pelegrí, C.; Duran-Vilaregut, J.; del Valle, J.; Crespo-Biel, N.; Ferrer, I.; Pallàs, M.; Camins, A.; Vilaplana, J. Cell cycle activation in striatal neurons from Huntington’s disease patients and rats treated with 3-nitropropionic acid. Int. J. Dev. Neurosci. Off. J. Int. Soc. Dev. Neurosci. 2008, 26, 665–671.
  31. Wang, Z.; Bai, Z.; Qin, X.; Cheng, Y. Aberrations in Oxidative Stress Markers in Amyotrophic Lateral Sclerosis: A Systematic Review and Meta-Analysis. Oxid. Med. Cell. Longev. 2019, 2019, 1712323.
  32. Pansarasa, O.; Bordoni, M.; Diamanti, L.; Sproviero, D.; Gagliardi, S.; Cereda, C. SOD1 in Amyotrophic Lateral Sclerosis: “Ambivalent” Behavior Connected to the Disease. Int. J. Mol. Sci. 2018, 19, 1345.
  33. Chen, H.Q.; Jin, Z.Y.; Wang, X.J.; Xu, X.M.; Deng, L.; Zhao, J.W. Luteolin protects dopaminergic neurons from inflammation-induced injury through inhibition of microglial activation. Neurosci. Lett. 2008, 448, 175–179.
  34. Jiang, T.; Sun, Q.; Chen, S. Oxidative stress: A major pathogenesis and potential therapeutic target of antioxidative agents in Parkinson’s disease and Alzheimer’s disease. Prog. Neurobiol. 2016, 147, 1–19.
  35. Conte, V.; Uryu, K.; Fujimoto, S.; Yao, Y.; Rokach, J.; Longhi, L.; Trojanowski, J.Q.; Lee, V.M.; McIntosh, T.K.; Praticò, D. Vitamin E reduces amyloidosis and improves cognitive function in Tg2576 mice following repetitive concussive brain injury. J. Neurochem. 2004, 90, 758–764.
  36. Cuadrado, A.; Moreno-Murciano, P.; Pedraza-Chaverri, J. The transcription factor Nrf2 as a new therapeutic target in Parkinson’s disease. Expert Opin. Ther. Targets 2009, 13, 319–329.
  37. Nakashima, H.; Ishihara, T.; Yokota, O.; Terada, S.; Trojanowski, J.Q.; Lee, V.M.; Kuroda, S. Effects of alpha-tocopherol on an animal model of tauopathies. Free Radic. Biol. Med. 2004, 37, 176–186.
  38. Karuppagounder, S.S.; Madathil, S.K.; Pandey, M.; Haobam, R.; Rajamma, U.; Mohanakumar, K.P. Quercetin up-regulates mitochondrial complex-I activity to protect against programmed cell death in rotenone model of Parkinson’s disease in rats. Neuroscience 2013, 236, 136–148.
  39. Aalinkeel, R.; Kutscher, H.L.; Singh, A.; Cwiklinski, K.; Khechen, N.; Schwartz, S.A.; Prasad, P.N.; Mahajan, S.D. Neuroprotective effects of a biodegradable poly(lactic-co-glycolic acid)-ginsenoside Rg3 nanoformulation: A potential nanotherapy for Alzheimer’s disease? J. Drug Target. 2018, 26, 182–193.
  40. Singhal, A.; Morris, V.B.; Labhasetwar, V.; Ghorpade, A. Nanoparticle-mediated catalase delivery protects human neurons from oxidative stress. Cell Death Dis 2013, 4, e903.
  41. Haney, M.J.; Klyachko, N.L.; Zhao, Y.; Gupta, R.; Plotnikova, E.G.; He, Z.; Patel, T.; Piroyan, A.; Sokolsky, M.; Kabanov, A.V.; et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J. Control. Release Off. J. Control. Release Soc. 2015, 207, 18–30.
  42. Wang, Z.H.; Wang, Z.Y.; Sun, C.S.; Wang, C.Y.; Jiang, T.Y.; Wang, S.L. Trimethylated chitosan-conjugated PLGA nanoparticles for the delivery of drugs to the brain. Biomaterials 2010, 31, 908–915.
  43. Picone, P.; Bondi, M.L.; Montana, G.; Bruno, A.; Pitarresi, G.; Giammona, G.; Di Carlo, M. Ferulic acid inhibits oxidative stress and cell death induced by Ab oligomers: Improved delivery by solid lipid nanoparticles. Free Radic. Res. 2009, 43, 1133–1145.
  44. Kraehenbuhl, J.P.; Neutra, M.R. Molecular and cellular basis of immune protection of mucosal surfaces. Physiol. Rev. 1992, 72, 853–879.
  45. Cheng, W.Y.; Ho, Y.S.; Chang, R.C. Linking circadian rhythms to microbiome-gut-brain axis in aging-associated neurodegenerative diseases. Ageing Res. Rev. 2022, 78, 101620.
  46. Zhuang, L.; Chen, H.; Zhang, S.; Zhuang, J.; Li, Q.; Feng, Z. Intestinal Microbiota in Early Life and Its Implications on Childhood Health. Genom. Proteom. Bioinform. 2019, 17, 13–25.
  47. Wells, J.M.; Brummer, R.J.; Derrien, M.; MacDonald, T.T.; Troost, F.; Cani, P.D.; Theodorou, V.; Dekker, J.; Méheust, A.; de Vos, W.M.; et al. Homeostasis of the gut barrier and potential biomarkers. Am. J. Physiol. Gastrointest. Liver Physiol. 2017, 312, G171–G193.
  48. Dumitrescu, L.; Popescu-Olaru, I.; Cozma, L.; Tulbă, D.; Hinescu, M.E.; Ceafalan, L.C.; Gherghiceanu, M.; Popescu, B.O. Oxidative Stress and the Microbiota-Gut-Brain Axis. Oxid. Med. Cell. Longev. 2018, 2018, 2406594.
  49. Wang, H.; Wang, G.; Banerjee, N.; Liang, Y.; Du, X.; Boor, P.J.; Hoffman, K.L.; Khan, M.F. Aberrant Gut Microbiome Contributes to Intestinal Oxidative Stress, Barrier Dysfunction, Inflammation and Systemic Autoimmune Responses in MRL/lpr Mice. Front. Immunol. 2021, 12, 651191.
  50. Dalile, B.; Van Oudenhove, L.; Vervliet, B.; Verbeke, K. The role of short-chain fatty acids in microbiota-gut-brain communication. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 461–478.
  51. Yarandi, S.S.; Peterson, D.A.; Treisman, G.J.; Moran, T.H.; Pasricha, P.J. Modulatory Effects of Gut Microbiota on the Central Nervous System: How Gut Could Play a Role in Neuropsychiatric Health and Diseases. J. Neurogastroenterol. Motil. 2016, 22, 201–212.
  52. Zhang, L.; Wang, Y.; Xiayu, X.; Shi, C.; Chen, W.; Song, N.; Fu, X.; Zhou, R.; Xu, Y.F.; Huang, L.; et al. Altered Gut Microbiota in a Mouse Model of Alzheimer’s Disease. J. Alzheimer’s Dis. JAD 2017, 60, 1241–1257.
  53. Cilia, R.; Piatti, M.; Cereda, E.; Bolliri, C.; Caronni, S.; Ferri, V.; Cassani, E.; Bonvegna, S.; Ferrarese, C.; Zecchinelli, A.L.; et al. Does Gut Microbiota Influence the Course of Parkinson’s Disease? A 3-Year Prospective Exploratory Study in de novo Patients. J. Park. Dis. 2021, 11, 159–170.
  54. Bäuerl, C.; Collado, M.C.; Diaz Cuevas, A.; Viña, J.; Pérez Martínez, G. Shifts in gut microbiota composition in an APP/PSS1 transgenic mouse model of Alzheimer’s disease during lifespan. Lett. Appl. Microbiol. 2018, 66, 464–471.
  55. Vogt, N.M.; Kerby, R.L.; Dill-McFarland, K.A.; Harding, S.J.; Merluzzi, A.P.; Johnson, S.C.; Carlsson, C.M.; Asthana, S.; Zetterberg, H.; Blennow, K.; et al. Gut microbiome alterations in Alzheimer’s disease. Sci. Rep. 2017, 7, 13537.
  56. Cattaneo, A.; Cattane, N.; Galluzzi, S.; Provasi, S.; Lopizzo, N.; Festari, C.; Ferrari, C.; Guerra, U.P.; Paghera, B.; Muscio, C.; et al. Association of brain amyloidosis with pro-inflammatory gut bacterial taxa and peripheral inflammation markers in cognitively impaired elderly. Neurobiol. Aging 2017, 49, 60–68.
  57. Saresella, M.; Marventano, I.; Barone, M.; La Rosa, F.; Piancone, F.; Mendozzi, L.; d’Arma, A.; Rossi, V.; Pugnetti, L.; Roda, G.; et al. Alterations in Circulating Fatty Acid Are Associated With Gut Microbiota Dysbiosis and Inflammation in Multiple Sclerosis. Front. Immunol. 2020, 11, 1390.
  58. Romano, S.; Savva, G.M.; Bedarf, J.R.; Charles, I.G.; Hildebrand, F.; Narbad, A. Meta-analysis of the Parkinson’s disease gut microbiome suggests alterations linked to intestinal inflammation. NPJ Park. Dis. 2021, 7, 27.
  59. Qian, Y.; Yang, X.; Xu, S.; Wu, C.; Song, Y.; Qin, N.; Chen, S.D.; Xiao, Q. Alteration of the fecal microbiota in Chinese patients with Parkinson’s disease. Brain Behav. Immun. 2018, 70, 194–202.
  60. Mazzini, L.; Mogna, L.; De Marchi, F.; Amoruso, A.; Pane, M.; Aloisio, I.; Cionci, N.B.; Gaggìa, F.; Lucenti, A.; Bersano, E.; et al. Potential Role of Gut Microbiota in ALS Pathogenesis and Possible Novel Therapeutic Strategies. J. Clin. Gastroenterol. 2018, 52, S68–S70.
  61. Tremlett, H.; Zhu, F.; Arnold, D.; Bar-Or, A.; Bernstein, C.N.; Bonner, C.; Forbes, J.D.; Graham, M.; Hart, J.; Knox, N.C.; et al. The gut microbiota in pediatric multiple sclerosis and demyelinating syndromes. Ann. Clin. Transl. Neurol. 2021, 8, 2252–2269.
  62. Quigley, E.M.M. Microbiota-Brain-Gut Axis and Neurodegenerative Diseases. Curr. Neurol. Neurosci. Rep. 2017, 17, 94.
  63. Powell, N.; Walker, M.M.; Talley, N.J. The mucosal immune system: Master regulator of bidirectional gut-brain communications. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 143–159.
  64. Natale, G.; Ryskalin, L.; Morucci, G.; Lazzeri, G.; Frati, A.; Fornai, F. The Baseline Structure of the Enteric Nervous System and Its Role in Parkinson’s Disease. Life 2021, 11, 732.
  65. Bravo, J.A.; Forsythe, P.; Chew, M.V.; Escaravage, E.; Savignac, H.M.; Dinan, T.G.; Bienenstock, J.; Cryan, J.F. Ingestion of Lactobacillus strain regulates emotional behavior and central GABA receptor expression in a mouse via the vagus nerve. Proc. Natl. Acad. Sci. USA 2011, 108, 16050–16055.
  66. Sgritta, M.; Dooling, S.W.; Buffington, S.A.; Momin, E.N.; Francis, M.B.; Britton, R.A.; Costa-Mattioli, M. Mechanisms Underlying Microbial-Mediated Changes in Social Behavior in Mouse Models of Autism Spectrum Disorder. Neuron 2019, 101, 246–259.e246.
  67. Tian, P.; Wang, G.; Zhao, J.; Zhang, H.; Chen, W. Bifidobacterium with the role of 5-hydroxytryptophan synthesis regulation alleviates the symptom of depression and related microbiota dysbiosis. J. Nutr. Biochem. 2019, 66, 43–51.
  68. Luck, B.; Engevik, M.A.; Ganesh, B.P.; Lackey, E.P.; Lin, T.; Balderas, M.; Major, A.; Runge, J.; Luna, R.A.; Sillitoe, R.V.; et al. Bifidobacteria shape host neural circuits during postnatal development by promoting synapse formation and microglial function. Sci. Rep. 2020, 10, 7737.
  69. Shandilya, S.; Kumar, S.; Kumar Jha, N.; Kumar Kesari, K.; Ruokolainen, J. Interplay of gut microbiota and oxidative stress: Perspective on neurodegeneration and neuroprotection. J. Adv. Res. 2022, 38, 223–244.
  70. Vaughn, A.C.; Cooper, E.M.; DiLorenzo, P.M.; O’Loughlin, L.J.; Konkel, M.E.; Peters, J.H.; Hajnal, A.; Sen, T.; Lee, S.H.; de La Serre, C.B.; et al. Energy-dense diet triggers changes in gut microbiota, reorganization of gut-brain vagal communication and increases body fat accumulation. Acta Neurobiol. Exp. 2017, 77, 18–30.
  71. Sen, T.; Cawthon, C.R.; Ihde, B.T.; Hajnal, A.; DiLorenzo, P.M.; de La Serre, C.B.; Czaja, K. Diet-driven microbiota dysbiosis is associated with vagal remodeling and obesity. Physiol. Behav. 2017, 173, 305–317.
  72. Reese, A.T.; Cho, E.H.; Klitzman, B.; Nichols, S.P.; Wisniewski, N.A.; Villa, M.M.; Durand, H.K.; Jiang, S.; Midani, F.S.; Nimmagadda, S.N.; et al. Antibiotic-induced changes in the microbiota disrupt redox dynamics in the gut. eLife 2018, 7, e35987.
  73. Zorov, D.B.; Plotnikov, E.Y.; Silachev, D.N.; Zorova, L.D.; Pevzner, I.B.; Zorov, S.D.; Babenko, V.A.; Jankauskas, S.S.; Popkov, V.A.; Savina, P.S. Microbiota and mitobiota. Putting an equal sign between mitochondria and bacteria. Biochemistry. Biokhimiia 2014, 79, 1017–1031.
  74. Migeotte, I.; Communi, D.; Parmentier, M. Formyl peptide receptors: A promiscuous subfamily of G protein-coupled receptors controlling immune responses. Cytokine Growth Factor Rev. 2006, 17, 501–519.
  75. Tiso, M.; Schechter, A.N. Nitrate reduction to nitrite, nitric oxide and ammonia by gut bacteria under physiological conditions. PLoS ONE 2015, 10, e0119712.
  76. Leschelle, X.; Goubern, M.; Andriamihaja, M.; Blottière, H.M.; Couplan, E.; Gonzalez-Barroso, M.D.; Petit, C.; Pagniez, A.; Chaumontet, C.; Mignotte, B.; et al. Adaptative metabolic response of human colonic epithelial cells to the adverse effects of the luminal compound sulfide. Biochim. Et Biophys. Acta 2005, 1725, 201–212.
  77. Beaumont, M.; Andriamihaja, M.; Lan, A.; Khodorova, N.; Audebert, M.; Blouin, J.M.; Grauso, M.; Lancha, L.; Benetti, P.H.; Benamouzig, R.; et al. Detrimental effects for colonocytes of an increased exposure to luminal hydrogen sulfide: The adaptive response. Free Radic. Biol. Med. 2016, 93, 155–164.
  78. Donertas Ayaz, B.; Zubcevic, J. Gut microbiota and neuroinflammation in pathogenesis of hypertension: A potential role for hydrogen sulfide. Pharm. Res 2020, 153, 104677.
  79. Cheignon, C.; Tomas, M.; Bonnefont-Rousselot, D.; Faller, P.; Hureau, C.; Collin, F. Oxidative stress and the amyloid beta peptide in Alzheimer’s disease. Redox Biol. 2018, 14, 450–464.
  80. Agostinho, P.; Cunha, R.A.; Oliveira, C. Neuroinflammation, oxidative stress and the pathogenesis of Alzheimer’s disease. Curr. Pharm. Des. 2010, 16, 2766–2778.
  81. Outeiro, T.F.; Koss, D.J.; Erskine, D.; Walker, L.; Kurzawa-Akanbi, M.; Burn, D.; Donaghy, P.; Morris, C.; Taylor, J.P.; Thomas, A.; et al. Dementia with Lewy bodies: An update and outlook. Mol. Neurodegener. 2019, 14, 5.
  82. Sato, C.; Barthélemy, N.R.; Mawuenyega, K.G.; Patterson, B.W.; Gordon, B.A.; Jockel-Balsarotti, J.; Sullivan, M.; Crisp, M.J.; Kasten, T.; Kirmess, K.M.; et al. Tau Kinetics in Neurons and the Human Central Nervous System. Neuron 2018, 97, 1284–1298.e1287.
  83. Markesbery, W.R. Oxidative stress hypothesis in Alzheimer’s disease. Free Radic. Biol. Med. 1997, 23, 134–147.
  84. Sharma, C.; Kim, S.R. Linking Oxidative Stress and Proteinopathy in Alzheimer’s Disease. Antioxidants 2021, 10, 1231.
  85. Abramov, A.Y.; Potapova, E.V.; Dremin, V.V.; Dunaev, A.V. Interaction of Oxidative Stress and Misfolded Proteins in the Mechanism of Neurodegeneration. Life 2020, 10, 101.
  86. Patten, D.A.; Germain, M.; Kelly, M.A.; Slack, R.S. Reactive oxygen species: Stuck in the middle of neurodegeneration. J. Alzheimer’s Dis. JAD 2010, 20 (Suppl. S2), S357–S367.
  87. Hong, W.K.; Han, E.H.; Kim, D.G.; Ahn, J.Y.; Park, J.S.; Han, B.G. Amyloid-beta-peptide reduces the expression level of mitochondrial cytochrome oxidase subunits. Neurochem. Res. 2007, 32, 1483–1488.
  88. Wang, Y.; Wu, F.; Pan, H.; Zheng, W.; Feng, C.; Wang, Y.; Deng, Z.; Wang, L.; Luo, J.; Chen, S. Lost region in amyloid precursor protein (APP) through TALEN-mediated genome editing alters mitochondrial morphology. Sci. Rep. 2016, 6, 22244.
  89. Peña-Bautista, C.; Baquero, M.; Vento, M.; Cháfer-Pericás, C. Free radicals in Alzheimer’s disease: Lipid peroxidation biomarkers. Clin. Chim. Acta Int. J. Clin. Chem. 2019, 491, 85–90.
  90. Hadidi Zavareh, A.H.; Haji Khani, R.; Pakpour, B.; Soheili, M.; Salami, M. Probiotic treatment differentially affects the behavioral and electrophysiological aspects in ethanol exposed animals. Iran. J. Basic Med. Sci. 2020, 23, 776–780.
  91. Soheili, M.; Alinaghipour, A.; Salami, M. Good bacteria, oxidative stress and neurological disorders: Possible therapeutical considerations. Life Sci. 2022, 301, 120605.
  92. Luca, M.; Di Mauro, M.; Di Mauro, M.; Luca, A. Gut Microbiota in Alzheimer’s Disease, Depression, and Type 2 Diabetes Mellitus: The Role of Oxidative Stress. Oxid. Med. Cell. Longev. 2019, 2019, 4730539.
  93. Ho, S.T.; Hsieh, Y.T.; Wang, S.Y.; Chen, M.J. Improving effect of a probiotic mixture on memory and learning abilities in d-galactose-treated aging mice. J. Dairy Sci. 2019, 102, 1901–1909.
  94. Wu, S.; Liu, X.; Jiang, R.; Yan, X.; Ling, Z. Roles and Mechanisms of Gut Microbiota in Patients With Alzheimer’s Disease. Front. Aging Neurosci. 2021, 13, 650047.
  95. Minter, M.R.; Zhang, C.; Leone, V.; Ringus, D.L.; Zhang, X.; Oyler-Castrillo, P.; Musch, M.W.; Liao, F.; Ward, J.F.; Holtzman, D.M.; et al. Antibiotic-induced perturbations in gut microbial diversity influences neuro-inflammation and amyloidosis in a murine model of Alzheimer’s disease. Sci. Rep. 2016, 6, 30028.
  96. Friedland, R.P.; Chapman, M.R. The role of microbial amyloid in neurodegeneration. PLoS Pathog. 2017, 13, e1006654.
  97. Yin, Y.N.; Yu, Q.F.; Fu, N.; Liu, X.W.; Lu, F.G. Effects of four Bifidobacteria on obesity in high-fat diet induced rats. World J. Gastroenterol. 2010, 16, 3394–3401.
  98. Bo, T.B.; Wen, J.; Zhao, Y.C.; Tian, S.J.; Zhang, X.Y.; Wang, D.H. Bifidobacterium pseudolongum reduces triglycerides by modulating gut microbiota in mice fed high-fat food. J. Steroid Biochem. Mol. Biol. 2020, 198, 105602.
  99. Jaunmuktane, Z.; Brandner, S. Invited Review: The role of prion-like mechanisms in neurodegenerative diseases. Neuropathol. Appl. Neurobiol. 2020, 46, 522–545.
  100. Morales, R.; Moreno-Gonzalez, I.; Soto, C. Cross-seeding of misfolded proteins: Implications for etiology and pathogenesis of protein misfolding diseases. PLoS Pathog. 2013, 9, e1003537.
  101. Frost, B.; Diamond, M.I. Prion-like mechanisms in neurodegenerative diseases. Nat. Rev. Neurosci. 2010, 11, 155–159.
  102. Poewe, W.; Seppi, K.; Tanner, C.M.; Halliday, G.M.; Brundin, P.; Volkmann, J.; Schrag, A.E.; Lang, A.E. Parkinson disease. Nat. Rev. Dis. Prim. 2017, 3, 17013.
  103. Ferrarin, M.; Carpinella, I.; Rabuffetti, M.; Calabrese, E.; Mazzoleni, P.; Nemni, R. Locomotor disorders in patients at early stages of Parkinson’s disease: A quantitative analysis. In Proceedings of the 2006 International Conference of the IEEE Engineering in Medicine and Biology Society, New York, NY, USA, 30 August–3 September 2006; pp. 1224–1227.
  104. Lu, G.; Wang, Y.; Shi, Y.; Zhang, Z.; Huang, C.; He, W.; Wang, C.; Shen, H.M. Autophagy in health and disease: From molecular mechanisms to therapeutic target. MedComm 2022, 3, e150.
  105. Braak, H.; Rüb, U.; Gai, W.P.; 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. 2003, 110, 517–536.
  106. Ferreiro, A.; Crook, N.; Gasparrini, A.J.; Dantas, G. Multiscale Evolutionary Dynamics of Host-Associated Microbiomes. Cell 2018, 172, 1216–1227.
  107. Sveinbjornsdottir, S. The clinical symptoms of Parkinson’s disease. J. Neurochem. 2016, 139 (Suppl. S1), 318–324.
  108. Fyfe, I. Movement disorders: Comparison of cognitive impairment in Parkinson disease and essential tremor. Nat. Rev. Neurol. 2017, 13, 260.
  109. Postuma, R.B.; Lang, A.E.; Munhoz, R.P.; Charland, K.; Pelletier, A.; Moscovich, M.; Filla, L.; Zanatta, D.; Rios Romenets, S.; Altman, R.; et al. Caffeine for treatment of Parkinson disease: A randomized controlled trial. Neurology 2012, 79, 651–658.
  110. Dias, V.; Junn, E.; Mouradian, M.M. The role of oxidative stress in Parkinson’s disease. J. Park. Dis. 2013, 3, 461–491.
  111. Stefanis, L. α-Synuclein in Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2012, 2, a009399.
  112. Emamzadeh, F.N. Alpha-synuclein structure, functions, and interactions. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2016, 21, 29.
  113. Simon, D.K.; Tanner, C.M.; Brundin, P. Parkinson Disease Epidemiology, Pathology, Genetics, and Pathophysiology. Clin. Geriatr. Med. 2020, 36, 1–12.
  114. Whitton, P.S. Inflammation as a causative factor in the aetiology of Parkinson’s disease. Br. J. Pharmacol. 2007, 150, 963–976.
  115. Gómez-Benito, M.; Granado, N.; García-Sanz, P.; Michel, A.; Dumoulin, M.; Moratalla, R. Modeling Parkinson’s Disease With the Alpha-Synuclein Protein. Front. Pharmacol. 2020, 11, 356.
  116. Menozzi, E.; Macnaughtan, J.; Schapira, A.H.V. The gut-brain axis and Parkinson disease: Clinical and pathogenetic relevance. Ann. Med. 2021, 53, 611–625.
  117. Hwang, O. Role of oxidative stress in Parkinson’s disease. Exp. Neurobiol. 2013, 22, 11–17.
  118. Jenner, P.; Olanow, C.W. Oxidative stress and the pathogenesis of Parkinson’s disease. Neurology 1996, 47, S161–S170.
  119. Puspita, L.; Chung, S.Y.; Shim, J.W. Oxidative stress and cellular pathologies in Parkinson’s disease. Mol. Brain 2017, 10, 53.
  120. Tong, H.; Zhang, X.; Meng, X.; Lu, L.; Mai, D.; Qu, S. Simvastatin Inhibits Activation of NADPH Oxidase/p38 MAPK Pathway and Enhances Expression of Antioxidant Protein in Parkinson Disease Models. Front. Mol. Neurosci. 2018, 11, 165.
  121. Han, C.; Shen, H.; Yang, Y.; Sheng, Y.; Wang, J.; Li, W.; Zhou, X.; Guo, L.; Zhai, L.; Guan, Q. Antrodia camphorata polysaccharide resists 6-OHDA-induced dopaminergic neuronal damage by inhibiting ROS-NLRP3 activation. Brain Behav. 2020, 10, e01824.
  122. Pfeiffer, R.F. Gastrointestinal dysfunction in Parkinson’s disease. Clin. Neurosci. 1998, 5, 136–146.
  123. Lubomski, M.; Davis, R.L.; Sue, C.M. The gut microbiota: A novel therapeutic target in Parkinson’s disease? Park. Relat. Disord. 2019, 66, 265–266.
  124. Chai, X.Y.; Diwakarla, S.; Pustovit, R.V.; McQuade, R.M.; Di Natale, M.; Ermine, C.M.; Parish, C.L.; Finkelstein, D.I.; Furness, J.B. Investigation of nerve pathways mediating colorectal dysfunction in Parkinson’s disease model produced by lesion of nigrostriatal dopaminergic neurons. Neurogastroenterol. Motil. Off. J. Eur. Gastrointest. Motil. Soc. 2020, 32, e13893.
  125. Perez-Pardo, P.; Dodiya, H.B.; Engen, P.A.; Forsyth, C.B.; Huschens, A.M.; Shaikh, M.; Voigt, R.M.; Naqib, A.; Green, S.J.; Kordower, J.H.; et al. Role of TLR4 in the gut-brain axis in Parkinson’s disease: A translational study from men to mice. Gut 2019, 68, 829–843.
  126. Rietdijk, C.D.; Perez-Pardo, P.; Garssen, J.; van Wezel, R.J.; Kraneveld, A.D. Exploring Braak’s Hypothesis of Parkinson’s Disease. Front. Neurol. 2017, 8, 37.
  127. Grassi, D.; Howard, S.; Zhou, M.; Diaz-Perez, N.; Urban, N.T.; Guerrero-Given, D.; Kamasawa, N.; Volpicelli-Daley, L.A.; LoGrasso, P.; Lasmézas, C.I. Identification of a highly neurotoxic α-synuclein species inducing mitochondrial damage and mitophagy in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2018, 115, E2634–E2643.
  128. Sampson, T.R.; Challis, C.; Jain, N.; Moiseyenko, A.; Ladinsky, M.S.; Shastri, G.G.; Thron, T.; Needham, B.D.; Horvath, I.; Debelius, J.W.; et al. A gut bacterial amyloid promotes α-synuclein aggregation and motor impairment in mice. eLife 2020, 9, e53111.
  129. Yang, X.; Qian, Y.; Xu, S.; Song, Y.; Xiao, Q. Longitudinal Analysis of Fecal Microbiome and Pathologic Processes in a Rotenone Induced Mice Model of Parkinson’s Disease. Front. Aging Neurosci. 2017, 9, 441.
  130. Ostojic, S.M. Inadequate Production of H(2) by Gut Microbiota and Parkinson Disease. Trends Endocrinol. Metab. TEM 2018, 29, 286–288.
  131. Ohta, S. Molecular hydrogen as a preventive and therapeutic medical gas: Initiation, development and potential of hydrogen medicine. Pharmacol. Ther. 2014, 144, 1–11.
  132. Dobson, R.; Giovannoni, G. Multiple sclerosis—A review. Eur. J. Neurol. 2019, 26, 27–40.
  133. Derfuss, T.; Linington, C.; Hohlfeld, R.; Meinl, E. Axo-glial antigens as targets in multiple sclerosis: Implications for axonal and grey matter injury. J. Mol. Med. 2010, 88, 753–761.
  134. Pender, M.P.; Greer, J.M. Immunology of multiple sclerosis. Curr. Allergy Asthma Rep. 2007, 7, 285–292.
  135. Tommasin, S.; Giannì, C.; De Giglio, L.; Pantano, P. Neuroimaging Techniques to Assess Inflammation in Multiple Sclerosis. Neuroscience 2019, 403, 4–16.
  136. Biernacki, T.; Sandi, D.; Bencsik, K.; Vécsei, L. Kynurenines in the Pathogenesis of Multiple Sclerosis: Therapeutic Perspectives. Cells 2020, 9, 1564.
  137. Mahad, D.H.; Trapp, B.D.; Lassmann, H. Pathological mechanisms in progressive multiple sclerosis. Lancet Neurol. 2015, 14, 183–193.
  138. Tobore, T.O. Oxidative/Nitroxidative Stress and Multiple Sclerosis. J. Mol. Neurosci. MN 2021, 71, 506–514.
  139. Schepici, G.; Silvestro, S.; Bramanti, P.; Mazzon, E. The Gut Microbiota in Multiple Sclerosis: An Overview of Clinical Trials. Cell Transplant. 2019, 28, 1507–1527.
  140. Cosorich, I.; Dalla-Costa, G.; Sorini, C.; Ferrarese, R.; Messina, M.J.; Dolpady, J.; Radice, E.; Mariani, A.; Testoni, P.A.; Canducci, F.; et al. High frequency of intestinal T(H)17 cells correlates with microbiota alterations and disease activity in multiple sclerosis. Sci. Adv. 2017, 3, e1700492.
  141. Berer, K.; Gerdes, L.A.; Cekanaviciute, E.; Jia, X.; Xiao, L.; Xia, Z.; Liu, C.; Klotz, L.; Stauffer, U.; Baranzini, S.E.; et al. Gut microbiota from multiple sclerosis patients enables spontaneous autoimmune encephalomyelitis in mice. Proc. Natl. Acad. Sci. USA 2017, 114, 10719–10724.
  142. Cekanaviciute, E.; Yoo, B.B.; Runia, T.F.; Debelius, J.W.; Singh, S.; Nelson, C.A.; Kanner, R.; Bencosme, Y.; Lee, Y.K.; Hauser, S.L.; et al. Gut bacteria from multiple sclerosis patients modulate human T cells and exacerbate symptoms in mouse models. Proc. Natl. Acad. Sci. USA 2017, 114, 10713–10718.
  143. Amato, M.P.; Derfuss, T.; Hemmer, B.; Liblau, R.; Montalban, X.; Soelberg Sørensen, P.; Miller, D.H. Environmental modifiable risk factors for multiple sclerosis: Report from the 2016 ECTRIMS focused workshop. Mult. Scler. 2018, 24, 590–603.
  144. Riccio, P.; Rossano, R.; Larocca, M.; Trotta, V.; Mennella, I.; Vitaglione, P.; Ettorre, M.; Graverini, A.; De Santis, A.; Di Monte, E.; et al. Anti-inflammatory nutritional intervention in patients with relapsing-remitting and primary-progressive multiple sclerosis: A pilot study. Exp. Biol. Med. 2016, 241, 620–635.
  145. Cignarella, F.; Cantoni, C.; Ghezzi, L.; Salter, A.; Dorsett, Y.; Chen, L.; Phillips, D.; Weinstock, G.M.; Fontana, L.; Cross, A.H.; et al. Intermittent Fasting Confers Protection in CNS Autoimmunity by Altering the Gut Microbiota. Cell Metab. 2018, 27, 1222–1235.e1226.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , ,
View Times: 388
Revisions: 2 times (View History)
Update Date: 30 Nov 2022
1000/1000
Video Production Service