IMR Press / JIN / Volume 22 / Issue 6 / DOI: 10.31083/j.jin2206153
Open Access Review
Targeting Mitochondrial Oxidative Stress: Potential Neuroprotective Therapy for Spinal Cord Injury
Show Less
1 School of Medical, Yan’an University, 716000 Yan’an, Shaanxi, China
*Correspondence: yadxgf@yau.edu.cn (Feng Gao)
J. Integr. Neurosci. 2023, 22(6), 153; https://doi.org/10.31083/j.jin2206153
Submitted: 17 May 2023 | Revised: 5 July 2023 | Accepted: 10 July 2023 | Published: 30 October 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Spinal cord injury (SCI) is a serious central nervous system (CNS) injury disease related to hypoxia-ischemia and inflammation. It is characterized by excessive reactive oxygen species (ROS) production, oxidative damage to nerve cells, and mitochondrial dysfunction. Mitochondria serve as the primary cellular origin of ROS, wherein the electron transfer chain complexes within oxidative phosphorylation frequently encounter electron leakage. These leaked electrons react with molecular oxygen, engendering the production of ROS, which culminates in the occurrence of oxidative stress. Oxidative stress is one of the common forms of secondary injury after SCI. Mitochondrial oxidative stress can lead to impaired mitochondrial function and disrupt cellular signal transduction pathways. Hence, restoring mitochondrial electron transport chain (ETC), reducing ROS production and enhancing mitochondrial function may be potential strategies for the treatment of SCI. This article focuses on the pathophysiological role of mitochondrial oxidative stress in SCI and evaluates in detail the neuroprotective effects of various mitochondrial-targeted antioxidant therapies in SCI, including both drug and non-drug therapy. The objective is to provide valuable insights and serve as a valuable reference for future research in the field of SCI.

Keywords
spinal cord injury
mitochondria oxidative stress
antioxidant therapy
mitochondrial electron transport chain
reactive oxygen stress
1. Introduction

Spinal cord injury (SCI) is a central nervous system traumatic disorder characterized by a high disability rate, resulting in impairments [1, 2]. SCI resulting from vehicular crashes, gunshot wounds, high-altitude falling injuries and heavy falling injuries is the most pr from diverse etiologies and manifests varying degrees of sensory and motor functional evalent form in clinical practice [3]. According to clinical statistics, the global annual prevalence of SCI ranges from 10.4 to 83 cases per million population [4]. Furthermore, the cost of treatment and rehabilitation care is as high as $3 million per person [5]. SCI can be categorized into primary injury and secondary injury based on its mechanism [6]. The former is attributed to extrinsic influences and various factors leading to vascular rupture and soft tissue injury surrounding the spinal cord. The latter encompasses a cascade of malign processes following the initial injury, comprising inflammatory response, oxidative stress, neuronal and glial cell apoptosis, as well as diverse intracellular organelle damage [7, 8]. Currently, the management options for SCI remain considerably restricted. Clinical intervention primarily revolves around surgical procedures aimed at alleviating compression and diminishing secondary pathological compression. Additionally, methylprednisolone, along with other pharmaceutical agents like riluzole and minocycline, is administered to mitigate inflammation and edema [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19]. The employment of high-dose methylprednisolone as a therapeutic approach remains controversial due to the prevalence of numerous potential complications, including sepsis, infection, gastrointestinal hemorrhage, pneumonia, pulmonary embolism, and corticosteroid-related adverse effects [20]. Therefore, looking for new treatment strategies and an in-depth study of their mechanism is the main focus of our efforts.

Mitochondria are organelles that are enveloped by two membranes and participate in a wide range of regulatory functions. Their primary role is to generate cellular energy in the form of adenosine triphosphate (ATP) via the respiratory chain (ETC) reaction, which provides the requisite energy for cellular activities [21]. SCI can induce mitochondrial dysfunction [22]. One of the primary culprits responsible for secondary injury is oxidative stress. Mitochondrial oxidative stress can result in impaired mitochondrial function and impact cellular signal transduction. Excessive production of mitochondrial reactive oxygen species (mtROS) has been shown to directly impact biological macromolecules such as mitochondrial DNA (mtDNA), proteins, and lipids, thereby compromising the integrity of mitochondrial structure and function [23]. ETC represents the primary source of mtROS. Thus, remedying respiratory chain abnormalities and mitigating excessive mtROS production can ameliorate mitochondrial oxidative stress. Researchers have proposed several drug and non-drug treatments that target mitochondria, which have been demonstrated to counteract mitochondrial oxidative stress and enhance motor function in SCI models. This article provides an in-depth analysis of the pathophysiological role of mitochondrial oxidative stress in SCI and the research advancements pertaining to treatments of SCI-related mitochondrial dysfunction.

2. Mechanism of Action of Mitochondrial Oxidative Stress and SCI

In normal mitochondria, the oxidation and antioxidant systems exist in a state of dynamic equilibrium. When there is an overactivity of the mitochondrial oxidation system or a deficiency in the antioxidant system, oxidative stress can occur. ROS are the most significant components of the mitochondrial oxidation system and are primarily present as the superoxide anion (O2-), hydrogen peroxide (H2O2), and hydroxyl radical (OH) [24]. Under normal circumstances, ROS in the body is kept at a minimal level to ensure normal mitochondrial function.

ETC localized in the mitochondria’s inner membrane is the primary site of ROS generation. It comprises four membrane-bound complexes and two mobile electron carriers, namely, coenzyme Q (CoQ) and cytochrome C. Two distinct pathways of electron transfer exist within the ETC: the reduced form of nicotinamide adenine dinucleotide (NADH)-dependent complexes I/III/IV and the succinate-dependent complexes II/III/IV [25]. The electron transfer in mitochondria can lead to electron leakages that result in the formation of O2- or H2O2 when some electrons cannot be transferred normally and react with oxygen. Complexes I and III in the ETC are known to contribute to this process [26]. Complex I is a site of prominent O2- production, which occurs at two distinct locations. Firstly, the flavin mononucleotide (FMN) cofactor situated within the complex acts as an electron acceptor, receiving electrons from NADH. Secondly, the CoQ binding site at the end of Fe-S transfers two electrons to CoQ, culminating in the formation of O2- [27]. During the forward electron transfer process, the electrons donated by NADH undergo a complete reduction of the FMN center within complex I, leading to the formation of O2- through reaction with oxygen. The extent of O2- generation at the FMN center is governed by the NADH/nicotinamide adenine dinucleotide-oxidized form (NAD+) ratio and remains relatively low under normal physiological conditions. Complex III, particularly at the CoQH2 oxidation site, contributes to the production of O2- in the ETC. Upon binding to complex III, CoQH2 donates its two electrons to the Fe-S center and cytochrome C. This electron transfer sequence results in the formation of an unsteady Q species on CoQ, which reacts with oxygen to produce O2- and generate ROS. Nevertheless, the pace of ROS generation at this locus stays modest during physiological circumstances [28].

At the 24-hour time point following contusion, a notable decline in respiratory control ratio (RCR) function is evident. Mitochondria are isolated from both the sham and injured animal groups at 6 to 24 hours post-injury, revealing an elevation in levels of 3-nitrotyrosine (3-NT), 4-hydroxynonenal (HNE), and protein carbonyls within the mitochondria subsequent to injury. SCI is accompanied by a detrimental cycle of heightened mtROS, culminating in amplified oxidative damage, ultimately leading to an escalation of mtROS to a pathological threshold [29]. Research findings indicate that within the rat SCI model, there is a notable 48% elevation in mitochondrial ROS levels after 4 hours of SCI compared to the control group. Furthermore, 24 hours post-injury, there is a substantial increase in catalase (CAT) activity and glutathione (GSH) concentration. These observations signify anomalous mitochondrial functionality and ROS levels subsequent to SCI [30]. Moreover, the ETC within the mitochondria is the first component to be adversely affected after SCI [31, 32]. Impairment of the ETC adversely affects the function of complex I, resulting in NADH accumulation, elevation of the NADH/NAD+ ratio, and increased O2- production. Complex I can generate a substantial quantity of O2- through the reverse electron transport (RET) mechanism. RET is activated under conditions of low CoQ pool height and high proton motive force, thereby forcing electrons to revert from CoQH2 to complex I and reducing NAD to NADH at the FMN site [33]. Moreover, complex III inhibition augments ROS generation. While complex II is not usually considered a significant contributor to ROS, some studies suggest that O2- can also be generated following complex II impairment, but the underlying mechanism requires further investigation [25].

Apart from the aforementioned sources, ROS can also be formed by enzymatic action, including glycerol aldehyde-3-phosphate dehydrogenase (GAPDH), monoamine oxidase (MAO), cytochrome b5 reductase (Cb5R), and cytochrome P450 (p450). Meanwhile, it can also be composed of matrix enzymes and complexes, including aconitase, α-ketoglutarate dehydrogenase (αKGDH), and pyruvate dehydrogenase (PDH) [34, 35, 36, 37]. In addition to the oxidation system, there is also an antioxidant system that maintains mitochondrial homeostasis. Notably, antioxidative proteins like glutathione peroxidase (GPX) and peroxiredoxin (Prxs) synergize to exert a pivotal function in the antioxidant system [38]. When SCI occurs, there is excessive production of ROS, leading to the disruption of the cell membrane structure. The maintenance of a stable state in the mitochondrial structure becomes challenging, resulting in impaired function. This directly impacts the cellular energy supply and ultimately leads to cell death. Therefore, it is imperative to preserve the stability of ETC function and regulate ROS levels within a safe concentration to suppress mitochondrial oxidative stress.

3. Role of Mitochondrial Oxidative Stress in Spinal Cord Injury

The preceding discussion provides a comprehensive account of how mitochondrial oxidative stress is generated in the context of spinal cord injury, including abnormalities in the ETC and the release of mtROS. Mitochondria primarily serve the vital physiological function of generating a substantial amount of ATP through oxidative phosphorylation (OXPHOS) to meet the body’s energy demands. Furthermore, mitochondria are involved in regulating various cellular metabolic processes through intricate mechanisms, such as maintaining intracellular calcium homeostasis, governing programmed cell death, and modulating immune responses. However, in the presence of oxidative stress within mitochondria, detrimental effects are present, such as the opening of the mitochondrial permeability transition pore (mPTP), aberrant mitophagy, inflammation, and mitochondrial DNA (mtDNA) damage. These mechanisms contribute to the worsening of SCI. The subsequent section will delve into a detailed exploration of the role played by mitochondrial oxidative stress in SCI (Fig. 1).

Fig. 1.

Role of mitochondrial oxidative stress in spinal cord injury. Following spinal cord injury, electron transfer within the mitochondria gives rise to a fraction of electrons engaging in electron leakage with oxygen, leading to the formation of superoxide anion (O2-) or hydrogen peroxide (H2O2). This phenomenon predominantly occurs in complexes I and III of the electron transport chain (ETC), resulting in elevated levels of reactive oxygen species (ROS) and subsequent initiation of the mitochondrial permeability transition pore (mPTP). Calcium ion (Ca2+) overload and increased ROS concentrations following spinal cord injury further promote the persistent opening of the mPTP. Simultaneously, the opening of the mPTP can be directly or indirectly inhibited through mitochondrial protective pathways such as SIRT3, which curtails the frequency of mPTP opening. The surplus mitochondrial ROS (mtROS) provokes inflammation, disrupts mitochondrial function, and causes damage to mitochondrial DNA (mtDNA). Moreover, mtROS can modulate the production of the pro-inflammatory cytokines IL-1β and IL-18 by activating the p65 subunit of the NF-κB signaling pathway and stimulating caspase-1 through the activation of NLRP3 inflammasomes. ROS-induced mtDNA damage manifests in two principal consequences: (i) structural impairments involving single and double strand breaks of DNA arising from direct ROS assault and (ii) the generation of 8-hydroxyguanine (8-OH-G).

3.1 Disruption of Mitochondrial Permeability Transition Pore Opening by Mitochondrial Oxidative Stress

The mPTP is a non-selective channel consisting of the voltage-dependent anion channel (VDAC) located in the outer mitochondrial membrane, the adenine nucleotide translocator (ANT), and cyclophilin D (Cyp-D) complex situated in the inner mitochondrial membrane [39]. An overabundance of mtROS can induce the opening of the mPTP. CypD functions as a modulator of mPTP and is capable of concealing the inhibitory binding site for phosphate on mPTP, thus heightening mPTP’s sensitivity to ROS and Ca2+ [40]. Nevertheless, the administration of TRO-19622, an inhibitor of VDAC, or Mito-TEMPO, a scavenger specifically targeting mROS, can effectively alleviate the aforementioned consequences [41]. Multiple studies have provided evidence that the generation of mtROS in response to oxidative stress can be modulated via mitochondrial protective pathways, such as SIRT3. These pathways exert direct or indirect inhibitory effects on the mPTP, thereby attenuating the frequency of mPTP opening events [42].

In cases where the range of adjustment cannot be controlled, a feedback loop of negative consequences may ensue, leading to a ‘vicious cycle’. Under pathophysiological circumstances, the reversible operation of ATP synthase occurs, wherein the hydrolysis of ATP takes place, creating an electrochemical gradient that traverses the mitochondrial inner membrane. These physiological phenomena may culminate in escalated ROS levels, consequently instigating the activation of mPTP [43]. Excessive accumulation of Ca2+ ions and heightened levels of ROS subsequent to SCI facilitate persistent activation of mPTP. The sustained activation of mPTP results in unimpeded diffusion of molecules with a mass greater than 1.5 kilodaltons (kDa) within the mitochondria, leading to a reduction in mitochondrial membrane potential [37]. The recurrent and persistent activation of mPTP obstructs electron transport and exacerbates the impairment of the mitochondrial electron transport chain. This leads to mtROS and Ca2+ overload, consequently intensifying the opening of mPTP [44]. Concomitantly, subsequent to mPTP activation, the persistent elevation of mtROS can precipitate the upregulation of pro-apoptotic pathways and elicit the translocation of pro-apoptotic factors (such as caspase3) to the mitochondria, thereby potentiating mPTP aperture [45]. Pyruvate carboxylase (PYC) exerts a stabilizing effect on mPTP by suppressing oxidative stress, which in turn ameliorates motor deficits following SCI and curtails neuronal apoptosis [46]. The aforementioned investigations have demonstrated an inextricable link between mPTP activation and mitochondrial oxidative stress. The initiation of mPTP will perpetuate oxidative stress, engendering a deleterious cycle that culminates in heightened cellular impairment.

3.2 Impact of Mitochondrial Oxidative Stress on Mitophagy

Autophagy is the process of combining damaged organelles or denatured proteins with lysosomes for self-degradation [47]. Mitophagy refers to a selective form of autophagy that functions to maintain mitochondrial homeostasis by selectively eliminating damaged or surplus mitochondria [48]. During this series of events, PTEN-induced putative kinase 1 (PINK1) perceives the depolarization of the mitochondrial membrane, leading to its accumulation on the outer membrane and the subsequent activation of Parkin. Upon activation, Parkin triggers the autophagic degradation process to achieve mitophagy. Additionally, activated Parkin can ubiquitinate other receptors located on the outer mitochondrial membrane, including NIP3-like protein X (NIX), FUNDC1, among others. These receptors can directly interact with microtubule-associated protein 1 light chain 3 (LC3), a molecule capable of sensitively detecting intracellular and extracellular signal changes and inducing the aggregation of autophagosomes [21].

The overproduction of mtROS causes a reduction in mitochondrial membrane potential (MMP). This, in turn, activates PINK1 located on the damaged outer mitochondrial membrane (OMM), engendering the initiation of ubiquitination and phosphorylation of OMM proteins, consequently promoting mitophagy [49]. Furthermore, heightened levels of ROS can trigger the initiation of mitophagy in vascular endothelial cells [50]. Mitophagy has the potential to reduce levels of ROS and may be involved in the amelioration of SCI [51]. Nonetheless, excessive mitochondrial oxidative stress can lead to an uncontrolled burst of ROS, ultimately disrupting the proper regulation of mitophagy, resulting in either an excess or deficiency of this process. This dysregulation of mitophagy is intricately associated with the onset and progression of certain pathologies. Following SCI, NIX can trigger an excessive level of mitophagy in neurons, subsequently promoting mitochondrial degeneration and neuronal cell death [52]. Research has demonstrated that betulinic acid (BA) can enhance autophagy in the context of SCI by activating the AMPK-mTOR-TFEB signaling pathway, thus promoting the induction of mitophagy and mitigating the accumulation of ROS [53]. Salidroside (Sal) has been shown to ameliorate mitochondrial dysfunction and structural abnormalities by mitigating the production of ROS. Moreover, Sal exerts its beneficial effects by augmenting the PTEN-induced PINK1-Parkin signaling pathway, which facilitates the initiation of mitophagy and enhances the clearance of impaired mitochondria [51]. These results indicate that mitochondrial oxidative stress may lead to abnormal mitophagy after SCI.

3.3 Induction of Inflammation by Mitochondrial Oxidative Stress

The generation of mtROS is intricately linked to the activation of pro-inflammatory cytokines [54]. The generation of ROS originating from the nicotinamide adenine dinucleotide phosphate (NADPH) oxidase system, as well as the production of O2- from the ETC within mitochondria, collectively promote the amplification of proinflammatory cytokine production [55]. mtROS exert a significant regulatory influence on the activation of the NF-κB signaling pathway and the modulation of inflammatory vesicular signaling, consequently impacting the inflammatory response. Notably, mitochondria have been reported to play a role in the activation of NLRP3 inflammasomes through mtROS, mtDNA, and cardiolipin, although the precise contribution of mitochondria in this process remains controversial. Simultaneously, NLRP3 inflammasomes induce caspase-1-dependent damage to mitochondria, resulting in escalated mtROS production, dissipation of mitochondrial membrane potential, and compromised integrity of both the inner and outer mitochondrial membranes [56]. These findings unveil the significant involvement of mitochondrial oxidative stress in the initiation of inflammation. Activation of NLRP3 inflammasomes prompts caspase-1 activation, orchestrating the production of pro-inflammatory cytokines IL-1β and IL-18. These cytokines, upon binding to specific intracellular receptors, potentiate the inflammatory response, ultimately culminating in organ dysfunction [57, 58]. The restitution of mitochondrial structural integrity and functional capacity impedes ROS generation, consequently mitigating the inflammatory response. Concurrently, mtROS triggers the activation of NLRP3 inflammasomes, and there is also a potential involvement of mtDNA oxidative damage in promoting a pro-inflammatory milieu [59]. Several studies have established a correlation between ETC activity and the activation of NLRP3 inflammasomes mediated by ROS. During cerebral ischemic injury, the elevated levels of succinic acid increased during ischemia, undergo oxidation. This oxidation, facilitated by reverse electron transport, stimulates ROS production within complex I, resulting in an initial surge of O2- that contributes to ischemia-reperfusion injury. Consequently, the NLRP3-dependent secretion of IL-1β is heightened, exacerbating the inflammatory response [58]. Simultaneously, it has been observed that lipopolysaccharide (LPS)-activated microglia exhibit an overproduction of chemokines, cytokines, and ROS. Within this pro-inflammatory state, the suppression of mtROS induced by LPS assumes a regulatory role in governing the synthesis of pro-inflammatory mediators by impeding the activation of MAPK and NF-κB signaling pathways [60].

While no direct evidence of the impact of mitochondrial oxidative stress on SCI inflammation exists, research has demonstrated that the knockout of mitochondrial membrane protein phosphoglycerate mutase 5 (PGAM5) can furnish neuroprotection by engaging in the oxidative stress and inflammation cascade in microglia. Moreover, PGAM5 knockout can proficiently govern mtROS and GSH levels and enhance mitochondrial dysfunction [61]. The transmission of miR-155 through exosomes is implicated in the initiation of EndoMT and the production of mtROS in bEnd.1 cells stimulated by M3-BMDMs. MtROS activates the NF-κB signaling cascade via targeting the downstream inhibitors of cytokine signaling 6 (SOCS6), thereby restraining SOCS6-mediated p65 ubiquitination and degradation [62]. Melatonin has the capability to restrain NLRP3 inflammation activation by inducing the Nrf3/ARE pathway. This lessens the ROS and malondialdehyde (MDA) expression and boosts the superoxide dismutase (SOD) and glutathione peroxidase (GSH-PX) activity, subsequently suppressing neuroinflammation, ameliorating mitochondrial dysfunction, and enhancing motor function recovery post-SCI [63]. This indirectly indicates that the correlation between mitochondrial oxidative damage and inflammation and requires further investigation.

3.4 MtDNA Damage Caused by Mitochondrial Oxidative Stress

mtDNA is a circular double-stranded molecule with a length of approximately 16.569, representing approximately 0.1%–1% of the total cellular DNA. The mtDNA coding region encompasses 37 genes, which consist of 22 transfer RNA (tRNA) genes, 2 ribosomal RNA (rRNA) genes, and 13 polypeptides [64]. The mtDNA is positioned in close proximity to the ETC. Due to its lack of nucleosomal protection and histone modification, mtDNA is vulnerable to damage, limited in its repair pathway, and prone to a high frequency of mutations, particularly under oxidative stress conditions. Thirteen peptides encoded by mtDNA play a crucial role in intracellular respiratory chain transmission and OXPHOS [65]. The mitochondrial transcription factor TFAM governs the modulation of mtDNA-encoded subunits in the ETC. Dysfunctional NRF1/2 signaling impedes the transcription of nuclear-encoded subunits of the respiratory chain complex and TFAM. The absence of TFAM packaging in the mtDNA prompts instability of the D-loop and impedes the transcription and replication of mtDNA. Consequently, diminished mtDNA transcription and replication accentuate the impairment of the respiratory chain [66]. One of the principal mechanisms by which ROS induces mtDNA damage is through oxidative modification of purine and pyrimidine bases, resulting in point mutations. ROS-mediated damage can lead to two main outcomes: (1) Structural damage, including single and double strand breaks of DNA, which are caused by direct ROS attack. The occurrence of DNA breakage following SCI was confirmed by comet assay. (2) Formation of 8-hydroxyguanine (8-OH-G). The primary products of mtDNA base damage include thymidinediol in pyrimidine nucleosides and 8-hydroxy-2deoxyguanosine (8-oxoG) in purines. The most extensively studied oxidative base damage is 8-oxodG, which is generated by ROS attack at the C8 position of guanine. 8-oxodG serves as a reliable marker for DNA oxidative damage [67]. Prior investigations have demonstrated that the levels of 8-oxodG increase following SCI, with significant increases in 8-oxodG observed at all time points from 3 hours post-injury (hpi) to 21 days post-injury (dpi) [68]. Nonetheless, 8-oxodG may exist in various forms of nucleic acids, including cytoplasmic RNA, nuclear DNA, and mitochondrial DNA. Further investigations are warranted to elucidate the precise source of 8-oxodG. During SCI, ROS can impact ETC function and OXPHOS via mtDNA-encoded peptides, which may lead to mtDNA damage [65]. These findings indicate that mtDNA is highly vulnerable to oxidative damage induced by ROS. In a rodent model of spinal cervical spondylosis, administration of riluzole mitigated oxidative DNA damage in the spinal cord post-decompression surgery and decreased the presence of 8-oxoG DNA. Moreover, in vitro investigations demonstrated that riluzole conserved mitochondrial functionality and decreased oxidative neuronal harm [69]. Currently, limited research has explored the mechanisms underlying mtDNA repair in SCI, with existing studies primarily focusing on evaluating alterations in gene expression subsequent to SCI. Further investigations are warranted to elucidate the association between mtDNA, mitochondrial oxidative stress, and their therapeutic interventions. As our scientific inquiry progresses, numerous gaps pertaining to the understanding of mtDNA damage and repair necessitate comprehensive investigation.

4. Treatment Strategies to Inhibit Mitochondrial Oxidative Stress

Mitochondrial oxidative stress triggers abnormalities in the ETC, leading to uncontrolled generation of mtROS, which subsequently triggers a cascade of pathophysiological changes such as opening of the mPTP, abnormal mitophagy, inflammation, and mtDNA damage. Therapeutic interventions targeting mitochondrial oxidative stress may represent a promising therapeutic approach. Despite its essential role in ATP production, the ETC also generates a considerable amount of ROS upon dysfunction. Following SCI, mitochondrial oxidative stress rapidly intensifies, creating a “vicious cycle”. Within the scope of this manuscript, we provide a detailed overview of recent advances in antioxidant therapy for mitochondria after SCI. Our focus is on two approaches: correcting ETC abnormalities under pathological conditions, and removing excessive mtROS, to facilitate a more comprehensive and in-depth understanding of SCI treatment Table 1 (Ref. [70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85]).

Table 1.Treatment strategies to inhibit mitochondrial oxidative stress.
Name Mechanism Dosage Animal strains Reference
NAC, NACA GSH↑, mtROS↓ NAC 100 mg/kg was injected intraperitoneally once a day for 28 days. NACA (150 or 300 mg/kg/day) was continuously administered 24 hours or 7 days after SCI. C57BL/6 mice; (Guo et al, [79] 2015; Patel et al., [80] 2014)
Female Sprague-Dawley rats
SOD/CAT (NPs) Restore ETC function At 6 hours after spinal cord injury, 30 mg/kg was injected into a caudal vein. Sprague-Dawley rats (Andrabi et al., [70] 2020)
mtROS↓
Mitochondrial Membrane Potential↑
ATP synthesis↑
5-HT1F receptor stimulant (LY344864) ATP Synβ, TFAM and NDUFB8↑ One hour after spinal cord injury, 2.0 mg/kg was injected intraperitoneally once a day for 21 days. Female C57BL/6J mice, male and female 5-HT1F receptor gene knockout (KO) mice (Simmons et al., [71] 2020)
Reduce mtDNA damage and improve mitochondrial dysfunction
Thiamine OGDHC↑ Within 15–20 hours after spinal cord injury, 200 mg/mL was injected intraperitoneally. Female Sprague-Dawley rats (Boyko et al., [72] 2021)
Repair TCA cycle
Improve ETC and OXPHOS
α5 (Including TCA cycle intermediates, EAAs, BCAA and cofactors thiamine and pyridoxine) Repair TCA cycle Oral administration was given 3 days after spinal cord injury and continued until the end of the experimental procedure. C57BL/6J mice (Dolci et al., [73] 2022)
Improve ETC and OXPHOS
Mitochondrial mass and respiration↑
oxidative stress↓
Ketogenic diet Activity of complexes I, II, III and IV↑ Ketogenic diet for 7 days after spinal cord injury Male Sprague-Dawley rats (Seira et al., [74] 2021)
mitochondrial oxidative damage↓
NS309 (SK/KCa channel activator) mitochondrial respiratory chain complex activity↑ After spinal cord ischemia-reperfusion injury, 2 mg/kg was injected intraperitoneally. Adult New Zealand white rabbits (Zhu et al., [75] 2019)
antioxidation
Correct mitochondrial dysfunction
Leptin Correct mitochondrial ETC abnormalities Subcutaneous injection of 1 mg/kg was performed 3 hours before MCAO. Male Sprague-Dawley rats (Hu et al., [76] 2019)
mitochondrial oxidative stress↓
NMES-RT Mitochondrial electron transport chain complex I↑ Twice a week for 12 weeks, 45–60 minutes per session Thirty-three individuals aged between 20 and 61 years with chronic (≥1 year post injury) SCI (Gorgey et al., [77] 2021)
PBM p-AMPK, PGC-1α, Nrf1, Sirt1 and TFAM↑ After spinal cord injury, the T10 spinal cord was exposed to PBM for 1 hour per day. Sprague-Dawley rats (Zhu et al., [78] 2022)
Restore mitochondrial respiratory chain complex activity
MitoQ Mitochondrial specific antioxidants MitoQ 5 mg/kg was intraperitoneally injected on day 0, 1 and 2 after spinal cord injury. C57BL/6J mice (Huang et al., [81] 2022)
mtROS↓
Polydatin Regulate mPTP, effectively remove mtROS, and reduce mitochondrial dysfunction. Spinal cord ischemia-reperfusion injury was induced by gastric injection of 30 mg/kg 2 days before the operation for 7 days. C57BL/6J mice (Zhan et al., [82] 2021)
XJB-5-131 mitochondrial-targeted ROS scavenger, inhibition of CL, reduces neuronal apoptosis 15 mg/kg was injected intraperitoneally 30 minutes after spinal cord injury. Female Sprague-Dawley rats (Liu et al., [83] 2022)
Zn mtROS↓ After spinal cord injury, ZnG (different concentrations) was injected intraperitoneally. C57BL/6J mice (Xu C et al., [84] 2023)
EPO Reduce mitochondrial damage and improve mitochondrial membrane potential in ferroptosis. After spinal cord injury, 1000 IU/kg and 5000 IU/kg were injected intraperitoneally once a week for 2 weeks. Female Sprague-Dawley rats (Kang et al., [85] 2023)

NAC, N-acetylcysteine; NACA, N-acetylcysteine amide; GSH, glutathione; mtROS, mitochondrial reactive oxygen species; SCI, Spinal cord injury; SOD, superoxide dismutase; CAT, catalase; ETC, electron transport chain; ATP, adenosine triphosphate; mtDNA, mitochondrial DNA; TCA, tricarboxylic acid; MCAO, middle cerebral artery occlusion; NMES-RT, neuromuscular electrical stimulation resistance training; PBM, photobiomodulation; mPTP, mitochondrial permeability transition pore; EPO, Erythropoietin; OXPHOS, oxidative phosphorylation; EAAs, essential amino acids; BCAA, branched-chain amino acids; CL, cardiolipin; ZnG, zinc gluconate. ↑ represents up regulation, and ↓ represents down regulation.

4.1 Correct Mitochondrial Oxidative Respiratory Chain Abnormalities
4.1.1 Drug Treatment

The administration of SOD/CAT (NPs) nanoparticles was observed to significantly ameliorate the excessive production of mtROS resulting from impaired ETC transmission components after SCI, leading to an increase in MMP, a reduction in Ca2+ content, and enhanced ATP synthesis [70]. LY344864 is a 5-hydroxytryptamine 1F (5-HT1F) receptor agonist, which has been demonstrated to attenuate the reduction in mtDNA content following SCI, reverse the downregulation of Nrf2, ATP Synβ, TFAM, and NDUFB8 protein expression after SCI, and improve the motor function of the lower limbs in mice. Thus, LY344864 holds promise as a potent 5-HT1F receptor agonist for developing SCI therapeutic interventions [71]. Furthermore, research has indicated that the decrease in the OGDHC level following SCI restricts the tricarboxylic acid (TCA) cycle, leading to reduced production of NADH for oxidation in the ETC. This ultimately impairs crucial mitochondrial functions, including oxidative phosphorylation. Interventions with thiamine, a pleiotropic modulator, may activate neuroprotective mechanisms, enhance motor function recovery following SCI, and potentially mitigate the downregulation of OGDHC expression post-SCI [72]. It is noteworthy that research has demonstrated the potential utilization of certain bioenergetic compounds, such as intermediates of the TCA cycle, the administration of essential amino acids (EAAs), branched-chain amino acids (BCAA) and coenzymes thiamine and pyridoxine, collectively termed as α5. Oral administration of α5 has been found to ameliorate the TCA cycle and OXPHOS in spinal cord tissue, augment mitochondrial biogenesis and respiration, and mitigate oxidative stress at the site of injury [73]. The ketogenic diet has been reported to enhance the function of mitochondrial respiratory complexes I, II, III, and IV post-SCI while mitigating acute SCI-induced mitochondrial oxidative damage and dysfunction [74]. In the model of spinal cord ischemia-reperfusion injury (SCI/R), administration of NS309, a pharmacological activator of small conductance calcium-activated K+ (SK/KCa) channels, has been found to preserve the activity of mitochondrial respiratory complexes, thus preventing SCI/R by exhibiting antioxidant activity and inhibiting mitochondrial dysfunction [75]. In the rat model of cerebral ischemic injury, leptin has been shown to modulate the ETC through the STAT3 pathway, resulting in decreased mitochondrial damage and protection of brain tissue against mitochondrial oxidative stress during cerebral ischemia [76]. Whether this approach can be employed for the treatment of SCI requires further investigation. By modulating the ETC, the drugs exert regulatory control over cellular OXPHOS and disrupt the process of oxidative stress, thereby effectively counteracting oxidative stress and preserving compromised spinal cord tissue. The respiratory chain complex, integral to biological oxidation and OXPHOS, exhibits abundant sites for inhibition and activation. Consequently, distinct drugs or specific drug components exert varying effects on the mitochondrial respiratory chain complex, providing avenues for drug development in the context of spinal cord injury. These strategies include targeted drug delivery to specific sites within the ETC and its complex, as well as augmenting the proportion of specific drug components, thereby enhancing the drug’s efficacy in disease treatment.

4.1.2 Non-Drug Therapy

Neuromuscular electrical stimulation resistance training (NMES-RT) has been shown to enhance peak oxygen uptake (Vo2peak) of skeletal muscle in individuals with chronic SCI, possibly attributed to alterations in complex I of the ETC [77]. In rat models of SCI, photobiomodulation (PBM) has been found to stimulate the AMPK signaling pathway, leading to the upregulation of p-AMPK, TFAM, Nrf1, Sirt1 and PGC-1α. This activation of PBM can re-establish the functional activity of mitochondrial respiratory chain complexes, which subsequently augment the production of ATP [78].

4.2 Clear Excess mtROS
4.2.1 Drug Treatment

Glutathione reductase (GR) plays a vital role in the reduction of oxidized glutathione (GSSG) to reduce GSH, facilitating the scavenging of ROS and safeguarding mitochondrial membrane integrity. N-acetylcysteine (NAC), a GSH precursor, exhibits inhibitory effects on oxidative stress injuries induced by mitochondrial dysfunction and demonstrates a certain capacity for improving mitochondrial dysfunction following SCI [79]. However, NAC’s biofilm permeability is limited. Studies have demonstrated that N-acetylcysteine amide (NACA) significantly elevates GSH levels within the body and reduces the production of mtROS, thereby mitigating mitochondrial oxidative stress subsequent to SCI [80]. In recent years, many mitochondrial antioxidants have been known. MitoQ can effectively remove mtROS, demonstrating a certain therapeutic effect of mitochondrial oxidative stress. Studies have reported that MitoQ enhances spinal cord angiogenesis by restoring impaired mitochondrial function and facilitating the recovery of mitochondrial structure and function following SCI [81]. Thus, MitoQ is considered a promising therapeutic intervention for mitigating mitochondrial oxidative stress subsequent to SCI. In the spinal cord ischemia-reperfusion model, Polydatin (PD) has been found to regulate MMP and mPTP opening by activating the Nrf2/ARE pathway, thereby effectively scavenging mtROS, restoring ATP synthesis and mitigating neuronal damage due to mitochondrial dysfunction [82]. The therapeutic potential of PD in SCI warrants further investigation. A novel mitochondrial-targeted ROS scavenger, XJB-5-131, has been developed, and studies have demonstrated its efficacy in inhibiting cardiolipin (CL) alterations in rats after SCI, reducing tissue damage, neuronal apoptosis, and improving motor function recovery following SCI [83]. Zinc (Zn) has been shown to promote autophagy via SIRT3, thus alleviating inflammation and mtROS production in damaged spinal cord and neurons [84]. Subsequent experiments demonstrated that Zn could also inhibit mtROS production via the PI3K/Akt signaling pathway [86]. Furthermore, dynasore has emerged as a promising combined ROS blocker for in vivo research and treatment of abnormal accumulation of mitochondrial ROS [87]. In recent years, researchers have been actively enhancing pharmaceutical agents through modulating administration modalities, augmenting the utilization of biologics, and continually refining delivery methodologies to establish a robust groundwork for clinical intervention. While the management of mtROS is not novel, it remains clear that the inhibition of ROS generation persists as a prominent contributor to mitochondrial oxidative stress.

4.2.2 Non-Drug Therapy

Research has indicated that the inhibition of the NF-κB pathway via down-regulation of exosome miR-155 in polarized M1 macrophages can enhance the proliferation of vascular endothelial cells and reduce the production of mtROS following traumatic SCI [62]. Currently, the research landscape concerning miRNA therapy targeting mitochondrial oxidative stress is relatively sparse, yet it remains a compelling and pertinent subject warranting our focused consideration.

4.3 Others

In addition to the aforementioned treatments, various drugs have been found to alleviate mitochondrial oxidative stress and have neuroprotective effects on SCI. These drugs include maltol, metformin, receptor-interacting protein (RIP) inhibitors, ligustrazine, ebselen, among others [88, 89, 90, 91, 92, 93]. Erythropoietin (EPO) has been found to contribute to the recovery of reduced mitochondrial membrane potential in ferroptosis and reduce mitochondrial damage after SCI [85]. Ferrostatin-1 has been shown to inhibit mitochondrial lipid peroxidation, reduce ROS and MDA levels, and promote the expression of GSH and GPX4 in the RSL-93-induced ferroptosis model of oligodendrocytes [94]. Most antioxidants have difficulty passing through the blood-brain barrier, and the selective permeability of the mitochondrial membrane limits the targeting efficiency of many drugs on mitochondria, leading to diminished clinical effects. Thus, the efficient delivery of drugs to neuronal mitochondria while avoiding toxicity remains a challenge in SCI treatment. Moreover, apart from the aforementioned therapeutic approaches, the modulation of molecules implicated in mitochondrial oxidative stress, such as uncoupling proteins, deacetylases, and others, either individually or in combination, holds promise for reinstating neuronal mitochondrial functionality in SCI, thus affording protection to the spinal cord tissue.

5. Conclusions

Through extensive research, it has been elucidated that mitochondrial oxidative stress exerts disruptive effects on mPTP opening, mtDNA damage, mitophagy impairment, inflammatory induction, and other detrimental processes. This implies the pivotal role of mitochondrial oxidative stress in the cascade of secondary injury subsequent to SCI. Hence, the inhibition of mitochondrial oxidative stress remains a compelling avenue for exploration in the treatment of SCI. Multiple studies have substantiated this proposition, highlighting strategies such as targeted mitigation of excess mtROS through interventions at the ETC and its complexes, utilization of biological materials, and augmentation of drug efficacy by optimizing dosages. Furthermore, non-pharmaceutical approaches that target ETC and mtROS demonstrate promising potential for future spinal cord injury recovery. Intriguingly, the intricate distribution of molecular proteins within mitochondria, closely intertwined with mitochondrial function, holds considerable significance in maintaining mitochondrial homeostasis and functionality. Nonetheless, research on gene therapy and combination therapy for mitigating mitochondrial oxidative stress in the context of treatment remains limited. Given the multifaceted nature of spinal cord injury, focusing solely on the treatment of a specific secondary loss is unlikely to yield comprehensive efficacy. Hence, the exploration of combination therapy and the quest for single-drug interventions capable of addressing multiple secondary injuries represent imperative and worthwhile research directions.

Author Contributions

ZH was resposible for conceptualizing, writing and revising this article; CZ drew the illustrations; JXL developed the search strategy and literature search; FFZ is responsible for combing references; XYQ and FG designed and critically revised the manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

We are grateful to three anonymous reviewers for their critical comments on the manuscript.

Funding

This research received no external funding.

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Zhao C, Rao JS, Duan H, Hao P, Shang J, Fan Y, et al. Chronic spinal cord injury repair by NT3-chitosan only occurs after clearance of the lesion scar. Signal Transduction and Targeted Therapy. 2022; 7: 184.
[2]
Matson KJE, Russ DE, Kathe C, Hua I, Maric D, Ding Y, et al. Single cell atlas of spinal cord injury in mice reveals a pro-regenerative signature in spinocerebellar neurons. Nature Communications. 2022; 13: 5628.
[3]
Gong W, Zhang T, Che M, Wang Y, He C, Liu L, et al. Recent advances in nanomaterials for the treatment of spinal cord injury. Materials Today Bio. 2022; 18: 100524.
[4]
Karsy M, Hawryluk G. Modern Medical Management of Spinal Cord Injury. Current Neurology and Neuroscience Reports. 2019; 19: 65.
[5]
Katoh H, Yokota K, Fehlings MG. Regeneration of Spinal Cord Connectivity Through Stem Cell Transplantation and Biomaterial Scaffolds. Frontiers in Cellular Neuroscience. 2019; 13: 248.
[6]
Liu K, Dong X, Wang Y, Wu X, Dai H. Dopamine-modified chitosan hydrogel for spinal cord injury. Carbohydrate Polymers. 2022; 298: 120047.
[7]
Tian T, Zhang S, Yang M. Recent progress and challenges in the treatment of spinal cord injury. Protein & Cell. 2023; pwad003.
[8]
Walsh CM, Gull K, Dooley D. Motor rehabilitation as a therapeutic tool for spinal cord injury: New perspectives in immunomodulation. Cytokine & Growth Factor Reviews. 2023; 69: 80–89.
[9]
Sámano C, Kaur J, Nistri A. A study of methylprednisolone neuroprotection against acute injury to the rat spinal cord in vitro. Neuroscience. 2016; 315: 136–149.
[10]
Onder C, Onder C, Akesen S, Yumusak E, Akesen B. Riluzole is Effective on Spinal Decompression for Treating Acute Spinal Injury When Compared With Methylprednisolone and the Combination of Two Drugs: In Vivo Rat Model. Global Spine Journal. 2023; 21925682231159068.
[11]
Bracken MB, Shepard MJ, Holford TR, Leo-Summers L, Aldrich EF, Fazl M, et al. Administration of methylprednisolone for 24 or 48 hours or tirilazad mesylate for 48 hours in the treatment of acute spinal cord injury. Results of the Third National Acute Spinal Cord Injury Randomized Controlled Trial. National Acute Spinal Cord Injury Study. Journal of American Medical Association. 1997; 277: 1597–1604.
[12]
Evaniew N, Noonan VK, Fallah N, Kwon BK, Rivers CS, Ahn H, et al. Methylprednisolone for the Treatment of Patients with Acute Spinal Cord Injuries: A Propensity Score-Matched Cohort Study from a Canadian Multi-Center Spinal Cord Injury Registry. Journal of Neurotrauma. 2015; 32: 1674–1683.
[13]
Daverey A, Agrawal SK. Neuroprotective effects of Riluzole and Curcumin in human astrocytes and spinal cord white matter hypoxia. Neuroscience Letters. 2020; 738: 135351.
[14]
Wu Q, Zhang W, Yuan S, Zhang Y, Zhang W, Zhang Y, et al. A Single Administration of Riluzole Applied Acutely After Spinal Cord Injury Attenuates Pro-inflammatory Activity and Improves Long-Term Functional Recovery in Rats. Journal of Molecular Neuroscience: MN. 2022; 72: 730–740.
[15]
Li HX, Cui J, Fan JS, Tong JZ. An observation of the clinical efficacy of combining Riluzole with mannitol and hyperbaric oxygen in treating acute spinal cord injury. Pakistan Journal of Medical Sciences. 2021; 37: 320–324.
[16]
Grossman RG, Fehlings MG, Frankowski RF, Burau KD, Chow DSL, Tator C, et al. A prospective, multicenter, phase I matched-comparison group trial of safety, pharmacokinetics, and preliminary efficacy of riluzole in patients with traumatic spinal cord injury. Journal of Neurotrauma. 2014; 31: 239–255.
[17]
Yune TY, Lee JY, Jung GY, Kim SJ, Jiang MH, Kim YC, et al. Minocycline alleviates death of oligodendrocytes by inhibiting pro-nerve growth factor production in microglia after spinal cord injury. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2007; 27: 7751–7761.
[18]
Bin S, Zhou N, Pan J, Pan F, Wu XF, Zhou ZH. Nano-carrier mediated co-delivery of methyl prednisolone and minocycline for improved post-traumatic spinal cord injury conditions in rats. Drug Development and Industrial Pharmacy. 2017; 43: 1033–1041.
[19]
Casha S, Zygun D, McGowan MD, Bains I, Yong VW, Hurlbert RJ. Results of a phase II placebo-controlled randomized trial of minocycline in acute spinal cord injury. Brain: a Journal of Neurology. 2012; 135: 1224–1236.
[20]
Luo W, Wang Y, Lin F, Liu Y, Gu R, Liu W, et al. Selenium-Doped Carbon Quantum Dots Efficiently Ameliorate Secondary Spinal Cord Injury via Scavenging Reactive Oxygen Species. International Journal of Nanomedicine. 2020; 15: 10113–10125.
[21]
Rabchevsky AG, Michael FM, Patel SP. Mitochondria focused neurotherapeutics for spinal cord injury. Experimental Neurology. 2020; 330: 113332.
[22]
Jia ZQ, Li G, Zhang ZY, Li HT, Wang JQ, Fan ZK, et al. Time representation of mitochondrial morphology and function after acute spinal cord injury. Neural Regeneration Research. 2016; 11: 137–143.
[23]
Nukolova NV, Aleksashkin AD, Abakumova TO, Morozova AY, Gubskiy IL, Kirzhanova ЕА, et al. Multilayer polyion complex nanoformulations of superoxide dismutase 1 for acute spinal cord injury. Journal of Controlled Release. 2018; 270: 226–236.
[24]
Zhang B, Bailey WM, McVicar AL, Gensel JC. Age increases reactive oxygen species production in macrophages and potentiates oxidative damage after spinal cord injury. Neurobiology of Aging. 2016; 47: 157–167.
[25]
Zhao RZ, Jiang S, Zhang L, Yu ZB. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). International Journal of Molecular Medicine. 2019; 44: 3–15.
[26]
Lewén A, Matz P, Chan PH. Free radical pathways in CNS injury. Journal of Neurotrauma. 2000; 17: 871–890.
[27]
Murphy MP. How mitochondria produce reactive oxygen species. The Biochemical Journal. 2009; 417: 1–13.
[28]
Nolfi-Donegan D, Braganza A, Shiva S. Mitochondrial electron transport chain: Oxidative phosphorylation, oxidant production, and methods of measurement. Redox Biology. 2020; 37: 101674.
[29]
Sullivan PG, Krishnamurthy S, Patel SP, Pandya JD, Rabchevsky AG. Temporal characterization of mitochondrial bioenergetics after spinal cord injury. Journal of Neurotrauma. 2007; 24: 991–999.
[30]
Azbill RD, Mu X, Bruce-Keller AJ, Mattson MP, Springer JE. Impaired mitochondrial function, oxidative stress and altered antioxidant enzyme activities following traumatic spinal cord injury. Brain Research. 1997; 765: 283–290.
[31]
Scholpa NE, Schnellmann RG. Mitochondrial-Based Therapeutics for the Treatment of Spinal Cord Injury: Mitochondrial Biogenesis as a Potential Pharmacological Target. The Journal of Pharmacology and Experimental Therapeutics. 2017; 363: 303–313.
[32]
Sullivan PG, Rabchevsky AG, Keller JN, Lovell M, Sodhi A, Hart RP, et al. Intrinsic differences in brain and spinal cord mitochondria: Implication for therapeutic interventions. The Journal of Comparative Neurology. 2004; 474: 524–534.
[33]
Mazat JP, Devin A, Ransac S. Modelling mitochondrial ROS production by the respiratory chain. Cellular and Molecular Life Sciences: CMLS. 2020; 77: 455–465.
[34]
Hussain ASM, Renno WM, Sadek HL, Kayali NM, Al-Salem A, Rao MS, et al. Monoamine oxidase-B inhibitor protects degenerating spinal neurons, enhances nerve regeneration and functional recovery in sciatic nerve crush injury model. Neuropharmacology. 2018; 128: 231–243.
[35]
Vigil TM, Frieler RA, Kilpatrick KL, Wang MM, Mortensen RM. Aconitate decarboxylase 1 suppresses cerebral ischemia-reperfusion injury in mice. Experimental Neurology. 2022; 347: 113902.
[36]
Traube FR, Özdemir D, Sahin H, Scheel C, Glück AF, Geserich AS, et al. Redirected nuclear glutamate dehydrogenase supplies Tet3 with α-ketoglutarate in neurons. Nature Communications. 2021; 12: 4100.
[37]
McEwen ML, Sullivan PG, Rabchevsky AG, Springer JE. Targeting mitochondrial function for the treatment of acute spinal cord injury. Neurotherapeutics: the Journal of the American Society for Experimental NeuroTherapeutics. 2011; 8: 168–179.
[38]
Yang N, Guan QW, Chen FH, Xia QX, Yin XX, Zhou HH, et al. Antioxidants Targeting Mitochondrial Oxidative Stress: Promising Neuroprotectants for Epilepsy. Oxidative Medicine and Cellular Longevity. 2020; 2020: 6687185.
[39]
Boyenle ID, Oyedele AK, Ogunlana AT, Adeyemo AF, Oyelere FS, Akinola OB, et al. Targeting the mitochondrial permeability transition pore for drug discovery: Challenges and opportunities. Mitochondrion. 2022; 63: 57–71.
[40]
Du H, Yan SS. Mitochondrial permeability transition pore in Alzheimer’s disease: cyclophilin D and amyloid beta. Biochimica et Biophysica Acta. 2010; 1802: 198–204.
[41]
Qu J, Chen W, Hu R, Feng H. The Injury and Therapy of Reactive Oxygen Species in Intracerebral Hemorrhage Looking at Mitochondria. Oxidative Medicine and Cellular Longevity. 2016; 2016: 2592935.
[42]
Yang Y, Tian Y, Guo X, Li S, Wang W, Shi J. Ischemia Injury induces mPTP opening by reducing Sirt3. Neuroscience. 2021; 468: 68–74.
[43]
Mnatsakanyan N, Jonas EA. The new role of F_1F_o ATP synthase in mitochondria-mediated neurodegeneration and neuroprotection. Experimental Neurology. 2020; 332: 113400.
[44]
Kent AC, El Baradie KBY, Hamrick MW. Targeting the Mitochondrial Permeability Transition Pore to Prevent Age-Associated Cell Damage and Neurodegeneration. Oxidative Medicine and Cellular Longevity. 2021; 2021: 6626484.
[45]
Ma Z, Xin Z, Di W, Yan X, Li X, Reiter RJ, et al. Melatonin and mitochondrial function during ischemia/reperfusion injury. Cellular and Molecular Life Sciences. 2017; 74: 3989–3998.
[46]
Tao L, Liu X, Da W, Tao Z, Zhu Y. Pycnogenol achieves neuroprotective effects in rats with spinal cord injury by stabilizing the mitochondrial membrane potential. Neurological Research. 2020; 42: 597–604.
[47]
Gorgey AS, Witt O, O’Brien L, Cardozo C, Chen Q, Lesnefsky EJ, et al. Mitochondrial health and muscle plasticity after spinal cord injury. European Journal of Applied Physiology. 2019; 119: 315–331.
[48]
Li Q, Gao S, Kang Z, Zhang M, Zhao X, Zhai Y, et al. Rapamycin Enhances Mitophagy and Attenuates Apoptosis After Spinal Ischemia-Reperfusion Injury. Frontiers in Neuroscience. 2018; 12: 865.
[49]
Lazarou M, Sliter DA, Kane LA, Sarraf SA, Wang C, Burman JL, et al. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature. 2015; 524: 309–314.
[50]
Fan P, Xie XH, Chen CH, Peng X, Zhang P, Yang C, et al. Molecular Regulation Mechanisms and Interactions Between Reactive Oxygen Species and Mitophagy. DNA and Cell Biology. 2019; 38: 10–22.
[51]
Gu C, Li L, Huang Y, Qian D, Liu W, Zhang C, et al. Salidroside Ameliorates Mitochondria-Dependent Neuronal Apoptosis after Spinal Cord Ischemia-Reperfusion Injury Partially through Inhibiting Oxidative Stress and Promoting Mitophagy. Oxidative Medicine and Cellular Longevity. 2020; 2020: 3549704.
[52]
Nie P, Wang H, Yu D, Wu H, Ni B, Kong J, et al. NIX Mediates Mitophagy in Spinal Cord Injury in Rats by Interacting with LC3. Cellular and Molecular Neurobiology. 2022; 42: 1983–1994.
[53]
Wu C, Chen H, Zhuang R, Zhang H, Wang Y, Hu X, et al. Betulinic acid inhibits pyroptosis in spinal cord injury by augmenting autophagy via the AMPK-mTOR-TFEB signaling pathway. International Journal of Biological Sciences. 2021; 17: 1138–1152.
[54]
Naik E, Dixit VM. Mitochondrial reactive oxygen species drive proinflammatory cytokine production. The Journal of Experimental Medicine. 2011; 208: 417–420.
[55]
McElroy PB, Liang LP, Day BJ, Patel M. Scavenging reactive oxygen species inhibits status epilepticus-induced neuroinflammation. Experimental Neurology. 2017; 298: 13–22.
[56]
Bader V, Winklhofer KF. Mitochondria at the interface between neurodegeneration and neuroinflammation. Seminars in Cell & Developmental Biology. 2020; 99: 163–171.
[57]
Han X, Xu T, Fang Q, Zhang H, Yue L, Hu G, et al. Quercetin hinders microglial activation to alleviate neurotoxicity via the interplay between NLRP3 inflammasome and mitophagy. Redox Biology. 2021; 44: 102010.
[58]
Vezzani B, Carinci M, Patergnani S, Pasquin MP, Guarino A, Aziz N, et al. The Dichotomous Role of Inflammation in the CNS: A Mitochondrial Point of View. Biomolecules. 2020; 10: 1437.
[59]
Wu X, Gong L, Xie L, Gu W, Wang X, Liu Z, et al. NLRP3 Deficiency Protects Against Intermittent Hypoxia-Induced Neuroinflammation and Mitochondrial ROS by Promoting the PINK1-Parkin Pathway of Mitophagy in a Murine Model of Sleep Apnea. Frontiers in Immunology. 2021; 12: 628168.
[60]
Nair S, Sobotka KS, Joshi P, Gressens P, Fleiss B, Thornton C, et al. Lipopolysaccharide-induced alteration of mitochondrial morphology induces a metabolic shift in microglia modulating the inflammatory response in vitro and in vivo. Glia. 2019; 67: 1047–1061.
[61]
Dai C, Qu B, Peng B, Liu B, Li Y, Niu C, et al. Phosphoglycerate mutase 5 facilitates mitochondrial dysfunction and neuroinflammation in spinal tissues after spinal cord injury. International Immunopharmacology. 2023; 116: 109773.
[62]
Ge X, Tang P, Rong Y, Jiang D, Lu X, Ji C, et al. Exosomal miR-155 from M1-polarized macrophages promotes EndoMT and impairs mitochondrial function via activating NF-κB signaling pathway in vascular endothelial cells after traumatic spinal cord injury. Redox Biology. 2021; 41: 101932.
[63]
Wang H, Wang H, Huang H, Qu Z, Ma D, Dang X, et al. Melatonin Attenuates Spinal Cord Injury in Mice by Activating the Nrf2/ARE Signaling Pathway to Inhibit the NLRP3 Inflammasome. Cells. 2022; 11: 2809.
[64]
Xu S, Zhang X, Liu C, Liu Q, Chai H, Luo Y, et al. Role of Mitochondria in Neurodegenerative Diseases: From an Epigenetic Perspective. Frontiers in Cell and Developmental Biology. 2021; 9: 688789.
[65]
Liu C, Liu Y, Ma B, Zhou M, Zhao X, Fu X, et al. Mitochondrial regulatory mechanisms in spinal cord injury: A narrative review. Medicine. 2022; 101: e31930.
[66]
Grünewald A, Rygiel KA, Hepplewhite PD, Morris CM, Picard M, Turnbull DM. Mitochondrial DNA Depletion in Respiratory Chain-Deficient Parkinson Disease Neurons. Annals of Neurology. 2016; 79: 366–378.
[67]
Zhang Q, Wang J, Gu Z, Zhang Q, Zheng H. Effect of lycopene on the blood-spinal cord barrier after spinal cord injury in mice. Bioscience Trends. 2016; 10: 288–293.
[68]
Scheijen EEM, Hendrix S, Wilson DM 3rd. Oxidative DNA Damage in the Pathophysiology of Spinal Cord Injury: Seems Obvious, but Where Is the Evidence? Antioxidants (Basel, Switzerland). 2022; 11: 1728.
[69]
Karadimas SK, Laliberte AM, Tetreault L, Chung YS, Arnold P, Foltz WD, et al. Riluzole blocks perioperative ischemia-reperfusion injury and enhances postdecompression outcomes in cervical spondylotic myelopathy. Science Translational Medicine. 2015; 7: 316ra194.
[70]
Andrabi SS, Yang J, Gao Y, Kuang Y, Labhasetwar V. Nanoparticles with antioxidant enzymes protect injured spinal cord from neuronal cell apoptosis by attenuating mitochondrial dysfunction. Journal of Controlled Release: Official Journal of the Controlled Release Society. 2020; 317: 300–311.
[71]
Simmons EC, Scholpa NE, Cleveland KH, Schnellmann RG. 5-hydroxytryptamine 1F Receptor Agonist Induces Mitochondrial Biogenesis and Promotes Recovery from Spinal Cord Injury. The Journal of Pharmacology and Experimental Therapeutics. 2020; 372: 216–223.
[72]
Boyko A, Tsepkova P, Aleshin V, Artiukhov A, Mkrtchyan G, Ksenofontov A, et al. Severe Spinal Cord Injury in Rats Induces Chronic Changes in the Spinal Cord and Cerebral Cortex Metabolism, Adjusted by Thiamine That Improves Locomotor Performance. Frontiers in Molecular Neuroscience. 2021; 14: 620593.
[73]
Dolci S, Mannino L, Bottani E, Campanelli A, Di Chio M, Zorzin S, et al. Therapeutic induction of energy metabolism reduces neural tissue damage and increases microglia activation in severe spinal cord injury. Pharmacological Research. 2022; 178: 106149.
[74]
Seira O, Kolehmainen K, Liu J, Streijger F, Haegert A, Lebihan S, et al. Ketogenesis controls mitochondrial gene expression and rescues mitochondrial bioenergetics after cervical spinal cord injury in rats. Scientific Reports. 2021; 11: 16359.
[75]
Zhu J, Yang LK, Chen WL, Lin W, Wang YH, Chen T. Activation of SK/KCa Channel Attenuates Spinal Cord Ischemia-Reperfusion Injury via Anti-oxidative Activity and Inhibition of Mitochondrial Dysfunction in Rabbits. Frontiers in Pharmacology. 2019; 10: 325.
[76]
Hu S, Cheng D, Peng D, Tan J, Huang Y, Chen C. Leptin attenuates cerebral ischemic injury in rats by modulating the mitochondrial electron transport chain via the mitochondrial STAT3 pathway. Brain and Behavior. 2019; 9: e01200.
[77]
Gorgey AS, Lai RE, Khalil RE, Rivers J, Cardozo C, Chen Q, et al. Neuromuscular electrical stimulation resistance training enhances oxygen uptake and ventilatory efficiency independent of mitochondrial complexes after spinal cord injury: a randomized clinical trial. Journal of Applied Physiology (Bethesda, Md.: 1985). 2021; 131: 265–276.
[78]
Zhu Z, Wang X, Song Z, Zuo X, Ma Y, Zhang Z, et al. Photobiomodulation promotes repair following spinal cord injury by restoring neuronal mitochondrial bioenergetics via AMPK/PGC-1α/TFAM pathway. Frontiers in Pharmacology. 2022; 13: 991421.
[79]
Guo J, Li Y, Chen Z, He Z, Zhang B, Li Y, et al. N-acetylcysteine treatment following spinal cord trauma reduces neural tissue damage and improves locomotor function in mice. Molecular Medicine Reports. 2015; 12: 37–44.
[80]
Patel SP, Sullivan PG, Pandya JD, Goldstein GA, VanRooyen JL, Yonutas HM, et al. N-acetylcysteine amide preserves mitochondrial bioenergetics and improves functional recovery following spinal trauma. Experimental Neurology. 2014; 257: 95–105.
[81]
Huang T, Shen J, Bao B, Hu W, Sun Y, Zhu T, et al. Mitochondrial-targeting antioxidant MitoQ modulates angiogenesis and promotes functional recovery after spinal cord injury. Brain Research. 2022; 1786: 147902.
[82]
Zhan J, Li X, Luo D, Yan W, Hou Y, Hou Y, et al. Polydatin Attenuates OGD/R-Induced Neuronal Injury and Spinal Cord Ischemia/Reperfusion Injury by Protecting Mitochondrial Function via Nrf2/ARE Signaling Pathway. Oxidative Medicine and Cellular Longevity. 2021; 2021: 6687212.
[83]
Liu NK, Deng LX, Wang M, Lu QB, Wang C, Wu X, et al. Restoring mitochondrial cardiolipin homeostasis reduces cell death and promotes recovery after spinal cord injury. Cell Death & Disease. 2022; 13: 1058.
[84]
Xu C, Zhou Z, Zhao H, Lin S, Zhang P, Tian H, et al. Zinc Promotes Spinal Cord Injury Recovery by Blocking the Activation of NLRP3 Inflammasome Through SIRT3-Mediated Autophagy. Neurochemical Research. 2023; 48: 435–446.
[85]
Kang Y, Zhu R, Li S, Qin KP, Tang H, Shan WS, et al. Erythropoietin inhibits ferroptosis and ameliorates neurological function after spinal cord injury. Neural Regeneration Research. 2023; 18: 881–888.
[86]
Liu J, Huang Z, Yin S, Jiang Y, Shao L. Protective effect of zinc oxide nanoparticles on spinal cord injury. Frontiers in Pharmacology. 2022; 13: 990586.
[87]
Clemente LP, Rabenau M, Tang S, Stanka J, Cors E, Stroh J, et al. Dynasore Blocks Ferroptosis through Combined Modulation of Iron Uptake and Inhibition of Mitochondrial Respiration. Cells. 2020; 9: 2259.
[88]
Mao Y, Du J, Chen X, Al Mamun A, Cao L, Yang Y, et al. Maltol Promotes Mitophagy and Inhibits Oxidative Stress via the Nrf2/PINK1/Parkin Pathway after Spinal Cord Injury. Oxidative Medicine and Cellular Longevity. 2022; 2022: 1337630.
[89]
Wang H, Zheng Z, Han W, Yuan Y, Li Y, Zhou K, et al. Metformin Promotes Axon Regeneration after Spinal Cord Injury through Inhibiting Oxidative Stress and Stabilizing Microtubule. Oxidative Medicine and Cellular Longevity. 2020; 2020: 9741369.
[90]
Wang Y, Jiao J, Zhang S, Zheng C, Wu M. RIP3 inhibition protects locomotion function through ameliorating mitochondrial antioxidative capacity after spinal cord injury. Biomedicine & Pharmacotherapy. 2019; 116: 109019.
[91]
Rao S, Lin Y, Du Y, He L, Huang G, Chen B, et al. Designing multifunctionalized selenium nanoparticles to reverse oxidative stress-induced spinal cord injury by attenuating ROS overproduction and mitochondria dysfunction. Journal of Materials Chemistry. B. 2019; 7: 2648–2656.
[92]
Slusarczyk W, Olakowska E, Larysz-Brysz M, Woszczycka-Korczyńska I, de Carrillo DG, Węglarz WP, et al. Use of ebselen as a neuroprotective agent in rat spinal cord subjected to traumatic injury. Neural Regeneration Research. 2019; 14: 1255–1261.
[93]
Jia ZQ, Li SQ, Qiao WQ, Xu WZ, Xing JW, Liu JT, et al. Ebselen protects mitochondrial function and oxidative stress while inhibiting the mitochondrial apoptosis pathway after acute spinal cord injury. Neuroscience Letters. 2018; 678: 110–117.
[94]
Ge H, Xue X, Xian J, Yuan L, Wang L, Zou Y, et al. Ferrostatin-1 Alleviates White Matter Injury Via Decreasing Ferroptosis Following Spinal Cord Injury. Molecular Neurobiology. 2022; 59: 161–176.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top