Open access peer-reviewed chapter

Reactive Oxygen Species in Neurodegenerative Diseases: Implications in Pathogenesis and Treatment Strategies

Written By

Johnson Olaleye Oladele, Adenike T. Oladiji, Oluwaseun Titilope Oladele and Oyedotun M. Oyeleke

Submitted: 18 June 2021 Reviewed: 18 August 2021 Published: 24 September 2021

DOI: 10.5772/intechopen.99976

From the Edited Volume

Reactive Oxygen Species

Edited by Rizwan Ahmad

Chapter metrics overview

470 Chapter Downloads

View Full Metrics

Abstract

Neurodegenerative diseases are debilitating disorders which compromise motor or cognitive functions and are rapidly becoming a global communal disorder with over 46.8 million people suffering dementia worldwide. Aetiological studies have showed that people who are exposed to agricultural, occupational and environmental toxic chemicals that can interfere and degenerate dopaminergic neurons are prone to developing neurodegenerative diseases such as Parkinson Disease. The complex pathogenesis of the neurodegenerative diseases remains largely unknown; however, mounting evidence suggests that oxidative stress, neuroinflammation, protein misfolding, and apoptosis are the hallmarks of the diseases. Reactive oxygen species (ROS) are chemically reactive molecules that have been implicated in the pathogenesis of neurodegenerative diseases. ROS play a critical role as high levels of oxidative stress are commonly observed in the brain of patients with neurodegenerative disorders. This chapter focus on the sources of ROS in the brain, its involvement in the pathogenesis of neurodegenerative diseases and possible ways to mitigate its damaging effects in the affected brain.

Keywords

  • oxidative neuronal damage
  • neuroinflammation
  • oxidative stress

1. Introduction

Neurodegenerative diseases are debilitating disorders which compromise motor or cognitive functions and are rapidly becoming a global communal disorder with over 46.8 million people suffering dementia worldwide. They are characterised by progressive damage in neural cells and neuronal loss. Common neurodegenerative diseases include amyotrophic lateral sclerosis, Parkinson’s disease, Alzheimer’s disease, Huntington’s disease, and spinocerebellar ataxia [1]. These diseases represent major health challenges especially in the ageing population [2]. For instance, PD is the second most prevalent neurodegenerative disease affecting 1 to 2% of the population above age of 65 while AD is ranked the top 6 leading causes of death in the United States [3, 4].

It is estimated that more than 10 million individuals with the disease will be domiciled in the top 10 most populous nation in the world by 2030. In Nigeria, the most populous nation in Africa, neurodegenerative disease related cases have a significant impact on the overall hospital frequency of neurological cases reported [5]. Some of the characterised clinical features of these diseases include bradykinesia, rigidity, postural instability, resting tremor, prolonged reaction times, and freezing of gait, which may degenerate to tightened facial expression and unconscious facial movement [6, 7]. Aetiological reports have documented that individual who are exposed to industrial, occupational and environmental toxic chemicals that can interfere with the functions of the central nervous system and degenerate dopaminergic neurons are prone to developing neurodegenerative diseases such as Alzheimer’s disease, Parkinson disease [8, 9].

The complex pathogenesis of the neurodegenerative diseases remains largely unknown; however, mounting evidence suggests that oxidative stress, neuroinflammation, protein misfolding, and apoptosis are the hallmarks of the diseases (Figure 1). ROS may play a critical role as high levels of oxidative stress are commonly observed in the brain of patients with neurodegenerative conditions [10]. Reactive oxygen species (ROS) are chemically reactive molecules that have been implicated in the pathogenesis of neurodegenerative diseases. They are naturally generated within the biological system, playing significant functions in mediating cellular activities including stressor responses, cell survival, and inflammation. They also play pivotal role in the pathogenesis of many diseases such as cancer, allergy, muscle dysfunction, and cardiovascular disorders [11, 12]. Due to their reactivity, Presence of ROS in high quantity may lead to oxidative stress and ultimately cell death if left uncontrolled or treated. Oxidative stress is defined as the disruption of balance between pro-oxidant and antioxidant levels in biological systems [11].

Figure 1.

Possible involvement of oxidative stress, apoptosis, and neuroinflammation in pathogenesis of neurodegenerative diseases.

A number of experimental studies have been carried out to elucidate the significances of oxidative stress in neurodegenerative diseases [13, 14]. ROS may not be sufficient itself to induce neurodegenerative diseases but they appear to exacerbate the diseases’ progression through oxidative macromolecule damage and interaction with mitochondria [10]. Interestingly, neuronal cells have been identified to be vulnerable to oxidative damage due to their high oxygen consumption, high polyunsaturated fatty acid content in membranes, and weak antioxidant defence [15]. Under basal or unstressed physiological conditions, free radicals and ROS generated from mitochondria, NADPH oxidase (Nox), and xanthine oxidase are kept at relatively low levels by endogenous antioxidants [11]. Nevertheless, abnormal mitochondrial function and/or neuro-inflammation can alter the redox status and interrupt the balance [15]. Accumulation of misfolded proteins is part of the hallmark of pathogenesis of some neurodegenerative diseases such as Alzheimer disease and Parkinson disease (Figure 2). The aggregation of these misfolded or modified proteins can in turn triggers inflammatory response in the brain, which induces marked ROS release and subsequent oxidative stress [16]. Mitochondrial dysfunction with concomitant aberrant ROS secretion is strongly associated with neurodegenerative disorders [17]. For instance, mutant huntingtin (mHTT) in HD may directly interact with mitochondria causing compromised and alteration in energy supply and increased production of ROS [18].

Figure 2.

Molecular mechanisms underlying pathogenesis of Parkinson’s disease and Alzheimer’s disease.

Another key player in the pathogenesis of neurodegenerative diseases is neuroinflammation. The existence of neuroinflammatory processes in human brain has also been confirmed during autopsy on a molecular basis. Mogi and colleagues reported an increase in concentrations of TNFα, β2-microglobulin, epidermal growth factor (EGF), transforming growth factor α (TGFα), TGFβ1, and interleukins 1β, 6, and 2 in the striatum of patients with Parkinson’s disease [19, 20, 21, 22]. TNFα, interleukin 1β, and interferon γ were also detected in the effects indirectly. Proinflammatory cytokines, such as TNFα, interleukin 1β, and interferon γ, can induce the expression of the inducible form of nitric oxide synthase (iNOS) [23, 24] or cyclooxygenase 2 (COX2) [25]. These enzymes produce toxic reactive species. Other enzymes involved in neuroinflammatory processes mediated by oxidative stress such as myeloperoxidase, NADPH oxidase, and COX2, also have increased concentrations in neurodegenerative diseases [26].

Apoptosis has been implicated as the major pathway involved in the progressive neuronal cell death/loss observed in neurodegenerative diseases. Degeneration of one or more nerve cell populations is a major feature in many acute and chronic neurological diseases. Many criteria for apoptotic cell death are also fulfilled during the course of chronic neurodegenerative diseases. Therefore, the development of new therapeutic strategies for the treatment of neurodegenerative diseases requires an understanding of the molecular mechanisms underlying neuronal apoptosis. Extrinsic and intrinsic apoptosis pathways and several possible avenues for crosstalk between them can be distinguished. Whereas the extrinsic pathway is initiated by cell surface activation of cytokine receptors of the tumour necrosis factor (TNF) family, the intrinsic pathway depends on the integrity and function of mitochondria within the cell [27].

Various evidences from biochemical, genetic, cellular, and neuropathological studies have shown that protein misfolding, oligomerization, and accumulation in the brain are the main events triggering pathological abnormalities responsible for neurodegenerative diseases [28, 29]. The proteins most commonly implicated in the accumulation of cerebral misfolded aggregates in neurodegenerative diseases include: amyloid-beta (Aβ) in Alzheimer disease; tau in Alzheimer disease, frontotemporal dementia, corticobasal degeneration, progressive supranuclear palsy, argyrophilic grain disease, and chronic traumatic encephalopathy; alpha-synuclein (α-Syn) in PD, multiple system atrophy, and dementia with Lewy bodies; TAR DNA-binding protein 43 (TDP-43) in amyotrophic lateral sclerosis and frontotemporal frontotemporal dementia; and prion proteins in PrDs (i.e., Creutzfeldt–Jakob disease (CJD), bovine spongiform encephalopathy, chronic wasting disease, and scrapie). Despite the fact that the protein aggregates involved in distinct neurodegenerative diseases are different, the process of protein misfolding, its intermediates, end-products, and main features are remarkably similar [30].

Considering the pivotal roles of oxidative stress, neuroinflammation, protein misfolding, and apoptosis in neurodegenerative diseases (Figure 1), the manipulation of major key players in each of the pathological mechanisms may represent a promising treatment option to slow down neurodegeneration and alleviate associated symptoms. This chapter examine the role of reactive oxygen species (ROS) and oxidative stress in the pathogenesis and progression of neurodegenerative diseases. This chapter focus on the sources of ROS in the brain, its involvement in the pathogenesis of neurodegenerative diseases and possible ways to mitigate its damaging effects in the brain.

Advertisement

2. Role of oxidative stress in pathogenesis of Parkinson’s disease (PD)

PD is the second most common neurodegenerative disorder, characterised by the degeneration of dopaminergic neurons in the brain’s substantia nigra pars compacta [31]. PD affects around 1–2 percent of the population over the age of 65, and the prevalence rises to 4% in people over the age of 85 [32]. Overabundance of ROS or other free radicals has been linked to the pathological mechanism underlying dopaminergic neuron degeneration. Mitochondrial dysfunction or inflammation may both cause excessive ROS production [10]. The proper role of redox-sensitive signalling proteins in neuron cells, as well as neuronal survival, is dependent on maintaining redox homeostasis [33]. Mitochondria in neurons and glia are the main sources of ROS in the brain [10]. The production of these free radicals is exacerbated in PD due to neuroinflammation, dopamine degradation, mitochondrial dysfunction, ageing, GSH depletion, and high levels of iron or Ca2+ [10].

Consequently, when people with PD are exposed to environmental factors including pesticides, neurotoxins, and dopamine, ROS deposition may be exacerbated [34]. This is supported by a strong link between pesticide exposure and an increased risk of Parkinson’s disease [34]. ROS have been shown to contribute significantly to dopaminergic neuronal loss [10]. Other research has indicated that the loss of dopaminergic neurons is linked to the existence of neuromelanin, since highly pigmented neurons are more vulnerable to damage [35]. The formation of neuromelanin appears to be related to dopamine auto-oxidation, a process induced by ROS overproduction [35].

Neurodegeneration produces reactive oxygen species (ROS), which can destroy key cellular proteins and disrupt lipid membranes, leading in oxidative stress. Mitochondrial dysfunction increases free radical generation in the respiratory chain [10]. Parkinson’s disease has been linked to deficiencies in mitochondrial complex I in particular. Certainly, a significant portion of the unfavourable neuronal apoptosis seen in Parkinson’s disease is due to a complex I deficiency [36]. A mutation in the PTEN-induced putative kinase 1 gene is associated to this impairment (PINK1). PINK1 is a protein found in all human tissues that plays a key role in keeping mitochondrial membrane potential and preventing oxidative stress [36]. The PINK1 mutation is linked to the onset of Parkinson’s disease [36]. Mutations of leucine-rich repeat kinase 2 (LRRK2), parkin, alpha-synuclein, and DJ-1 have all been linked to the pathogenesis of Parkinson’s disease. These mutations may impair mitochondrial function, resulting in an increase in reactive oxygen species (ROS) production and oxidative stress vulnerability. Mutant parkin may play key roles in the development of autosomal recessive PD due to its involvement in lowering ROS and limiting the production of neurotoxic proteins produced by ubiquitination [36]. Additionally, alpha synuclein aggregation has been demonstrated to disrupt mitochondrial complex I activities, causing ATP production impairment and mitochondrial malfunction [37]. Proteasomal dysfunction which is exacerbated by dopamine-derived ROS, has been linked to neurodegeneration in Parkinson’s disease [37].

Currently, there is no effective cure for the treatment of Parkinson’s disease, however, deeper insights into the role of ROS in the disease pathogenesis (initiation and progression) should lead to more effective treatments for PD symptoms. Many neuroprotective approaches have been discovered to minimise mitochondrial oxidative stress in dopaminergic neurons. Free radicals damage has been proven to be reduced by antioxidants [38]. GSH, ascorbic acid and tocopherol are essential antioxidants that the antioxidant lipoic acid can recycle. Secretion of GSH which enhance reduction of lipid peroxide is one of the mechanisms by which lipoic acid offered beneficial effects against oxidative damage in oxidative stress-induced mitochondrial dysfunction [39]. In an animal study, it was discovered that treatment with lipoic acid enhanced motor coordination and ATP efficiency resulting in neuroprotection [40]. Furthermore, treatment of lipoic acid in a rotenone rats’ model of parkinsonian rats showed enhanced motor performance and marked reduction in neuronal lipid peroxide in the brain [40]. Neuroprotective ability of phytochemicals and antioxidant substances including polyphenols, Ginko biloba, docosahexaenoic acid (DHA), tocopherol, ascorbic acid, and coenzyme Q10, and have all been studied in animal experiments with remarkable findings [41, 42, 43, 44, 45, 46]. However, no convincing evidence of their neuroprotective benefits has been found in human [47]. Failures of such antioxidant medications should provide future recommendations for treating PD patients with combination therapies aimed at limiting ROS production in the brain and improving mitochondrial function [48].

Advertisement

3. Role of oxidative stress in pathogenesis of Alzheimer’s disease (AD)

Alzheimer’s disease (AD) is the most common neurodegenerative disease, characterised by gradual declines in memory, behaviour, and functionality that severely limit day-to-day activities [49]. The pathophysiology of Alzheimer’s disease is primarily linked to the formation of extracellular amyloid beta (Aβ) plaques and intracellular tau neurofibrillary tangles (NFT) [50]. Plaques in the endoplasmic reticulum (ER) can deplete calcium ions (Ca2+) storage, resulting in cytosolic Ca2+ overload. Endogenous GSH levels are reduced in response to an increase in cytosolic Ca2+, and ROS will accumulate within the cells [51]. ROS-induced ROS overproduction is believed to play a critical role in the aggregation and secretion of Aβ in AD, and oxidative stress is emerging as a significant factor in the pathogenesis of AD [52]. Mitochondrial dysfunction can result in increased production of reactive oxygen species (ROS), decreased ATP production, altered Ca2+ homeostasis, and excitotoxicity. All these alterations may be implicated in the development of AD [53].

Overactivation of N-methyl-D-aspartate-type glutamate receptors (NMDARs) can cause severe oxidative stress in Alzheimer’s patients. NMDAR activation has been showed to trigger excessive Ca2+ influx by increasing cell permeability and resulting in the production of neurotoxic levels of reactive oxygen and nitrogen species (RNS) [54, 55]. JNK/stress-activated protein kinase pathways can be mediated by reactive oxygen species (ROS). The hyperphosphorylation of tau proteins and Aβ-induced cell death have both been linked to the activation of these cascades [56]. Furthermore, Aβ proteins can directly cause formation of free radicals by inducing NADPH oxidase [57]. The activation of p38 mitogen activated protein kinase (p38 MAPK) by Aβ-induced ROS overproduction modifies cellular signalling pathways and initiates tau hyperphosphorylation. Intracellular NFT formation may be caused by an abnormal aggregation of hyperphosphorylated tau proteins [58, 59]. Consequently, Aβ has been shown to play a key role in the induction of cellular apoptosis [60]. Aβ may boost the activity of calcineurin, which then activates the Bcl-2-associated death promoter, causing mitochondrial cytochrome c release [61]. Aβ can also interact directly with caspases, resulting in neuron apoptosis [61].

Environmental stress, ageing, inflammation, and certain dietary factors (e.g., redox-active metals) may all trigger an increase in Aβ output by inducing additional oxidative stress [62]. Oxidative stress is more common in the elderly, which helps to explain why older people are more susceptible to Alzheimer’s disease [62]. Increased expression of cytokines, ROS levels, and cellular toxicity are all caused by inflammation, which accelerates the development of Alzheimer’s disease [63]. Aβ deposition results in microglial activation [64]. It’s becoming clear that sustained activation of microglia results in the release of pro-inflammatory cytokines, triggering a pro-inflammatory cascade and leading to neuronal loss and damage [65]. Environmental factors such as toxins, chemicals, and radiation may cause oxidative stress [66]. The production of reactive oxygen species (ROS) increases, where there are excess iron deposits [66]. Aβ itself can interact with metal ions to generate free radicals, therefore methionine 35 plays an important role in these reactions [67]. Cu2+/Zn2+-bound Aβ has been showed to have a structure identical to superoxide dismutase (SOD), suggesting that it could have antioxidant properties [68]. As a result, Cu2+ and Zn2+ supplementation has been considered as a novel strategy to reduce Aβ-induced ROS generation and metal catalysed Aβ deposition [68].

Drugs for Alzheimer’s disease are aimed at lowering Aβ oligomers and phosphorylated tau levels, lowering oxidative stress, and regulating epigenetic changes [69]. The majority of Alzheimer’s disease therapies depend on compounds with neuroprotective, anti-inflammatory, and antioxidant properties [70]. Medications that target ROS-mediated cascades like JNK and NF-B (e.g., tocopherol, resveratrol, and rutin) have demonstrated some promising results in vitro and in vivo [49]. When using antioxidants, significant factors including reaction kinetics and bioavailability (permeability, retention in the targeted region, distribution, and transport) must be taken into account [70]. Several ROS-related neuroprotective therapeutic techniques have shown great promise in the treatment of Alzheimer’s disease. The antioxidant response element (ARE) pathway regulated by nuclear factor erythroid 2-related factor 2 (Nrf2) is known to be an important conditioned response against oxidative stress [71]. The binding of Nrf2 to ARE activates the expression of several antioxidant genes in a synchronised manner that can work together for oxidative detoxification. Weakened Nrf2-ARE pathways were observed in the brains of transgenic mice with AD symptoms, while the enhancement of Nrf2-ARE cascades using adenoviral Nrf2 gene transfer has shown protective effects against the toxicity of Aβ deposition [71]. As a result, transcriptional modulation of endogenous antioxidants could hold great promise in the treatment of Alzheimer’s disease symptoms [71].

Advertisement

4. Role of oxidative stress in pathogenesis of spinocerebellar ataxia disease

Spinocerebellar ataxia is a progressive neurodegenerative illness caused by an autosomal dominant gene. Cognitive impairments, dysarthria, osculomotor abnormalities, and ataxic gait are all well-known signs of spinocerebellar ataxia, which can lead to mortality. Based on genetic descriptions, about 20 forms of spinocerebellar ataxia have been identified [72, 73]. The main pathogenic mutation in spinocerebellar ataxia has been linked to the expansion of repeated CAG trinucleotides [74]. The mutant ataxin 1 (ATXN1) protein, which has an enlarged polyglutamine, is overexpressed as a result of the mutation from expansion of repeated CAG trinucleotides. RAR-related orphan receptor alpha, which plays a key role in Purkinje cell activities, can be affected by mutant ataxin 1. Reduced RAR-related orphan receptor alpha gene expression has been linked to cerebellar hypoplasia and ataxia [75].

Majority of spinocerebellar ataxia are thought to be genetic disorders linked to ATXN mutations, however, different pathogenic pathways involving mitochondrial malfunction have been hypothesised [75]. Hakonen et al. [76] reported mitochondrial DNA depletion and respiratory complex I deficiency in the brain of infantile-onset spinocerebellar ataxia patients. Small concentration of ROS has been documented to be beneficial for cellular activities including cell signalling, nonetheless, higher concentration is dangerous to the brain being neurotoxic and have been established to cause neurodegeneration [49]. A study conducted by Stucki et al. have reported marked mitochondrial alterations and excessive accumulation of oxidative stress in the Purkinje cells of Spinocerebellar ataxia 1. It was suggested that there exists a connection between oxidative stress mediated mitochondrial impairments and the progression of spinocerebellar ataxia [75]. Similarly, the study evaluated the possible neuroprotective roles of MitoQ (a mitochondrial antioxidant) in a spinocerebellar ataxia mouse model. The result revealed long-term treatment of MitoQ markedly improved mitochondrial morphology and enhanced its functions in Purkinje cells resulting in amelioration of spinocerebellar ataxia 1-related symptoms including motor incoordination [75]. This report demonstrated the neuroprotective potential of mitochondria-targeted antioxidants as a potential treatment for spinocerebellar ataxia 1.

Similar to previous neurodegenerative diseases discussed, pathogenesis of spinocerebellar ataxia is associated with mitochondrial dysfunction [77]. For instance, Friedreich ataxia, is characterised by the absence of frataxin, an iron transporter protein located on the mitochondrial inner membrane. Decrease in the level of frataxin, leads to increase in concentration of iron in the mitochondrial matrix, thus stimulating the Fenton reaction which convert of H2O2 to ˙OH. The highly reactive ˙OH molecules can compromise the efficiency of energy production in neuron cells by causing oxidative damage to mitochondria [77]. Therefore, antioxidant supplementation, such as coenzyme Q10 and tocopherol, has been proven to increase energy production in many Friedreich ataxia patients by decreasing oxidative stress and restoring mitochondrial activity [78].

Because the brain contains so many mitochondria, mitochondrial malfunction can have a considerable deleterious impact on the nervous system. ROS are created spontaneously by the mitochondrial respiratory chain and are vital for sustaining mitochondrial function as well as brain cell resilience. However, there has been little study done to determine the potential involvement of ROS in spinocerebellar ataxia illnesses and establish optimum therapy options. More research is required to better understand the redox mechanisms driving various forms of spinocerebellar ataxias, with an emphasis on ROS-targeted therapy.

Advertisement

5. Role of oxidative stress in pathogenesis of Huntington’s disease (HD)

Huntington’s Disease, a neurological disorder is associated with unstable amplification of cytosine, adenine, and guanine (CAG) repeats in the HTT gene [79]. Development of CAG repeats within exon 1 of the huntingtin (HTT) gene results in a mutation that causes the polyglutamine tract to elongate, resulting in an HTT protein product that is prone to aggregation [79]. The mutant huntingtin (mHTT) aggregates are accrued throughout the brain of the affected persons, which can disturb transcription process and protein quality control. Those alterations are potentially responsible for the impaired cognitive functions and aberrant motor observed in HD are caused by mutant huntingtin aggregations and concomitant alterations om transcription process and protein quality control [79]. Currently available meditations for HD is palliative as it only inhibits the degree of severeness of symptoms. No meditation/remedy has effectively treated or markedly reversed or arrested the progression of the disease [79]. The mutant huntingtin has been demonstrated to suppress the expression of peroxisome proliferator-activated receptor-coactivator-1 and reduce the concentration of striatal mitochondrial [79, 80]. Similarly, mutant huntingtin has been documented as mutant of HD which has been implicated in the development of neuronal nuclear inclusion in HD as a result of excessive accumulation of cytoplasmic plaque [81]. Notwithstanding the well-proven connection between HD and OS, researches focused at providing treatment for the disease using antioxidant approach have not been successful [82].

A number of studies have documented that there exists link between irreversible neuronal damage and elevated oxidative markers [83]. The concentrations of well-established indicators of oxidative damage in HD such as neuron-specific enolase (NSE) and 8-hydroxy-2-deoxyguanosine (8-OHdG) have been monitored in one study to determine the benefits of neuro rehabilitation exercise [84]. Furthermore, Cu/Zn-SOD (SOD1) was documented as a possible peripheral indicator of neuronal oxidative damage, with levels considerably higher in HD patients compared to controls, implying a compensatory response to increasing oxidative levels in HD patients [84]. Nevertheless, consideration of SOD1 as an oxidative biomarker in HD remains undecided due to varied results obtained displaying different activity and concentration levels of SOD in HD [85]. After the end of the three weeks regimen neurorehabilitation exercise program, significant reduction in the levels of 8-OHdG and NSE were documented while SOD1 level remained high, indicating the possible neuroprotective role of SOD1 as an antioxidant enzyme mitigating against oxidative stress and scavenging free radicals [84]. Taken together, physical exercise was suggested for HD patients as it may possibly inhibit the disease progression and enhance redox homeostasis [86].

The consequence of HD on brain energy levels has stimulated researchers’ interest. In HD patients, reduced glucose consumption and higher lactate levels have been observed, supporting the theory that HD reduces energy levels [81]. According to new researches, oxidative damage is connected to reduced expression of the glucose transporter (GLUT)-3, which consequences lead to lactate build-up and glucose uptake inhibition [87]. Most of ATP synthesis take place via the production of proton motive force through processes of the electron transport chain [88]. mHTT has been demonstrated to perform a crucial function in mitochondrial dysfunction. Panov et al. [89] used electron microscopy to detect that the interaction between mitochondrial membranes and the N-terminal of mHTT leads to mitochondrial calcium abnormalities. Furthermore, mHTT inhibits respiratory complex II in a direct manner [90]. This alteration of the mitochondrial electron transport could lead to over production ROS with concomitant reduction in production of ATP [90].

According to a new mechanism hypothesised in 2015 for mitochondrial damage in HD, oxidative stress could incapacitate glyceraldehyde-3-phosphate dehydrogenase catalytic activities. The incapacitated glyceraldehyde-3-phosphate dehydrogenase is linked with impaired mitochondria which serve as a signalling molecule to initiate the damaged mitochondria towards lysosome engulfment through selective degradation. However, in the existence of mHTT, incapacitate glyceraldehyde-3-phosphate dehydrogenase can react unusually with the long polyglutamine of mHTT at the mitochondrial outer membrane, which result in the inhibition of degradation pathway mediated by incapacitate glyceraldehyde-3-phosphate dehydrogenase. As a result, impaired mitochondria are unable to be engulfed by lysosomes resulting into excessive accumulation of mHTT-expressing cells, thus, facilitating cell death [91]. ROS and mitochondrial alterations can both encourage the positive feedback loops, exacerbating neuronal loss in the cortex and striatum and increases oxidative stress [79]. Excessive generation of ROS and mitochondrial alterations have been implicated in the pathogenesis of HD, however, the event that occurred first remain elusive [92].

3-nitrotyrosine, thiobarbituric acid reactive substances (TBARS), and protein carbonyls are some of the other oxidative biomarkers often used in HD models [93]. Likewise, elevated levels of F2-isoprostane have been reported in the cerebrospinal fluid and brain tissue of Alzheimer’s disease and HD patients. As a result, measuring F2-isoprostane could be a useful way to assess the relevance of oxidative stress in HD patients. It’s worth noting that F2-isoprostane levels between the HD and control groups may overlap in the early stages of HD development [94]. Thus, interpretation of modifications of oxidative biomarkers in HD should be done with caution due to involvement of oxidative stress in other pathological conditions such as ageing, cancer, and soon. Additionally, oxidative biomarkers alterations levels may not reveal adequate evidence on whether the oxidative alterations perform a significant role on the neuronal cell death or disease pathogenesis [94]. The use of more sensitive and specific indicators or biomarkers would be essential to give detailed information and elucidate the specific functions performed by free radical and oxidative stress in pathogenesis of neurodegenerative diseases, which will provide a mechanistic approach to finding a suitable drug candidate for the effective treatment of HD.

Advertisement

6. Role of oxidative stress in pathogenesis of amyotrophic lateral sclerosis

Amyotrophic lateral sclerosis is a disease in which motor neurons in the anterior horn of the spinal cord gradually diminish [95]. Depending on whether there is a strongly outlined inherited genetic factor, amyotrophic lateral sclerosis is characterised as familial or sporadic. Sporadic amyotrophic lateral sclerosis usually appears between the ages of 50 and 60 [96]. Because the cause of sporadic amyotrophic lateral sclerosis is unknown, finding causal genes and environmental variables has been difficult. About 20% of instances of familial amyotrophic lateral sclerosis were caused by mutations in the SOD1 gene [97]. SOD1 has many activities, including posttranslational modification, energy consumption, controlling cellular respiration, and scavenging superoxide radicals (O2•–) [98]. Despite the fact that SOD malfunction results in a loss of antioxidant capacity, research suggests that genetic ablation of SOD1 in mice does not result in neurodegenerative diseases [14]. In divergence, the gain-of-function of mutant SOD1 protein has been markedly documented in the motor neuron diseases [14]. For example, a study has exhibited that mutant SOD1 can altered the amino acid biosynthesis of cells in a yeast model and induced cellular destruction, responsible for the neural degeneration in amyotrophic lateral sclerosis [99].

Rac1 is directly regulated by SOD1 via endosome connection, which then activates Nox. Redoxosomes which as Nox-containing endosomes play an essential role in NF-kB-mediated regulation of proinflammatory signals. Nox converts molecular oxygen into O2•–, which has vital functions in antibacterial activity, enzyme control, and cell signalling (Li et al., 2011). The ratio of reactive oxygen species to antioxidative molecules is balanced under normal physiological conditions. On the other hand, during pathological conditions, there is always rapid fluctuations in ROS levels and disturbances in antioxidant function, which result in elevated level of apoptosis, lipid peroxidation, and DNA damage during disease states [49]. SOD1 is an enzyme which convert O2•– into hydrogen peroxide (H2O2) and molecular oxygen. SOD1 mutants increase Nox2-dependent ROS generation, which is assumed to be the cause of motor neuron death in amyotrophic lateral sclerosis [100]. SOD1 that has been oxidised or misfolded has been found to cause mitochondrial dysfunction, which has been linked to the aetiology of sporadic amyotrophic lateral sclerosis [101].

Mutant SOD1 may enhance the progression of familial amyotrophic lateral sclerosis via the alterations of signal transduction pathways in motor neurons and in the activity of supportive glial cells [100]. SOD1, for instance, is regarded to be a key cell-signalling molecule with neuromodulatory functions. SOD1 is secreted via the microvesicular secretory pathway, according to studies in vitro and in transgenic mice models. SOD1 secreted into the environment binds to muscanaric receptors on nearby neurons, increasing intracellular Ca2+ concentration and ERK/AKT signalling [102]. SOD1 preserves motor neuron integrity by activating ERK/AKT signalling, and it has been demonstrated that SOD1 secretion can be enhanced in neurons under oxidative stress conditions [103]. Propofol conditioning treatment was demonstrated to protect the spinal cord against ischemia–reperfusion injury in rats by boosting PI3K/AKT signalling, which could be mediated by enhanced SOD1 activity [104]. Furthermore, oxidative stress can cause neuron cell death by blocking the neuroprotective IGF-I/AKT pathway, implying that the role of AKT signalling in neurodegeneration should be investigated further [105].

In conclusion, over secretion of ROS in the brain leads to oxidative stress which if not suppressed or inhibited could lead to oxidative damage of essential components of the central nervous system. This can also initiate or enhance some reactions which may have detrimental effects on the physiological functions and health of the brain. These reactions such as neuroinflammation, progressive neuronal cell loss via apoptosis if not abated can exacerbate protein misfolding and formation of protein aggregates resulting into neurodegeneration and associated neurobehavioural incompetence. Considering the pivotal roles of oxidative stress, neuroinflammation, protein misfolding, and apoptosis in neurodegenerative diseases (Figure 1), the manipulation of these major players in each of the pathological mechanisms may represent a promising treatment option to slow down neurodegeneration and alleviate associated symptoms.

References

  1. 1. Matilla-Duenas, T. Ashizawa, A. Brice et al. consensus paper: Pathological mechanisms underlying neurodegeneration in spinocerebellar ataxias, Cerebellum, vol. 13, pp. 269-302, 2014
  2. 2. M. Hamer and Y. Chida, Physical activity and risk of neurodegenerative disease: A systematic review of prospective evidence, Psychological Medicine, vol. 39, pp. 3-11, 2009
  3. 3. L. M. Bekris, I. F. Mata, and C. P. Zabetian, The genetics of Parkinson disease, Journal of Geriatric Psychiatry and Neurology, vol. 23, pp. 228-242, 2010
  4. 4. Alzheimer’s Association, 2016 Alzheimer’s disease facts and figures, Alzheimer’s and Dementia, vol. 12, pp. 459-509, 2016
  5. 5. Okubadejo NU, Ojo OO, Oshinaike OO, 2010, Clinical profile of parkinsonism and parkinson’s disease in Lagos, South Western Nigeria, BMC Neurol, 10, 1
  6. 6. Blandini F, 2013, Neural and immune mechanisms in the pathogenesis of Parkinson’s disease, J. Neuroimmune Pharmacol., 8, 189-201
  7. 7. Chaudhuri KR, Schapira AH, 2009, Non-motor symptoms of Parkinson’s disease: Dopaminergic pathophysiology and treatment, Lancet Neurol., 8, 464-474
  8. 8. Costello S, Cockburn M, Bronstein J, Zhang X, Ritz B, 2009, Parkinson’s disease and residential exposure to maneb and paraquat from agricultural applications in the central valley of California, Am. J. Epidemiol., 169, 919-926
  9. 9. Gao, H.M., Zhang, F., Zhou, H., Kam, W., Wilson, B., Hong, J.S. (2011). Neuroinflammation and alpha-synuclein dysfunction potentiate each other, driving chronic progression of neurodegeneration in a mouse model of Parkinson’s disease. Environ. Health Perspect. 119 (6), 807-814
  10. 10. V. Dias, E. Junn, and M. M. Mouradian, The role of oxidative stress in Parkinson’s disease, Journal of Parkinson’s Disease, vol. 3, pp. 461-491, 2013
  11. 11. L. Zuo, T. Zhou, B. K. Pannell, A. C. Ziegler, and T. M. Best, Biological and physiological role of reactive oxygen species – The good, the bad and the ugly, Acta Physiologica, vol. 214, pp. 329-348, 2015
  12. 12. F. He and L. Zuo, Redox roles of reactive oxygen species in cardiovascular diseases, International Journal of Molecular Sciences, vol. 16, pp. 27770-27780, 2015
  13. 13. J. St-Pierre, S. Drori, M. Uldry et al., Suppression of reactive oxygen species and neurodegeneration by the PGC-1 transcriptional coactivators, Cell, vol. 127, pp. 397-408, 2006
  14. 14. K. Hensley, M. Mhatre, S. Mou et al., On the relation of oxidative stress to neuroinflammation: Lessons learned from the G93A-SOD1 mouse model of amyotrophic lateral sclerosis, Antioxidants and Redox Signaling, vol. 8, pp. 2075-2087, 2006
  15. 15. A. C. Rego and C. R. Oliveira, Mitochondrial dysfunction and reactive oxygen species in excitotoxicity and apoptosis: Implications for the pathogenesis of neurodegenerative diseases, Neurochemical Research, vol. 28, pp. 1563-1574, 2003
  16. 16. T. Wyss-Coray and L. Mucke, Inflammation in neurodegenerative disease – A double-edged sword, Neuron, vol. 35, pp. 419-432, 2002
  17. 17. M. T. Lin and M. F. Beal, Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases, Nature, vol. 443, pp. 787-795, 2006
  18. 18. A. Ross and S. J. Tabrizi, Huntington’s disease: From molecular pathogenesis to clinical treatment, Lancet Neurology, vol. 10, pp. 83-98, 2011
  19. 19. Mogi M, Harada M, Kondo J. (1994). Interleukin-1 beta, interleukin-6, epidermal growth factor and transforming growth factor-alpha are elevated in the brain from parkinsonian patients. Neurosci Lett; 180: 147-150
  20. 20. Mogi M, Harada M, Kondo T, Narabayashi H, Riederer P, Nagatsu T. (1995). Transforming growth factor-beta 1 levels are elevated in the striatum and in ventricular cerebrospinal fl uid in Parkinson’s disease. Neurosci Lett; 193: 129-132
  21. 21. Mogi M, Harada M, Kondo T, Riederer P, Nagatsu T. (1995). Brain beta 2-microglobulin levels are elevated in the striatum in Parkinson’s disease. J Neural Transm Park Dis Dement Sect; 9: 87-92
  22. 22. Mogi M, Harada M, Riederer P, Narabayashi H, Fujita K, Nagatsu T. (1994). Tumor necrosis factor-alpha (TNF-alpha) increases both in the brain and in the cerebrospinal fluid from parkinsonian patients. Neurosci Lett; 165: 208-210
  23. 23. Chao CC, Hu S, Molitor TW, Shaskan EG, Peterson PK. (1992). Activated microglia mediate neuronal cell injury via a nitric oxide mechanism. J Immunol 149: 2736-2741
  24. 24. Boje KM, Arora PK. (1992). Microglial-produced nitric oxide and reactive nitrogen oxides mediate neuronal cell death. Brain Res; 587: 250-256
  25. 25. Vane JR, Bakhle YS, Botting RM. (2000). Cyclooxygenases 1 and 2. Annu Rev Pharmacol Toxicol; 38: 97-120
  26. 26. Knott C, Stern G, Wilkin GP. (2000). Inflammatory regulators in Parkinson’s disease: iNOS, lipocortin-1, and cyclooxygenases-1 and -2. Mol Cell Neurosci; 16: 724-739
  27. 27. Reed JC: Mechanisms of apoptosis. Am J Pathol 2000;157:1415-1430
  28. 28. Ross, C. A. and Poirier, M. A. Protein aggregation and neurodegenerative disease. Nat. Med. 10(Suppl), S10–S17 (2004)
  29. 29. Goedert, M. Neurodegeneration. Alzheimer’s and Parkinson’s diseases: The prion concept in relation to assembled Aβ, tau, and α-synuclein. Science 349, 1255555 (2015)
  30. 30. Soto, C. Unfolding the role of protein misfolding in neurodegenerative diseases. Nat. Rev. Neurosci. 4, 49-60 (2003)
  31. 31. A. L. McCormack, M. Thiruchelvam, A. B. Manning-Bog et al., Environmental risk factors and Parkinson’s disease: Selective degeneration of nigral dopaminergic neurons caused by the herbicide paraquat, Neurobiology of Disease, vol. 10, pp. 119-127, 2002
  32. 32. M. J. Farrer, Genetics of Parkinson disease: Paradigm shifts and future prospects, Nature Reviews Genetics, vol. 7, pp. 306-318, 2006
  33. 33. S. J. Chinta and J. K. Andersen, Redox imbalance in Parkinson’s disease, Biochimica et Biophysica Acta, vol. 1780, pp. 1362-1367, 2008
  34. 34. S. Gangemi, E. Gofita, C. Costa et al., Occupational and environmental exposure to pesticides and cytokine pathways in chronic diseases (review), International Journal of Molecular Medicine, vol. 38, pp. 1012-1020, 2016
  35. 35. R. Perfeito, T. Cunha-Oliveira, and A. C. Rego, Revisiting oxidative stress and mitochondrial dysfunction in the pathogenesis of Parkinson disease – Resemblance to the effect of amphetamine drugs of abuse, Free Radical Biology and Medicine, vol. 53, pp. 1791-1806, 2012
  36. 36. L. Zuo and M. S. Motherwell, The impact of reactive oxygen species and genetic mitochondrial mutations in Parkinson’s disease, Gene, vol. 532, pp. 18-23, 2013
  37. 37. G. Ganguly, S. Chakrabarti, U. Chatterjee, and L. Saso, Proteinopathy, oxidative stress and mitochondrial dysfunction: Cross talk in Alzheimer’s disease and Parkinson’s disease, drug design, Development and Therapy, vol. 11, pp. 797-810, 2017
  38. 38. N. A. Mazo, V. Echeverria, R. Cabezas et al., Medicinal plants as protective strategies against Parkinson’s disease, Current Pharmaceutical Design, vol. 23, 2017
  39. 39. P. I. Moreira, X. Zhu, X. Wang et al., Mitochondria: A therapeutic target in neurodegeneration, Biochimica et Biophysica Acta, vol. 1802, pp. 212-220, 2010
  40. 40. S. A. Zaitone, D. M. Abo-Elmatty, and A. A. Shaalan, Acetyl-L-carnitine and alpha-lipoic acid affect rotenone induced damage in nigral dopaminergic neurons of rat brain, implication for Parkinson’s disease therapy, Pharmacology, Biochemistry, and Behavior, vol. 100, pp. 347-360, 2012
  41. 41. Oladele JO., Oladele OT. Oyeleke OM., Oladiji AT. (2021). Neurological complications in COVID-19: Implications on international health security and possible interventions of phytochemicals. Contemporary developments and perspectives in international health security – Volume 2. DOI:10.5772/intechopen.96039
  42. 42. Oladele JO., Oyeleke OM., Oladele OT. Olaniyan MD. (2020). Neuroprotective mechanism of Vernonia amygdalina in a rat model of neurodegenerative diseases. Toxicology report. 7: 1223-1232. DOI:10.1016/j.toxrep.2020.09.005
  43. 43. Oyewole O.I., Olabiyi B.F. and Oladele J.O. (2017). Antioxidative potential of Ricinus communis leaf extract on cadmium-induced liver and brain toxicity in rats. UNIOSUN Journal of Sciences. 2, 2. 84-90
  44. 44. Oladele, J.O., Oyewole, O.I., Bello, O.K. and Oladele, O.T. (2017). Assessment of protective potentials of Ficus exasperata leaf on arsenate-mediated Dyslipidemia and oxidative damage in Rat’s brain. J. basic appl. Res. 3(3): 77-82
  45. 45. Oladele, JO., Adewale, OO., Oyeleke, OM., Oyewole IO., Salami, MO., Owoade G. (2020). Annona muricata protects against cadmium mediated oxidative damage in brain and liver of rats. Acta facultatis medicae Naissensis. 37 (3): 252-260. DOI:10.5937/afmnai2003252O
  46. 46. Oladele JO, Ajayi EIO, Oyeleke OM, Oladele OT, Olowookere BD, Adeniyi BM, Oyewole OI, Oladiji AT. (2020). A systematic review on COVID-19 pandemic with special emphasis on curative potentials of medicinal plants. Heliyon. 6: 1-17. DOI:10.1016/j.heliyon.2020.e04897
  47. 47. M. Etminan, S. S. Gill, and A. Samii, Intake of vitamin E, vitamin C, and carotenoids and the risk of Parkinson’s disease: A meta-analysis, Lancet Neurology, vol. 4, pp. 362-365, 2005
  48. 48. M. H. Yan, X. Wang, and X. Zhu, Mitochondrial defects and oxidative stress in Alzheimer disease and Parkinson disease, Free Radical Biology and Medicine, vol. 62, pp. 90-101, 2013
  49. 49. L. Zuo, B. T. Hemmelgarn, C. C. Chuang, and T. M. Best, The role of oxidative stress-induced epigenetic alterations in amyloid-beta production in Alzheimer’s disease, Oxidative Medicine and Cellular Longevity, vol. 2015, Article ID 604658, 13 pages, 2015
  50. 50. A. Butterfield, The 2013 SFRBM discovery award: Selected discoveries from the Butterfield laboratory of oxidative stress and its sequela in brain in cognitive disorders exemplified by Alzheimer disease and chemotherapy induced cognitive impairment, Free Radical Biology and Medicine, vol. 74, pp. 157-174, 2014
  51. 51. Ferreiro, C. R. Oliveira, and C. M. Pereira, The release of calcium from the endoplasmic reticulum induced by amyloid-beta and prion peptides activates the mitochondrial apoptotic pathway, Neurobiology of Disease, vol. 30, pp. 331-342, 2008
  52. 52. J. Bonda, X. Wang, G. Perry et al., Oxidative stress in Alzheimer disease: A possibility for prevention, Neuropharmacology, vol. 59, pp. 290-294, 2010
  53. 53. W. J. Huang, X. Zhang, and W. W. Chen, Role of oxidative stress in Alzheimer’s disease, Biomedical Reports, vol. 4, pp. 519-522, 2016
  54. 54. T. Nakamura and S. A. Lipton, Redox modulation by S-nitrosylation contributes to protein misfolding, mitochondrial dynamics, and neuronal synaptic damage in neurodegenerative diseases, Cell Death and Differentiation, vol. 18, pp. 1478-1486, 2011
  55. 55. T. Nakamura and S. A. Lipton, Preventing Ca2+−mediated nitrosative stress in neurodegenerative diseases: Possible pharmacological strategies, Cell Calcium, vol. 47, pp. 190-197, 2010
  56. 56. D. A. Patten, M. Germain, M. A. Kelly, and R. S. Slack, Reactive oxygen species: stuck in the middle of neurodegeneration, Journal of Alzheimer’s Disease, vol. 20, Supplement 2, pp. S357–S367, 2010
  57. 57. P. B. Shelat, M. Chalimoniuk, J. H. Wang et al., Amyloid beta peptide and NMDA induce ROS from NADPH oxidase and AA release from cytosolic phospholipase A2 in cortical neurons, Journal of Neurochemistry, vol. 106, pp. 45-55, 2008
  58. 58. Giraldo, A. Lloret, T. Fuchsberger, and J. Vina, Abeta and tau toxicities in Alzheimer’s are linked via oxidative stress induced p 38 activation: Protective role of vitamin E, Redox Biology, vol. 2, pp. 873-877, 2014
  59. 59. N. Bulat and C. Widmann, Caspase substrates and neurodegenerative diseases, Brain Research Bulletin, vol. 80, pp. 251-267, 2009
  60. 60. P. Agostinho, J. P. Lopes, Z. Velez, and C. R. Oliveira, Overactivation of calcineurin induced by amyloid-beta and prion proteins, Neurochemistry International, vol. 52, pp. 1226-1233, 2008
  61. 61. A. Awasthi, Y. Matsunaga, and T. Yamada, Amyloid-beta causes apoptosis of neuronal cells via caspase cascade, which can be prevented by amyloid-beta-derived short peptides, Experimental Neurology, vol. 196, pp. 282-289, 2005
  62. 62. Hamilton and C. Holscher, The effect of ageing on neurogenesis and oxidative stress in the APPswe/PS1deltaE9 mouse model of Alzheimer’s disease, Brain Research, vol. 1449, pp. 83-93, 2012
  63. 63. Holmes, C. Cunningham, E. Zotova et al., systemic inflammation and disease progression in Alzheimer disease, Neurology, vol. 73, pp. 768-774, 2009
  64. 64. T. J. Seabrook, L. Y. Jiang, M. Maier, and C. A. Lemere, Minocycline affects microglia activation, a beta deposition, and behavior in APP-tg mice, Glia, vol. 53, pp. 776-782, 2006
  65. 65. W. Y. Wang, M. S. Tan, J. T. Yu, and L. Tan, Role of pro-inflammatory cytokines released from microglia in Alzheimer’s disease, Annals of Translational Medicine, vol. 3, p. 136, 2015
  66. 66. M. Nizzari, S. Thellung, A. Corsaro et al., Neurodegeneration in Alzheimer disease: role of amyloid precursor protein and presenilin 1 intracellular signaling, Journal of Toxicology, vol. 2012, Article ID 187297, 13 pages, 2012
  67. 67. A. Butterfield and D. Boyd-Kimball, The critical role of methionine 35 in Alzheimer’s amyloid beta-peptide (1-42)-induced oxidative stress and neurotoxicity, Biochimica et Biophysica Acta, vol. 1703, pp. 149-156, 2005
  68. 68. C. Curtain, F. Ali, I. Volitakis et al., Alzheimer’s disease amyloid-beta binds copper and zinc to generate an allosterically ordered membrane-penetrating structure containing superoxide dismutase-like subunits, The Journal of Biological Chemistry, vol. 276, pp. 20466-20473, 2001
  69. 69. M. W. Dysken, P. D. Guarino, J. E. Vertrees et al., Vitamin E and memantine in Alzheimer’s disease: Clinical trial methods and baseline data, Alzheimer’s and Dementia, vol. 10, pp. 36-44, 2014
  70. 70. M. Dumont and M. F. Beal, Neuroprotective strategies involving ROS in Alzheimer disease, Free Radical Biology and Medicine, vol. 51, pp. 1014-1026, 2011
  71. 71. K. Kanninen, T. M. Malm, H. K. Jyrkkanen et al., Nuclear factor erythroid 2-related factor 2 protects against beta amyloid, Molecular and Cellular Neurosciences, vol. 39, pp. 302-313, 2008
  72. 72. M. Rossi, S. Perez-Lloret, L. Doldan et al., Autosomal dominant cerebellar ataxias: A systematic review of clinical features, European Journal of Neurology, vol. 21, pp. 607-615, 2014
  73. 73. Y. M. Sun, C. Lu, and Z. Y. Wu, Spinocerebellar ataxia: Relationship between phenotype and genotype – A review, Clinical Genetics, vol. 90, pp. 305-314, 2016
  74. 74. M. U. Manto, The wide spectrum of spinocerebellar ataxias (SCAs), Cerebellum, vol. 4, pp. 2-6, 2005
  75. 75. M. Stucki, C. Ruegsegger, S. Steiner et al., Mitochondrial impairments contribute to spinocerebellar ataxia type 1 progression and can be ameliorated by the mitochondria targeted antioxidant MitoQ, Free Radical Biology and Medicine, vol. 97, pp. 427-440, 2016
  76. 76. A. H. Hakonen, S. Goffart, S. Marjavaara et al., Infantileonset spinocerebellar ataxia and mitochondrial recessive ataxia syndrome are associated with neuronal complex I defect and mtDNA depletion, Human Molecular Genetics, vol. 17, pp. 3822-3835, 2008
  77. 77. V. Campuzano, L. Montermini, Y. Lutz et al., Frataxin is reduced in Friedreich ataxia patients and is associated with mitochondrial membranes, Human Molecular Genetics, vol. 6, pp. 1771-1780, 1997
  78. 78. R. Lodi, P. E. Hart, B. Rajagopalan et al., Antioxidant treatment improves in vivo cardiac and skeletal muscle bioenergetics in patients with Friedreich’s ataxia, Annals of Neurology, vol. 49, pp. 590-596, 2001
  79. 79. J. Gil-Mohapel, P. S. Brocardo, and B. R. Christie, The role of oxidative stress in Huntington’s disease: Are antioxidants good therapeutic candidates?, Current Drug Targets, vol. 15, pp. 454-468, 2014
  80. 80. P. Weydt, V. V. Pineda, A. E. Torrence et al., Thermoregulatory and metabolic defects in Huntington’s disease transgenic mice implicate PGC-1alpha in Huntington’s disease neurodegeneration, Cell Metabolism, vol. 4, pp. 349-362, 2006
  81. 81. Mochel and R. G. Haller, Energy deficit in Huntington disease: why it matters, The Journal of Clinical Investigation, vol. 121, pp. 493-499, 2011
  82. 82. A. Kumar and R. R. Ratan, Oxidative stress and Huntington’s disease: The good, the bad, and the ugly, Journal of Huntington’s Disease, vol. 5, pp. 217-237, 2016
  83. 83. Tunez, F. Sanchez-Lopez, E. Aguera, R. Fernandez-Bolanos, F. M. Sanchez, and I. Tasset-Cuevas, important role of oxidative stress biomarkers in Huntington’s disease, Journal of Medicinal Chemistry, vol. 54, pp. 5602-5606, 2011
  84. 84. Ciancarelli, D. AmicisDe, C. MassimoDi et al., influence of intensive multifunctional neurorehabilitation on neuronal oxidative damage in patients with Huntington’s disease, Functional Neurology, vol. 30, pp. 47-52, 2015
  85. 85. M. A. Sorolla, G. Reverter-Branchat, J. Tamarit, I. Ferrer, J. Q. Ros, and E. Cabiscol, Proteomic and oxidative stress analysis in human brain samples of Huntington disease, Free Radical Biology and Medicine, vol. 45, pp. 667-678, 2008
  86. 86. R. M. Arida, F. A. Scorza, and E. A. Cavalheiro, Role of physical exercise as complementary treatment for epilepsy and other brain disorders, Current Pharmaceutical Design, vol. 19, pp. 6720-6725, 2013
  87. 87. A. Covarrubias-Pinto, P. Moll, M. Solis-Maldonado et al., Beyond the redox imbalance: Oxidative stress contributes to an impaired GLUT3 modulation in Huntington’s disease, Free Radical Biology and Medicine, vol. 89, pp. 1085-1096, 2015
  88. 88. M. Bonora, M. R. Wieckowsk, C. Chinopoulos et al., Molecular mechanisms of cell death: Central implication of ATP synthase in mitochondrial permeability transition, Oncogene, vol. 34, p. 1608, 2015
  89. 89. A. V. Panov, C. A. Gutekunst, B. R. Leavitt et al., Early mitochondrial calcium defects in Huntington’s disease are a direct effect of polyglutamines, Nature Neuroscience, vol. 5, pp. 731-736, 2002
  90. 90. E. Bossy-Wetzel, A. Petrilli, and A. B. Knott, Mutant huntingtin and mitochondrial dysfunction, Trends in Neurosciences, vol. 31, pp. 609-616, 2008
  91. 91. Liot, J. Valette, J. Pepin, J. Flament, and E. Brouillet, Energy defects in Huntington’s disease: why “in vivo” evidence matters, Biochemical and Biophysical Research Communications, vol. 483, no. 4, pp. 1084-1095, 2017
  92. 92. C. Zuccato, M. Valenza, and E. Cattaneo, Molecular mechanisms and potential therapeutical targets in Huntington’s disease, Physiological Reviews, vol. 90, pp. 905-981, 2010
  93. 93. M. C. Polidori, P. Mecocci, S. E. Browne, U. Senin, and M. F. Beal, Oxidative damage to mitochondrial DNA in Huntington’s disease parietal cortex, Neuroscience Letters, vol. 272, pp. 53-56, 1999
  94. 94. E. Miller, A. Morel, L. Saso, and J. Saluk, Isoprostanes and neuroprostanes as biomarkers of oxidative stress in neurodegenerative diseases, Oxidative Medicine and Cellular Longevity, vol. 2014, Article ID 572491, 10 pages, 2014
  95. 95. P. Taylor, R. H. Brown Jr., and D. W. Cleveland, Decoding ALS: From genes to mechanism, Nature, vol. 539, pp. 197-206, 2016
  96. 96. C. Ingre, P. M. Roos, F. Piehl, F. Kamel, and F. Fang, Risk factors for amyotrophic lateral sclerosis, Clinical Epidemiology, vol. 7, pp. 181-193, 2015
  97. 97. Gamez, M. Corbera-Bellalta, G. Nogales et al., mutational analysis of the cu/Zn superoxide dismutase gene in a Catalan ALS population: Should all sporadic ALS cases also be screened for SOD1?, Journal of the Neurological Sciences, vol. 247, pp. 21-28, 2006
  98. 98. R. A. Saccon, R. K. Bunton-Stasyshyn, E. M. Fisher, and P. Fratta, Is SOD1 loss of function involved in amyotrophic lateral sclerosis?, Brain, vol. 136, pp. 2342-2358, 2013
  99. 99. E. L. Bastow, A. R. Peswani, D. S. J. Tarrant et al., New links between SOD1 and metabolic dysfunction from a yeast model of amyotrophic lateral sclerosis, Journal of Cell Science, vol. 129, pp. 4118-4129, 2016
  100. 100. Q. Li, N. Y. Spencer, N. J. Pantazis, and J. F. Engelhardt, Alsin and SOD1(G93A) proteins regulate endosomal reactive oxygen species production by glial cells and proinflammatory pathways responsible for neurotoxicity, The Journal of Biological Chemistry, vol. 286, pp. 40151-40162, 2011
  101. 101. E. D’Amico, P. Factor-Litvak, R. M. Santella, and H. Mitsumoto, Clinical perspective on oxidative stress in sporadic amyotrophic lateral sclerosis, Free Radical Biology and Medicine, vol. 65, pp. 509-527, 2013
  102. 102. P. Mondola, S. Damiano, A. Sasso, and M. Santillo, The cu, Zn superoxide dismutase: Not only a dismutase enzyme, Frontiers in Physiology, vol. 7, p. 594, 2016
  103. 103. S. Damiano, T. Petrozziello, V. Ucci, S. Amente, M. Santillo, and P. Mondola, Cu-Zn superoxide dismutase activates muscarinic acetylcholine M1 receptor pathway in neuroblastoma cells, Molecular and Cellular Neurosciences, vol. 52, pp. 31-37, 2013
  104. 104. Q. J. Yu and Y. Yang, Function of SOD1, SOD2, and PI3K/AKT signaling pathways in the protection of propofol on spinal cord ischemic reperfusion injury in a rabbit model, Life Sciences, vol. 148, pp. 86-92, 2016
  105. 105. D. Davila and I. Torres-Aleman, Neuronal death by oxidative stress involves activation of FOXO3 through a two-arm pathway that activates stress kinases and attenuates insulin like growth factor I signaling, Molecular Biology of the Cell, vol. 19, pp. 2014-2025, 2008

Written By

Johnson Olaleye Oladele, Adenike T. Oladiji, Oluwaseun Titilope Oladele and Oyedotun M. Oyeleke

Submitted: 18 June 2021 Reviewed: 18 August 2021 Published: 24 September 2021