Next Article in Journal
Pequi (Caryocar brasiliense, Camb) Bark Extract Reduces ROS Production in Diabetic Human Coronary Endothelial Cells
Previous Article in Journal
Role of Advanced Glycation End Products and Mitohormesis in Cancer Development and Progression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Stress, Inflammation, and Cellular Senescence in Neuropathic Pain: Mechanistic Crosstalk

1
Department of Surgery, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
2
Center for Molecular Medicine and Stem Cell Research, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
3
City Medical Emergency Department, 11000 Belgrade, Serbia
4
Department of Pathology, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
5
Department of Pathophysiology, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
6
Department of Internal Medicine, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
7
Department of Neurology, Faculty of Medical Sciences, University of Kragujevac, 34000 Kragujevac, Serbia
*
Authors to whom correspondence should be addressed.
These authors have contributed equally to this work and share first authorship.
Antioxidants 2025, 14(10), 1166; https://doi.org/10.3390/antiox14101166
Submission received: 8 September 2025 / Revised: 22 September 2025 / Accepted: 24 September 2025 / Published: 25 September 2025
(This article belongs to the Special Issue Chronic Pain and Oxidative Stress)

Abstract

Neuropathic pain is a chronic condition driven by intertwined mechanisms of oxidative stress, inflammation, and cellular senescence. Nerve injury and metabolic stress elevate reactive oxygen and nitrogen species, disrupt mitochondrial function, and activate the DNA-damage response, which stabilizes p53 and induces p16/p21-mediated cell-cycle arrest. These events promote a senescence-associated secretory phenotype (SASP) rich in cytokines, chemokines, and prostanoids that amplify neuroimmune signaling. In the spinal dorsal horn and dorsal root ganglia, microglia and astroglia respond to redox imbalance and danger cues by engaging NF-κB and MAPK pathways, increasing COX-2–dependent prostaglandin synthesis, and releasing mediators such as IL-1β and BDNF that enhance synaptic transmission and reduce inhibitory tone through KCC2 dysfunction. At the periphery, persistent immune-glial cross-talk lowers activation thresholds of nociceptors and sustains ectopic firing, while impaired autophagy and mitophagy further exacerbate mitochondrial dysfunction and ROS production. Collectively, these processes establish a feed-forward loop in which redox imbalance triggers senescence programs and SASP, SASP perpetuates neuroinflammation, and neuroinflammation maintains central sensitization—thereby consolidating a self-sustaining redox–senescence–inflammatory circuit underlying neuropathic pain chronicity.

1. Introduction

Neuropathic pain persists through tightly interwoven mechanisms of oxidative stress, inflammation, and cellular senescence, with Nrf2 functioning as an integrative hub that links these pathways to the initiation and maintenance of pain. Yet, despite rapid advances across redox biology, glial neuroimmunology, and senescence research, the field lacks a coherent framework that explains how mitochondrial dysfunction and ROS/RNS trigger DNA-damage responses and p16/p21-driven arrest, how the senescence-associated secretory phenotype amplifies microglial and astroglial signaling (e.g., NF-κB/MAPK, COX-2, IL-1β, BDNF), and how these processes remodel nociceptive circuits from primary afferents to the dorsal horn to consolidate central sensitization. This review’s objective is to articulate an integrative, pathway-level model that maps the crosstalk among oxidative stress, inflammation, and senescence across relevant cell types (neurons, microglia, astroglia, immune cells) and etiologies of neuropathic pain, closing the knowledge gap by synthesizing and critically interpreting contemporary experimental, translational, and clinical findings into a unified mechanistic narrative.

2. Foundations of Pain: From Protective Mechanism to Pathological Persistence

Pain represents a multidimensional sensory and emotional experience that functions as an essential protective mechanism, alerting the organism to actual or potential tissue injury [1]. It arises through the activation of specialized peripheral receptors and neural pathways that convey noxious information to higher brain centers, where it is integrated with cognitive and emotional components [2,3]. The International Association for the Study of Pain defines it as an unpleasant sensory and emotional experience related to, or resembling that associated with, tissue damage, thereby acknowledging its occurrence in both the presence and absence of ongoing structural injury [4]. Pain may arise from peripheral nociceptor activation subsequent to tissue injury, from pathological alterations within the somatosensory nervous system, or from modified nociceptive processing in the absence of definitive injury or disease [5,6]. The first mechanism, which is usually linked to acute, well-localized pain, typically subsides as the underlying damage heals [6]. In contrast, injury or dysfunction of neural pathways can produce pain that persists independently of tissue recovery, reflecting maladaptive signaling within the nervous system [7]. A third pathway involves disturbances in nociceptive modulation, often manifesting in conditions such as fibromyalgia or migraine, where central sensitization dominates [8]. In specific contexts, pain is exacerbated by inflammatory processes wherein immune-mediated responses sensitize peripheral receptors and sustain nociceptive input [9]. These mechanisms are not mutually exclusive and often coexist in chronic pain conditions, resulting in intricate clinical presentations that require meticulous mechanistic evaluation to inform targeted therapeutic approaches.

2.1. Definitional Evolution and Conceptual Challenges in Neuropathic Pain

The concept of neuropathic pain has undergone significant refinement over the past decades as experts have sought to achieve greater diagnostic precision [10]. The earliest widely adopted definition by the International Association for the Study of Pain (IASP) described it as pain initiated or caused by a primary lesion or dysfunction of the nervous system [11]. Although widely used, this wording was criticized because the term “dysfunction” was seen by many as too broad and unclear for diagnosis [12]. With this definition, some long-term pain conditions without clear nerve damage, such as fibromyalgia or irritable bowel syndrome, where the brain processes pain differently, could be wrongly labeled as neuropathic [13]. In 2008, a revised definition was proposed, restricting neuropathic pain to that arising as a direct consequence of a lesion or disease affecting the somatosensory system [14]. This update, approved by the IASP in 2011, left out conditions where problems in the motor system cause pain and removed “dysfunction” as a diagnostic criterion because it cannot be reliably confirmed [15].
Although the new definition is clearer, it still has limits. A problem in the sensory nervous system is required, but it is not enough on its own, since only some people with nerve damage actually develop neuropathic pain [15]. The pain experience is more often the result of maladaptive peripheral and central neuroplastic changes—such as sensitization, altered inhibition, and aberrant signaling—triggered by the lesion rather than the lesion itself [7,16].

2.2. Clinical Symptomatology of Neuropathic Pain

Neuropathic pain presents with a complex constellation of sensory abnormalities that reflect both gain and loss of function within the somatosensory system [17]. Patients frequently describe the pain as burning, shooting, stabbing, or accompanied by tingling sensations, which may arise spontaneously without any external stimulus [17,18]. A hallmark feature is the paradoxical combination of hypersensitivity and sensory deficit within the same affected area [19]. Hypersensitivity can manifest as pain evoked by normally non-painful stimuli, as in allodynia, or as an exaggerated response to painful stimuli, as in hyperalgesia [10]. These positive sensory phenomena are often accompanied by dysesthesia, characterized by unpleasant abnormal sensations, and paresthesia, involving abnormal but typically non-painful sensations such as pins and needles or numbness, which can nonetheless be distressing [10,20]. Negative sensory signs, including reduced sensitivity or complete loss of sensation in the painful region, frequently coexist, reflecting the underlying neuronal damage [17,21]. In many cases, symptoms persist or recur long after the initial injury has healed, and the pain may be exacerbated by movement or touch [21]. The chronicity of neuropathic pain is sustained by aberrant neural signaling and maladaptive neuroplastic changes, which contribute to the persistent mismatch between sensory input and perception, ultimately impairing physical function and psychological well-being [21].

2.3. Epidemiology of Neuropathic Pain

Estimating the true global burden of neuropathic pain remains challenging, largely due to the variability in diagnostic criteria and the absence of a universally accepted definition [22]. Population-based studies and systematic reviews suggest that chronic neuropathic pain affects a substantial proportion of adults, with prevalence rates ranging from approximately 3% to 17% depending on the population studied and the assessment tools employed [23]. Incidence estimates vary widely across specific syndromes: post-herpetic neuralgia and painful diabetic neuropathy represent some of the most common forms, while trigeminal and glossopharyngeal neuralgia occur less frequently [24]. The distribution of neuropathic pain is not uniform across demographic groups. Women appear disproportionately affected, accounting for over half of reported cases, and prevalence tends to rise with age, peaking in individuals between 50 and 64 years [25]. Occupational and environmental factors also influence disease distribution, with higher rates observed among manual laborers and residents of rural areas, suggesting that lifestyle, access to healthcare, and physical workload may contribute to risk [23].

2.4. Pathogenesis of Neuropathic Pain

Neuropathic pain emerges from a sustained neuroimmune dialog that reshapes excitability in peripheral and central pathways [26,27,28]. The increase in pro-inflammatory cytokines, particularly interleukin-1β (IL-1β), released by immune cells, microglia, and astroglia in the spinal cord is a key initiating event [29,30]. Through IL-1 receptor signaling, IL-1β activates nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and mitogen-activated protein kinase (MAPK) pathways, thereby amplifying the transcription of inflammatory genes and priming cyclo-oxygenase-2 (COX-2) induction [30]. COX-2 then drives prostanoid synthesis, particularly prostaglandin E2 (PGE2), which engages E-type prostanoid (EP) receptors to enhance neuronal firing and synaptic transmission (Figure 1) [31]. In addition to maintaining nociceptive gain and lowering activation thresholds, this cytokine–eicosanoid axis also initiates a feed-forward inflammatory loop that supports chronicity and treatment refractoriness [32,33].
At the periphery, nerve injury triggers a rapid, compartmentalized inflammatory response [34]. Schwann cells dedifferentiate and initiate myelin clearance at the lesion, while resident mast cells degranulate, releasing histamine, serotonin, leukotrienes, and growth factors that immediately sensitize nociceptors and recruit neutrophils [9,35]. Neutrophils, in turn, secrete algogenic mediators and promote the influx of monocyte-derived macrophages and T cells [9]. Together with resident macrophages and Schwann cells, these populations phagocytose degenerated axons/myelin and secrete cytokines/chemokines that further heighten excitability [9]. Through their direct action on injury-upregulated receptors on nociceptors, T-helper (Th) subsets provide layered control with pro-inflammatory cytokines such as IL-1β, tumor necrosis factor-alpha (TNF-α), and interleukin-17 (IL-17), as well as counter-regulatory cytokines including interleukin-4 (IL-4) and interleukin-10 (IL-10), thereby shaping continuous input to the central nervous system [9,35,36].
Altered impulse traffic from injured afferents is reinforced by channel remodeling [37]. Pain signaling in Aδ and C fibers depends on voltage-gated sodium channels (NaV) that generate and propagate action potentials. By adjusting peripheral NaV activity, nociceptive drive can be reduced before central amplification takes place [38]. Thermo-transient receptor potential channels (thermo-TRPs) on sensory nerve endings connect signals from inflammation with temperature changes. Substances like prostanoids and cytokines make these channels more sensitive, linking tissue inflammation with stronger pain responses [39,40]. Therefore, a logical way to prevent early amplification of neuropathic pain is to strategically target peripheral channel function and inflammatory mediators.
Peripheral events propagate centrally through both neural and vascular routes. Although the spinal cord is protected by the blood-medullary (blood–spinal cord) barrier, peripheral nerve injury can disrupt its integrity, permitting influx of circulating immune cells into the dorsal horn milieu [37]. At the same time, glial cells change their behavior [41]. Microglia multiply, become motile and phagocytic, show new receptors, and release pro-inflammatory signals [41,42]. Astrocytes also shift into a reactive state, keeping up the production of cytokines and chemokines [43]. Coordinated glial cell recruitment and activation cause a temporally excessive release of cytokines and chemokines, which fuel neuroinflammation and central sensitization and sustain neuropathic pain states [44]. The precise upstream signals that enforce prolonged glial overproduction after peripheral injury remain incompletely defined and are an active area of investigation [45].
Clinically, these processes result in a diverse set of symptoms that combine “positive” characteristics like spontaneous pain, paresthesia, and dysesthesia with “negative” indicators like sensory loss and, occasionally, cognitive or motor impairment [46]. From a mechanistic perspective, this represents long-lasting modifications in ion-channel expression, synaptic strength, inhibitory–excitatory balance, and gene transcription along pain pathways [1]. Understanding how cytokine signaling, peripheral neuroinflammation, ion-channel plasticity, barrier dysfunction, and glial activation contribute in stages aids in early diagnosis and mechanism-targeted treatments meant to break the vicious cycle of inflammation and excitation.

3. Inflammation and Neuropathic Pain

Neuropathic pain unfolds within a sustained inflammatory context that spans the injured peripheral nerve, dorsal root ganglia (DRG), and spinal cord [47,48]. Following nerve damage, resident Schwann cells and macrophages rapidly release danger-associated molecular patterns (DAMPs) and eicosanoids, while endothelial activation increases leukocyte adhesion and extravasation [27]. Monocytes, macrophages, lymphocytes, and plasma cells accumulate at the lesion, where they generate reactive oxygen and nitrogen species (ROS/RNS) and secrete cytokines and chemokines that intensify local excitability [27,49]. This influx of cells is not just a side effect. It changes the properties of pain-sensing nerves, making them more excitable, lowering the threshold needed to activate them, and causing abnormal ongoing firing. As a result, these nerves keep sending signals that continuously stimulate central pain circuits [50].
Pattern-recognition receptors (PRRs), especially Toll-like receptor 4 (TLR4) found on macrophages, microglia, and even pain-sensing neurons, detect danger signals. When activated, they trigger intracellular signaling through the myeloid differentiation primary response 88 (MyD88) and TIR-domain-containing adapter-inducing interferon-β (TRIF) pathways [51]. These signaling routes converge at NF-κB and MAPKs, including p38, c-Jun N-terminal kinase (JNK), and extracellular signal-regulated kinase (ERK), which switch on genes that produce TNF, IL-1β, IL-6, COX-2, and chemokines such as C-C motif chemokine ligand 2 (CCL2) and C-X3-C motif chemokine ligand 1 (CX3CL1) [52,53]. At the same time, interleukin-6 family cytokines activate the glycoprotein 130 (gp130)/Janus kinase–signal transducer and activator of transcription (JAK–STAT) pathway, which helps maintain inflammation and promotes astrocyte reactivity [53,54]. Prostaglandin E2, produced by COX-2, acts on EP receptors located on primary sensory neurons and dorsal horn neurons. This interaction enhances cyclic adenosine monophosphate–protein kinase A/protein kinase C (cAMP–PKA/PKC) signaling, promotes phosphorylation of ion channels, and strengthens synaptic transmission. In this way, cytokine production is directly linked to increased pain sensitivity [53,54].
Peripheral inflammation directly modulates nociceptor transduction [55]. Tumor necrosis factor and IL-1β increase Nav channel availability and suppress repolarizing potassium (K+) currents, thereby favoring repetitive neuronal firing [56,57]. Prostaglandin E2, oxidized lipids, and electrophiles sensitize transient receptor potential vanilloid 1 (TRPV1) and activate transient receptor potential ankyrin 1 (TRPA1) through covalent or redox interactions, thereby linking lipid peroxidation to thermal and chemical hyperalgesia [58]. Chemokine axes, such as C-C motif chemokine ligand 2–C-C chemokine receptor type 2 (CCL2–CCR2) and C-X3-C motif chemokine ligand 1–C-X3-C chemokine receptor 1 (CX3CL1–CX3CR1), recruit and position immune cells but also signal within neurons to enhance excitability [59,60]. Sensory neurons reciprocate by releasing pro-inflammatory mediators such as adenosine triphosphate (ATP), substance P, calcitonin gene-related peptide (CGRP), and various cytokines, which amplify vascular permeability and immune cell influx—thereby forming a bidirectional neuron–immune loop that resists spontaneous resolution [55].
Central neuroinflammation consolidates chronicity [61]. The movement of immune cells and DAMPs into the dorsal horn is made possible by peripheral input and barrier disruption at the blood–nerve and blood–spinal cord interfaces [62]. Microglia adopt a reactive phenotype characterized by NADPH oxidase 2 (NOX2)-driven ROS production and p38 MAPK activation, releasing IL-1β, TNF, and brain-derived neurotrophic factor (BDNF) [63,64]. Brain-derived neurotrophic factor activates its receptor tropomyosin receptor kinase B (TrkB), leading to downregulation of potassium-chloride cotransporter 2 (KCC2) in dorsal horn neurons. This process diminishes the inhibitory effects normally mediated by gamma-aminobutyric acid (GABA) and glycine, resulting in disinhibition—a defining feature of central sensitization [65]. Astrocytes sustain the inflammatory response through STAT3 and NF-κB signaling. These pathways maintain COX-2 activity, disrupt the physiological balance of glutamate, and promote the release of chemokines [66]. As a consequence, microglia remain activated and immune cells persist within the tissue. Together, these glial changes turn short-term pain signals into long-lasting strengthening of synapses and increased network excitability.
Oxidative stress interlocks with inflammation at every tier. Cytokine and TLR signaling activate NOX enzymes and inducible nitric oxide synthase (iNOS), which increase the production of ROS and RNS [67]. These molecules further stimulate NF-κB and MAPK pathways, while also inflicting damage on mitochondrial and nuclear DNA [68]. Mitochondrial danger signals and oxidized lipids activate the cyclic GMP–AMP synthase–stimulator of interferon genes (cGAS–STING) pathway and the NOD-, LRR-, and pyrin domain-containing protein 3 (NLRP3) inflammasome. This activation promotes the maturation of IL-1β and further amplifies the release of pro-inflammatory cytokines [69]. This cycle of oxidative stress and inflammation explains what is often seen in practice. Stronger inflammation is linked to stronger pain. By calming either the inflammatory or oxidative stress pathways, pain sensitivity can be reduced, even if the original injury itself does not change.
Neuropathic pain develops when several processes work together: immune cells move into the area, receptors sense danger signals and trigger cytokine release, prostanoids are produced, and glial cells remain active. Together, these create an inflammatory environment that keeps lowering the threshold for neurons to fire. To break this cycle, treatments need to match the mechanisms involved. Examples include blocking TLR4/NF-κB or JAK–STAT signaling, inhibiting COX-2 or EP receptors, and targeting chemokine systems like CCL2/CCR2 or CX3CL1/CX3CR1. Adding therapies that control oxidative stress, such as NOX or iNOS modulation, and treatments aimed at glial cells can strengthen these effects. The goal of these combined approaches is to dismantle the inflammatory support that maintains central sensitization and drives ongoing neuropathic pain.
Peripheral and Central Inflammation and the Role of Glial Cells in Neuropathic Pain
Peripheral inflammation sets the stage for chronic pain by recruiting and reprogramming innate and adaptive immune cells at and around the injured nerve [9,70]. As shown in Figure 2, neutrophils arrive first and can be analgesic in the acute phase by releasing endogenous opioids (β-endorphin, met-enkephalin, dynorphin A) and clearing pathogens [32]. When inflammation persists, they form neutrophil extracellular traps and amplify tissue damage [32]. Efficient efferocytosis—macrophage uptake of apoptotic neutrophils—helps terminate this response and accelerates pain resolution [71]. Monocytes differentiate into macrophages that orchestrate phagocytosis, antigen presentation, and cytokine production [72]. In the area around pain-sensing neurons, macrophages release molecules such as TNF, IL-1β, IL-6, PGE2, nerve growth factor (NGF), insulin-like growth factor 1 (IGF-1), and chemokines including CCL2 and C-X-C motif chemokine ligand 1 (CXCL1). These substances directly make the primary sensory neurons more sensitive by changing how ion channels, for example, by altering TRPA1 and TRPV1 activity, or by increasing the availability of voltage-gated sodium channels Nav1.7–Nav1.9 [73,74]. As a result, neurons become more excitable and may exhibit abnormal firing. Neuron–macrophage crosstalk is reinforced by CCL2/CCR2 signaling between DRG neurons and infiltrating macrophages [75]. In parallel, DAMPs such as high-mobility group box 1 (HMGB1) engage TLR4 and the receptor for advanced glycation end products (RAGE), and accelerate a CXCL12/CXCR4 axis within peripheral nerves, deepening nociceptor sensitization [73,76,77].
Macrophages exhibit functional plasticity that shapes pain trajectories [72]. Pro-inflammatory “M1-like” macrophages, triggered by signals such as lipopolysaccharide (LPS), granulocyte-macrophage colony-stimulating factor (GM-CSF), interferon-gamma (IFN-γ), and TNF, increase neuronal excitability and exacerbate tissue damage. In contrast, “M2-like” macrophages, activated by IL-4, IL-10, IL-13, transforming growth factor-beta (TGF-β), and macrophage colony-stimulating factor (M-CSF), contribute to debris clearance, secrete anti-inflammatory mediators, and promote tissue repair [72,78]. Specialized pro-resolving mediators facilitate the transition of macrophages toward a pro-healing phenotype [79]. For instance, neuroprotectin D1 (NPD1) acts on the G protein-coupled receptor 37 (GPR37) in macrophages, enhancing their capacity for debris clearance and increasing IL-10 production. This action shortens the duration of inflammation and contributes to pain resolution [80]. Thus, inflammation can be both foe (sensitization) and friend (resolution), depending on cellular programming and timing [32].
Central inflammation consolidates chronicity through glial mechanisms that convert transient peripheral input into durable changes in dorsal horn circuitry [81]. Microglia are the first responders in the spinal cord: within days of nerve injury they proliferate, change morphology, and adopt reactive transcriptional states (“microgliosis”) [42]. Signals from sensory neurons, such as IL-1β, caspase-6, colony-stimulating factor 1 (CSF1), and extracellular proteases, activate microglia. Once activated, microglia release pro-inflammatory and modulatory molecules, including TNF, IL-1β, PGE2, and BDNF. These factors further amplify pain signaling and contribute to central sensitization [42,72]. Brain-derived neurotrophic factor activates its receptor TrkB on dorsal horn neurons, leading to reduced activity of KCC2. Impaired KCC2 function disrupts chloride homeostasis, thereby weakening inhibition mediated by GABA and glycine. This loss of inhibition, termed disinhibition, represents a hallmark feature of central sensitization [65]. Microglia express a variety of receptors and ion channels that render them highly responsive to environmental signals. Adenosine triphosphate-gated receptors P2X purinoceptor 4 (P2X4) and P2X purinoceptor 7 (P2X7) activate microglia to release cytokines and initiate assembly of NLRP3 inflammasome through caspase-1 [82,83]. The voltage-gated proton channel H_v1 facilitates the generation of ROS via NOX enzymes and supports communication between microglia and astrocytes [84]. Following peripheral nerve injury, transient receptor potential vanilloid 4 (TRPV4) is upregulated, resulting in the release of lipocalin-2 (LCN2), which contributes to enhanced pain sensitivity [85]. Notably, microglial activation is most prominent early and tends to wane after weeks, which helps explain why microglia-targeted inhibitors show greatest efficacy as preventive or early interventions [63,86]. In line with current consensus, the M1/M2 nomenclature is considered obsolete; instead, microglia are described using context-specific, transcriptionally defined states by transcriptional programs and function [87]. Single-cell studies reveal discrete non-homeostatic states—such as disease-associated microglia (DAM) and injury/activated-response states—that emerge in neurodegeneration and after nerve injury [87].
After peripheral nerve injury, microglia engage a neuron–microglia communication pathway centered on fractalkine (CX3CL1) and its microglial receptor CX3CR1, which has emerged as a consistent feature across neuropathic pain models [88]. In the dorsal horn, membrane-anchored neuronal CX3CL1 is proteolytically shed by microglial cathepsin-S to generate soluble CX3CL1 that activates CX3CR1 and propagates pain signaling [89]. Blocking spinal cathepsin-S reverses established mechanical hypersensitivity in nerve-injured rats, functionally linking Cat-S–driven CX3CL1 release to maintenance of neuropathic pain [90].
In experimental autoimmune neuritis—the rat model of Guillain–Barré syndrome—mechanical allodynia appears days before neurological deficits and coincides with increased dorsal-horn microglia, indicating early central sensitization [91]. Within this model, spinal CX3CL1/CX3CR1 expression is upregulated, supporting a role for fractalkine signaling in pain associated with peripheral inflammatory neuropathies [91].
Astrocytes, more abundant than microglia, enter a reactive state (“astrogliosis”) slightly later but persist for months, thereby maintaining central sensitization [92]. When astrocytes adopt a reactive phenotype, they upregulate glial fibrillary acidic protein (GFAP) expression and begin to proliferate. Rather than carrying out their normal supportive functions—such as maintaining K+ homeostasis, clearing glutamate, and regulating water balance—they shift toward the production of pro-inflammatory mediators [32]. In this state, astrocytes continuously express COX-2 and PGE2, release cytokines and chemokines including CCL2, CXCL1, and CX3CL1, and generate NO via iNOS. Collectively, these changes sustain inflammation and amplify pain signaling. These signals potentiate excitatory transmission, impair inhibitory tone, and stabilize microglial activation, creating a self-reinforcing inflammatory network [32].
Oligodendrocytes, although less frequently studied in the context of pain, also contribute to its modulation [93]. Following injury in the dorsal spinal cord, they release interleukin-33 (IL-33). This cytokine binds to its receptor suppression of tumorigenicity 2 (ST2) expressed on neurons and glial cells, thereby activating phosphoinositide 3-kinase (PI3K) and MAPK pathways, including p38, ERK, JNK, as well as nuclear factor NF-κB signaling. These cascades enhance the sensitivity of pain pathways and promote nociceptive transmission [94]. Interactions among oligodendrocytes, microglia, astrocytes, and primary afferents likely fine-tune conduction and synaptic integration in ways that remain under active investigation. Figure 3 summarizes the key cellular interactions that sustain this process.
In sum, Neuropathic pain develops when immune, neuronal, and glial mechanisms reinforce one another to maintain a chronic inflammatory state. At the periphery, immune cell recruitment and macrophage signaling sensitize nociceptors, lowering their activation threshold and causing abnormal firing. In the spinal cord, microglia respond early to neuronal cues through receptors such as P2X4, P2X7, TLR4, and CSF1, releasing pro-inflammatory factors that embed these signals into neural circuits. Over time, astrocytes take on the role of sustaining inflammation through COX-2 activity, prostaglandin release, chemokine output, and glutamate imbalance, while even oligodendrocytes contribute by releasing IL-33 that further activates pain pathways. Because these processes evolve with time, therapeutic timing is crucial: early interventions may be most effective when targeting microglial receptors or upstream triggers, whereas persistent pain may respond better to astrocyte-directed strategies such as COX-2 or EP receptor inhibition and chemokine blockade. Combining these with approaches that promote resolution—such as specialized pro-resolving mediators, macrophage reprogramming toward an M2 phenotype, and enhanced efferocytosis—offers a way to dismantle the inflammatory support system that stabilizes central sensitization and sustains neuropathic pain. To integrate these findings, Table 1 maps peripheral and central inflammatory cascades to their effects on nociceptive processing and to actionable clinical strategies.

4. Oxidative Stress in Neuropathic Pain: Sources, Redox Signaling, and Effects on Nociceptive Circuits

Oxidative stress emerges from a pathological shift in the cellular redox balance, where ROS and RNS exceed the buffering capacity of antioxidant systems [95,96]. This imbalance disrupts redox signaling and promotes damage to DNA, proteins, and lipids, ultimately contributing to cellular dysfunction and degeneration [97]. Central to this phenomenon is the overaccumulation of ROS, such as superoxide anion (O2), hydrogen peroxide (H2O2), and hydroxyl radical (•OH) (Figure 4), alongside RNS species including nitric oxide (NO) and peroxynitrite (ONOO) [98]. Although these species can participate in physiological signaling at low concentrations, their excess triggers a cascade of deleterious biochemical reactions that impair cellular integrity and survival [97].
In neurons, mitochondria represent the principal endogenous source of ROS [99]. Within the mitochondrial electron transport chain (ETC), complexes I and III leak electrons that prematurely reduce molecular oxygen, generating superoxide [100]. This process is exacerbated under metabolic stress, aging, or impaired glucose utilization—conditions known to compromise mitochondrial efficiency and elevate radical production [101,102]. Additionally, peroxisomes, endoplasmic reticulum, and enzymatic systems such as NADPH oxidase isoforms (NOX1–5), dual oxidases (Duox1/2), and various nitric oxide synthases (NOS1–3) further contribute to intracellular ROS and RNS flux [103,104]. Notably, NOX enzymes are dedicated sources of regulated ROS for immune and signaling functions, yet their hyperactivation results in oxidative injury, especially within the nervous system [105,106,107].
Exogenous factors also intensify oxidative load, especially in environmentally exposed tissues. Ultraviolet (UV) radiation promotes the formation of mutagenic DNA lesions like 8-oxo-guanine and depletes key antioxidants such as glutathione (GSH) [108,109]. Ionizing radiation increases intracellular peroxide accumulation, while alpha particles enhance oxygen consumption and accelerate ROS generation in fibroblasts [110,111]. Toxic metals such as iron, copper, cadmium, and arsenic can accelerate chemical reactions that generate highly reactive molecules. In particular, they drive the Fenton and Haber–Weiss reactions, which produce hydroxyl radicals. These radicals are extremely damaging, as they attack and oxidize DNA, RNA, and proteins, leading to structural injury and loss of function in cells [107]. Exposure to ozone leads to airway inflammation and oxidant-mediated respiratory dysfunction, even in healthy individuals. Furthermore, xenobiotics such as pesticides, cigarette smoke, and industrial pollutants serve as potent ROS inducers through diverse mechanisms of redox cycling and mitochondrial interference [107].
Beyond cellular damage, redox imbalance critically modulates intracellular signaling cascades. One key redox-sensitive transcription factor is NF-κB, whose activation governs pro-inflammatory gene expression. ROS can modify cysteine residues on NF-κB or its upstream kinases, altering DNA binding capacity and transcriptional activity [95,112]. However, cellular redox regulators such as thioredoxin (Trx) and redox factor-1 (Ref-1) may restore NF-κB function by reducing oxidized cysteines [113,114]. Electrophilic lipid mediators, including cyclopentenone prostaglandins, antagonize ROS-mediated inflammation by covalently modifying redox-sensitive proteins, thereby mitigating NF-κB activation [115].
Importantly, sustained oxidative stress induces genomic instability, particularly double-strand DNA breaks, which are frequently observed in cancer and neurodegeneration [116,117]. These lesions may arise from non-enzymatic mitochondrial processes or from aberrant enzymatic activity under inflammatory conditions [96]. Thus, oxidative stress is not merely a byproduct of cellular dysfunction but a driver of pathophysiology, especially in tissues with high metabolic demands such as the central nervous system.

4.1. Oxidative Stress and Neuropathic Pain

Oxidative stress is a core driver of neuropathic pain, operating both as a consequence of nerve injury and as an amplifier of neuroinflammation and neuronal hyperexcitability [99]. The nervous system is intrinsically vulnerable because of high oxygen use, abundant polyunsaturated lipids, and long, mitochondria-rich axons [118,119]. After a peripheral or central injury, the normal balance between oxidants and antioxidants is disrupted. Instead of being tightly controlled, ROS and RNS stay elevated for long periods. Signals that would normally help the cell adapt are turned into harmful processes, causing oxidation and nitration of proteins, lipids, and nucleic acids throughout the tissue [104]. This redox imbalance couples to innate immune pathways, sustains cytokine output, and lowers activation thresholds across pain circuits [95,120].
Reactive oxygen and nitrogen species in the body primarily originate from specific enzymes and organelles [104]. Mitochondria represent a major source: when electrons leak from complex I or from the ubiquinol oxidation (Q_o) site of complex III, molecular oxygen (O2) is reduced to superoxide (O2) [121]. This unstable molecule is rapidly converted into hydrogen peroxide (H2O2) by mitochondrial superoxide dismutase 2 (SOD2) [122]. In the presence of free ferrous iron (Fe2+) or cuprous copper (Cu+) ions, H2O2 undergoes the Fenton reaction, generating hydroxyl radicals (•OH) [123]. These radicals are highly reactive and capable of inflicting severe cellular and molecular damage. Nicotinamide adenine dinucleotide phosphate oxidases—including NOX1, NOX2, NOX4, and dual oxidases 1 and 2 (DUOX1/2)—function as stimulus-coupled generators of superoxide (O2) and hydrogen peroxide (H2O2) in neurons, Schwann cells, microglia, and endothelial cells [124,125,126]. Nitric oxide synthases (nNOS, iNOS, eNOS) produce nitric oxide (NO•), which reacts with O2 to yield peroxynitrite (ONOO), a potent oxidant capable of inactivating mitochondrial enzymes such as aconitase and complex I, as well as damaging nuclear and mitochondrial DNA [127,128,129]. Additional sources of oxidative stress include endoplasmic reticulum (ER) oxidative protein folding mediated by endoplasmic reticulum oxidoreductin 1–protein disulfide isomerase (ERO1–PDI) and peroxisomal β-oxidation, both of which contribute to elevated H2O2 levels [130]. Pathological conditions such as tissue injury, hyperglycemia, chemotherapeutic exposure, aging, and defective mitophagy further exacerbate these fluxes, while insufficient nuclear factor erythroid 2–related factor 2 (Nrf2) activity or depletion of GSH reduces antioxidant buffering capacity—ultimately locking tissues into a chronic pro-oxidant state [97,131].
At the nociceptor, ROS or RNS directly remodel excitability [132]. Electrophiles and lipid peroxidation products, such as 4-hydroxynonenal (4-HNE), isoprostanes, and oxidized phospholipids, covalently modify TRPA1 and sensitize TRPV1, thereby lowering thermal and chemical activation thresholds and increasing channel open probability [133,134]. Nerve growth factor- driven lipid oxidation further potentiates TRPV1 function [135]. Redox modifications significantly influence voltage-gated ion channel activity. Nitration or S-nitrosylation of NaV channel subunits facilitates repetitive neuronal firing, while oxidation of voltage-gated potassium (Kv) channels diminishes their repolarizing capacity. In addition, hyperpolarization-activated cyclic nucleotide-gated (HCN) channel gating can shift to favor ectopic activity [132,136]. Mitochondrial dysfunction within axons impairs ATP-dependent pumps and Ca2+ homeostasis, promoting spontaneous discharges [137]. Collectively, these peripheral effects magnify the afferent barrage entering the spinal cord.
In the dorsal horn, redox signaling stabilizes central sensitization [107]. Microglia transition to a pro-inflammatory state with NOX2- and mitochondria-derived ROS, activating p38/JNK and NF-κB pathways to release IL-1β, TNF, and BDNF [63,64]. BDNF–TrkB signaling downregulates the K+–Cl cotransporter KCC2, weakening GABA/glycine-mediated inhibition and producing disinhibition [65]. Astrocytes, via STAT3 and NF-κB, sustain COX-2–PGE2 production and glutamate–glutamine cycle dysregulation; PGE2 acting at EP receptors enhances synaptic gain [138]. Redox changes can directly modulate neuronal mechanisms that amplify pain signaling. Specific residues on N-methyl-D-aspartate (NMDA) receptor subunits, including glutamate ionotropic receptor NMDA type subunit 1 (GluN1) and subunit 2 (GluN2), are highly sensitive to oxidative modifications, which enhance receptor activity [139]. Moreover, S-nitrosylation—a process in which NO covalently attaches to cysteine residues on synaptic proteins—further augments excitatory signaling [140]. Collectively, these modifications facilitate long-term potentiation within pain pathways, reinforcing and amplifying nociceptive input [139,140]. Peroxynitrite adds nitrate groups to proteins inside both mitochondria and the cytoplasm of dorsal horn neurons and glial cells. This damages enzymes needed for energy production and weakens antioxidant defenses. As a result, the cells lose stability, and their excitability remains high, which helps keep pain pathways overactive [141].
Oxidants also act as immune amplifiers. Mitochondrial DNA (mtDNA) leakage and oxidized DAMPs activate cGAS–STING pathway as well as TLR and RAGE signaling [142]. These cascades enhance NF-κB and interferon regulatory factor (IRF) transcriptional programs, while also promoting NLRP3 inflammasome activation, which drives the maturation of IL-1β and propagates neuroinflammation [143]. This biochemical–immune coupling links peripheral injury to sustained central cytokine excess.

4.2. Mitochondrial Dysfunction in Neuropathic Pain

Neurons rely on mitochondria to meet the high energetic cost of excitability and synaptic transmission [144]. ATP generated by oxidative phosphorylation powers Na+/K+-ATPase to re-establish membrane potentials after spikes and fuels presynaptic vesicle cycling and axonal transport [145]. Beyond bioenergetics, mitochondria shape intracellular Ca2+ dynamics. They buffer activity-dependent Ca2+ elevations through the mitochondrial calcium uniporter (MCU) and release Ca2+ via the mitochondrial sodium/calcium exchanger (NCLX), thereby modulating neurotransmitter release probability and shaping short-term synaptic plasticity [137,146]. In sensory neurons with long axons, the correct positioning of mitochondria at nodes of Ranvier and terminals is essential; impaired transport or local failure lowers energy reserve and predisposes to ectopic firing [147].
Mitochondrial quality control is preserved through a tightly regulated cycle of fusion and fission, coupled with organelle turnover. Fusion is mediated by mitofusin-1/2 (MFN1/2) and optic atrophy protein 1 (OPA1), whereas fission is driven by dynamin-related protein 1 (DRP1). These dynamics are coordinated with biogenesis, regulated by peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), nuclear respiratory factors 1 and 2 (NRF1/2), and mitochondrial transcription factor A (TFAM). Damaged mitochondria are selectively removed through mitophagy, a process orchestrated by PTEN-induced kinase 1 (PINK1) and the E3 ubiquitin ligase Parkin [148]. Disruption at any point in this network—such as DRP1-driven mitochondrial fragmentation, mitofusin-2 (MFN2) dysfunction, impaired biogenesis, or defective mitophagy—leads to the accumulation of depolarized, ROS-rich mitochondria within axons and somata [149]. Membrane contact sites with ER, termed mitochondria-associated membranes (MAMs), coordinate lipid exchange and inositol 1,4,5-trisphosphate receptor (IP3R)-mediated Ca2+ transfer. Under pathological conditions, maladaptive ER–mitochondria coupling promotes Ca2+ overload, cyclophilin D-dependent opening of the mitochondrial permeability transition pore (mPTP), and ultimately bioenergetic collapse [149,150].
Mitochondria are the dominant endogenous source of reactive oxygen species in neurons [100]. Electron leak at complex I and at the Qo site of complex III reduces O2 to superoxide, which SOD2 converts to H2O2; in the presence of labile iron or copper, H2O2 yields hydroxyl radicals via Fenton chemistry [123]. When cardiolipin, a key lipid in the inner mitochondrial membrane, becomes oxidized, it disrupts the stability of respiratory supercomplexes that normally organize the electron transport chain. This destabilization weakens efficient energy transfer and loosens cytochrome c from the cristae structure. The freed cytochrome c both amplifies the production of reactive oxygen and nitrogen species and promotes swelling of mitochondria. Ultimately, it leaks into the cytosol, where it signals cell stress or death, while the mitochondria themselves lose their ability to carry out normal oxidative phosphorylation and energy production [151,152]. When antioxidant systems (e.g., glutathione, peroxiredoxins, thioredoxin) are depleted or outpaced, ROS or RNS oxidize proteins, lipids, and DNA, initiating DNA-damage responses and long-lived changes in neuronal function [149].
These mitochondrial derangements couple directly to the initiation and persistence of neuropathic pain [148]. When energy production drops, the Na+/K+-ATPase pump works less effectively. This makes it harder for neurons to maintain their resting potential, so Aδ and C fibers become unstable and start firing on their own [132]. At the same time, calcium balance is disturbed, leading to excess neurotransmitter release and activation of calcium-dependent kinases. These kinases add phosphate groups to ion channels, which makes neurons even more excitable [132]. Reactive oxygen and nitrogen species and lipid peroxidation products—such as 4-HNE and isoprostanes—covalently modify TRPA1and sensitize TRPV1, thereby lowering thermal and chemical activation thresholds. In parallel, redox modifications of NaV channels and Kv channels facilitate high-frequency neuronal firing and diminish the repolarizing reserve [153]. Mitochondrial NO• and peroxynitrite nitration deactivate key metabolic enzymes (e.g., aconitase, complex I), deepening energy failure and increasing oxidative load—a self-reinforcing loop that sustains peripheral hyperexcitability [148,153].
Structural and functional evidence supports this link across pain models. Ultrastructural studies show clear mitochondrial damage in different types of nerve cells. In peripheral sensory axons, DRG neurons, and Schwann cells, mitochondria appear swollen and full of vacuoles. This pattern is seen in conditions such as diabetes, chemotherapy-induced neuropathy, and experimental models like sciatic nerve constriction. These changes are consistent with mitochondria losing their ability to breathe properly and regulate calcium, which are both essential for healthy nerve function [154]. Manipulating mitochondrial function in nociceptors modifies pain phenotypes: pharmacologic or genetic interventions that restore respiration, limit ROS, or normalize Ca2+ flux reduce mechanical and thermal hypersensitivity in neuropathic and inflammatory models [154,155]. Hyperalgesic priming offers a mechanistic example of how mitochondrial dysfunction sustains pain. Transient inflammation elevates ATP synthase c-subunit lysine N-methyltransferase (ATPSc-KMT), which alters respiratory efficiency and impairs the resolution of pain. Experimental interventions—such as knocking down ATPSc-KMT, moderating mitochondrial respiration, or supplementing depleted metabolites—restore resolution and prevent the transition to chronic pain [156].
Therapeutically, mitochondria offer multiple actionable targets. Upstream strategies reduce oxidant production (e.g., limiting reverse electron transport at complex I, optimizing substrate supply, or recoupling NOS), while downstream approaches stabilize the organelle (cardiolipin-targeted peptides), enhance quality control (PGC-1α agonism, mitophagy support), or directly scavenge mitochondrial ROS (matrix-targeted antioxidants). Because energetic, Ca2+, and redox axes are interdependent, combination regimens—pairing mitochondrial support with glial modulators or redox-gated channel antagonists—are most likely to disrupt the bioenergetic–inflammatory–excitatory loop that maintains neuropathic pain. A concise mapping of oxidative nodes to pain mechanisms and actionable therapies is provided in Table 2.

5. Cellular Senescence in Neuropathic Pain: Triggers, SASP, and Neural Consequences

Cellular senescence, first described by Hayflick and Moorhead in the early 1960s, refers to the irreversible cessation of cell proliferation after a finite number of divisions [96]. Initially characterized in human fibroblasts, this process—termed replicative senescence—was later attributed to progressive telomere shortening, which induces chromosomal instability and functions as a tumor-suppressive mechanism [157,158]. Over time, senescence has been recognized as a multifaceted biological program that operates not only as a safeguard against malignant transformation but also as a participant in diverse physiological processes, including embryonic development, wound healing, and tissue regeneration [159,160]. Senescent cells remain metabolically active and adopt a distinctive secretory profile known as the senescence-associated secretory phenotype (SASP), which comprises pro-inflammatory cytokines, chemokines, growth factors, and proteases [161]. While SASP can facilitate tissue remodeling and repair, its chronic activation drives persistent low-grade inflammation, thereby contributing to aging and the pathogenesis of age-related diseases [162].
Different forms of cellular senescence, such as replicative, oncogene-induced, therapy-induced, stress-related, mitochondrial dysfunction–associated, and immune-driven, share common traits [163]. They all involve the activation of cell cycle inhibitors, including p53, p21WAF1/Cip1, and p16INK4a, as well as the activity of senescence-associated β-galactosidase (SA-β-gal) [164]. However, the exact molecular features and the functional consequences vary depending on the type of stress that triggered the process and the specific cellular environment.

5.1. Sources and Types of Cellular Senescence

Senescence can be initiated by a variety of physiological and pathological stressors that converge on pathways enforcing irreversible cell cycle arrest (Figure 5) [164]. Classic triggers include telomere attrition, which underlies replicative senescence, and persistent DNA damage, commonly manifesting as double-strand breaks that activate the DNA damage response (DDR) [165]. The DNA damage response signaling cascade recruits key factors such as phosphorylated histone H2AX (γH2AX), mediator of DNA damage checkpoint protein 1 (MDC1), and p53-binding protein 1 (53BP1) to sites of DNA lesions [166,167,168]. This recruitment sustains activation of the ataxia-telangiectasia mutated (ATM), ataxia-telangiectasia and Rad3-related (ATR), checkpoint kinase 1 (CHK1), and checkpoint kinase 2 (CHK2) pathways, which in turn stabilize tumor protein p53 (p53) and promote p21WAF1/Cip1-mediated inhibition of cyclin-dependent kinases (CDKs) [168,169].
Oncogene-induced senescence results from aberrant activation of oncogenes such as NRASG12V or BRAFV600E, enforcing growth arrest through p21 and p16 pathways [161,170]. Therapy-induced senescence occurs in response to genotoxic cancer treatments, including chemotherapy and radiation, which trigger DDR activation [171]. Stress-induced senescence is typically telomere-independent and arises from physical or chemical insults, including oxidative stress and DNA-damaging agents [96]. Mitochondrial dysfunction-associated senescence (MiDAS) develops following mitochondrial damage, producing a secretory phenotype distinct from classical SASP and notably lacking IL-1-driven inflammation [172]. Immunologically induced senescence emerges in the presence of excessive pro-inflammatory cytokines, such as IL-17, and is associated with altered Wingless/Integrated (WNT) signaling and extracellular matrix remodeling [173].

5.2. Phenotypic Features of Cellular Senescence and the Senescence-Associated Secretory Phenotype

Cellular senescence represents a complex and multifaceted cellular program characterized primarily by irreversible cell cycle arrest and profound changes in cellular morphology, gene expression, and secretory behavior [164]. The arrest is predominantly mediated by CDKs inhibitors such as p21Cip1/WAF1 and p16INK4a, which suppress CDK4/6 activity, enforce retinoblastoma (Rb) hypophosphorylation, and consequently silence E2 promoter binding factor (E2F)-dependent transcription required for G1-S progression [164,174,175]. This sustained proliferative arrest distinguishes senescent cells from quiescent or terminally differentiated cells.
Senescent cells acquire distinct morphological features, including an enlarged and flattened shape, increased granularity, and often nuclear irregularities [176,177]. Biochemically, they accumulate lipofuscin and show enhanced activity of lysosomal β-galactosidase, commonly detected as SA-β-gal [177,178]. However, the specificity of SA-β-gal as a standalone biomarker is limited due to its expression in certain non-senescent cells, such as macrophages [179]. Therefore, contemporary approaches recommend a multi-marker strategy, combining SA-β-gal with cell cycle arrest markers such as p16INK4a and p21Cip1/WAF1, indicators of DNA damage including γH2AX, and the absence of proliferation markers such as Ki-67, proliferating cell nuclear antigen (PCNA), or bromodeoxyuridine (BrdU) incorporation [180]. Novel detection approaches have emerged, including mass cytometry (CyTOF)-based panels for simultaneous multiparameter assessment and transcriptomic classifiers such as the SenMayo gene set and SenCID, a machine-learning algorithm trained on large senescence transcriptome datasets. These tools aim to delineate senescent cell populations across tissues and pathophysiological states with higher specificity and resolution [181].
A hallmark of senescent cells is their resistance to apoptosis, coupled with the acquisition of a pro-inflammatory and tissue-remodeling secretory profile known as the senescence-associated secretory phenotype (SASP) [182]. The senescence-associated secretory phenotype encompasses a wide repertoire of factors, including interleukins such as IL-1β, IL-6, and IL-8; chemokines such as CCL2, CCL5, and CXCL1; growth factors including TGF-β, growth differentiation factor 15 (GDF15), and epidermal growth factor (EGF); matrix metalloproteinases such as MMP-1 and MMP-3; as well as bioactive lipids [164]. Additionally, SASP includes extracellular vesicles, cytoplasmic chromatin fragments, and non-coding RNAs, which contribute to intercellular communication and paracrine propagation of senescence [164].
Multiple signaling pathways converge to regulate the SASP, notably NF-κB, p38 MAPK, mechanistic target of rapamycin (mTOR), CCAAT/enhancer-binding protein beta (C/EBPβ), and cGAS–STING axis [164]. These pathways are activated in response to persistent DNA damage and other cellular stressors, driving transcriptional programs that reinforce the inflammatory and pro-fibrotic features of the SASP [164]. While SASP factors can exert beneficial effects in acute contexts, such as wound healing and tumor suppression, their chronic activation is implicated in tissue degeneration, immunosenescence, and age-related pathologies [162].

5.3. Oxidative Stress as a Driver and Sustainer of Cellular Senescence

Oxidative stress is a central biological force in the initiation and perpetuation of cellular senescence [96]. Excess ROS disturb redox homeostasis [183]. They damage important molecules, especially DNA and mitochondria, and they activate stress signals that force the cell to stop dividing permanently [184,185]. One of the earliest consequences of ROS overload is the activation of DDR, often via oxidative lesions at telomeric repeats [186]. This triggers p53 stabilization and transcriptional induction of cyclin-dependent kinase inhibitors, notably p21CIP1 and p16INK4a, culminating in irreversible growth arrest and the formation of senescence-associated heterochromatic foci (SAHF) [168,187,188,189].
Mitochondrial dysfunction represents both a hallmark and a driver of oxidative stress–induced senescence [96]. Damage to mtDNA, which is not histone-protected and is not effectively repaired, compromises the integrity of the respiratory chain, hinders oxidative phosphorylation, and increases electron leakage from complexes I and III, all of which contribute to the production of additional ROS [149,185,190]. This mitochondrial ROS production feeds back to exacerbate nuclear and organellar damage, creating a self-sustaining oxidative loop that stabilizes the senescent phenotype [191]. Peroxisomal insufficiency further aggravates redox imbalance, as catalase-mediated detoxification of hydrogen peroxide is diminished in senescent cells [192,193].
In response to mitochondrial distress, senescent cells attempt to restore energy homeostasis through a signaling cascade involving ATM, protein kinase B, mechanistic target of rapamycin complex 1 (mTORC1), and PGC-1α, which collectively drive mitochondrial biogenesis [194]. However, in the senescent context, these newly formed mitochondria are often functionally compromised, maintaining high ROS output [191]. A parallel metabolic shift toward aerobic glycolysis (Warburg-like effect) provides ATP independent of the impaired electron transport chain but does not resolve chronic energetic stress [195]. This is reflected in altered nicotinamide adenine dinucleotide (NAD+) or reduced NADH ratios, sustained activation of AMP-activated protein kinase (AMPK), and stabilization of p21Cip1/WAF1 and p16INK4a messenger RNA (mRNA) transcripts via human antigen R (HuR) inhibition, collectively reinforcing p53–retinoblastoma protein checkpoint activation [196,197].
ROS not only activate DDR but also shape the SASP [96]. Through redox-sensitive signaling pathways such as growth arrest and DNA damage-inducible alpha (GADD45A)–p38 MAPK–growth factor receptor-bound protein 2 (GRB2)–TGF-β, oxidative stress sustains p21Cip1/WAF1 expression and promotes pro-inflammatory SASP secretion [96]. This secretome propagates paracrine senescence, maintains chronic inflammation, and contributes to tissue degeneration [198]. Mitochondrial ROS and telomeric ROS affect each other in both directions. ROS from mitochondria speed up telomere shortening, while oxidative damage at telomeres weakens mitochondrial function. This two-way interaction keeps redox imbalance locked into the processes that drive cell senescence [199].
The DDR sits at the intersection of oxidative stress and mitochondrial decline [96]. Persistent double-strand breaks, marked by γH2AX and 53BP1 foci, maintain checkpoint signaling and SASP expression [166]. DDR activation feeds forward into mitochondrial dysfunction via GADD45A–p38 MAPK signaling, enhancing ROS output and locking cells into a senescent state [96]. Pharmacological ROS scavenging, particularly when combined with mTOR inhibition, can restore mitochondrial membrane potential, reduce double-strand breaks, and partially reverse senescence-associated defects.
The p53–p21 and p16–Rb pathways are principal executors of oxidative stress–driven growth arrest [96]. Reactive oxygen species -activated ATM and ATR kinases phosphorylate CHK1 and CHK2, leading to p53 stabilization, p21Cip1/WAF1 induction, and enforcement of G1/S and G2/M cell cycle checkpoint blockade [166]. In parallel, ROS or p38 MAPK–driven transcriptional programs increase p16INK4a expression, maintaining Rb in its hypophosphorylated, growth-suppressive state [164]. Together, these pathways silence E2F target genes required for S-phase entry and mitotic progression [200].
An important amplifier of this process is thrombospondin-1 (TSP1), a matricellular glycoprotein upregulated in senescent cells by DNA damage, TGF-β1, and oxidative stress [201]. TSP1 engages CD47 to activate Nox1, increasing ROS production and further stabilizing p53–p21 signaling [201]. This TSP1–CD47–Nox1 axis also inhibits nitric oxide signaling, tilting the redox balance toward oxidative dominance, impairing angiogenesis, and promoting extracellular matrix remodeling in aging tissues [202].
Finally, oxidative stress remodels the senescent cell transcriptome by repressing genes involved in DNA replication and mitosis, altering vesicular trafficking and cell adhesion networks, and upregulating anti-apoptotic B-cell lymphoma 2 (BCL-2) family members. These adaptations enable long-term survival of senescent cells despite cumulative cellular and molecular damage [203]. In this way, ROS are not passive byproducts but active architects of the senescent phenotype, driving both intracellular damage programs and extracellular inflammatory signaling.

5.4. Senescence and Inflammation

Inflammation and cellular senescence are intricately linked through a self-reinforcing cycle in which inflammatory signaling not only triggers but also sustains and amplifies the senescent phenotype [204]. Chronic exposure to pro-inflammatory cytokines, chemokines, and lipotoxic signals contributes to the induction of senescence in multiple cell types, often in synergy with oxidative stress and DNA damage [204,205]. This relationship is central to age-related dysfunction, cancer biology, and impaired tissue regeneration [204,205,206].
Senescence can be promoted by persistent inflammatory stimuli, including IL-1β, IL-6, IL-8, TNF-α, and C-C motif chemokine ligand 11 (CCL11), which act through redox-sensitive pathways and transcriptional programs such as NF-κB, C/EBPβ, and p38 MAPK [198]. These pathways converge on cell cycle regulators like p16INK4a, p21CIP1, and p15INK4b, enforcing growth arrest [207]. Experimental models have demonstrated that exposure to moderate levels of IL-6 or IL-8 induces senescence in pre-damaged or primed cells, with IL-6 acting predominantly via autocrine loops, while IL-8 functions through CXCR2-mediated paracrine signaling [207,208,209]. Inhibition of either cytokine or their receptors delays senescence, underscoring their essential regulatory roles [207].
This inflammatory network is further amplified by feedback mechanisms [207]. For instance, C/EBPβ upregulates IL-6, which in turn sustains C/EBPβ expression, reinforcing the pro-senescent circuit [210]. Simultaneously, the induction of CXCR2 and its ligands (e.g., IL-8, GROα) contributes to a broad chemokine signaling program that stabilizes the senescent state [207]. This mechanism is evident in both replicative and oncogene-induced senescence and is mirrored by the expression patterns observed in premalignant lesions, fibrotic tissue, and aged organs [207].
Inflammation-induced senescence is not restricted to classical cytokines. Lipid-mediated inflammation, such as that driven by dysfunctional adipocytes, promotes immune senescence through the expansion of pro-inflammatory B and T cell subsets, and is linked to telomere attrition and tissue degeneration in metabolic diseases [198]. Furthermore, persistent low-grade inflammation in tissues such as endothelium and epithelium contributes to senescence by driving mitochondrial dysfunction, telomeric instability, and DNA damage—often through ROS-dependent mechanisms [205].
Senescent cells themselves become potent sources of inflammatory mediators via the SASP, perpetuating inflammation both locally and systemically [211]. This SASP-driven immune modulation not only exacerbates tissue dysfunction but also attracts immune cells [162,212]. Interestingly, in cancer models, this process facilitates the clearance of senescent tumor cells by macrophages, NK cells, and neutrophils, demonstrating a dual role of inflammation in both executing and resolving senescence [213].

5.5. Senescence and Neuropathic Pain

Once considered exclusive to dividing cells, senescence hallmarks are now documented in post-mitotic neurons and glia that shape pain processing [214,215]. The presence of senescence-like signatures in human pyramidal and cortical neurons, such as p21CIP1/p16INK4a induction, SA-β-gal activity, persistent DNA-damage foci, and a pro-inflammatory secretome, suggests a conserved stress program that prevents maladaptive cell-cycle re-entry and may temporarily prevent neuronal loss [215,216,217]. In neurodegenerative settings, genetic or pharmacologic clearance of senescent neurons and glia improves molecular and behavioral outcomes, establishing senescence as a modifiable disease node rather than a mere epiphenomenon [215]. In pain pathways, peripheral nerve injury increases p53-positive neurons, astrocytes, and microglia in the spinal cord, while age and injury drive a senescent phenotype in primary sensory neurons of lumbar DRG, identified by p21/p16 expression, IL-6 upregulation, and SA-β-gal activity [215,218].
Mechanistically, “post-mitotic cellular senescence” (PoMiCS) in neurons is best viewed as a fail-safe response to stress-driven cell-cycle re-entry [214]. Different stressors relevant to neuropathic pain, such as axotomy, restriction of growth factors, chemotherapy, oxidative stress, or genotoxic stress, can push mature neurons into an abnormal attempt to re-enter the S phase of the cell cycle [219]. Induction of cyclin-dependent kinase inhibitors (p21, p16) arrests this trajectory, averts mitotic catastrophe, and preserves cell number [214,220]. The result is functional: continuous activation of checkpoint and DNA-damage pathways traps neurons in a state with altered metabolism and increased secretion [215]. This state changes their normal excitability and disrupts how they integrate signals at the synapse. Thus, transient neuronal senescence may be adaptive early after injury, whereas persistent senescence becomes a pathogenic driver of chronic pain [215].
Senescent nociceptors act as cytokine hubs. After a peripheral injury, some DRG neurons enter a senescent program and begin producing IL-6 as part of their SASP. This IL-6 does not just mirror the signals from macrophages and Schwann cells but actively adds to them, amplifying the local inflammatory response [215,221]. Interleukin-6 signals through gp130/ JAK–STAT pathways in autocrine and paracrine loops to enhance membrane excitability, modulate ion channel expression, and reinforce pain-related transcriptional programs [222,223]. Notably, TRPV1+ nociceptors—a neuronal population central to thermal and chemical nociception—expand with age and following injury, while co-expressing p21Cip1/WAF1 and IL-6 [215,224]. This provides a direct cellular conduit linking senescence to hyperalgesia [215]. Across traumatic, diabetic, and chemotherapy-induced models, nociceptors consistently show high enrichment of senescence features, linking their sustained SASP output to the maintenance of hypersensitivity.
Glial senescence consolidates central sensitization [218]. Chronic peripheral injury can induce a senescent phenotype in spinal astrocytes, characterized by SA-β-gal positivity, p21/p16 induction, and γH2AX foci, alongside a SASP rich in cytokines, chemokines, and prostanoids that potentiates dorsal horn transmission [218]. This mirrors astrocytic SASP profiles seen in aging-related brain disorders and helps explain the long tail of central neuroinflammation after peripheral lesions [225]. Microglia, likewise, can adopt senescence-like states that shift from trophic support to cytokine production, amplifying neuronal disinhibition and synaptic gain [218]. After spinal cord injury, most senescence-marker-positive cells around lesion borders are neurons, and senescent cortical neurons possess a SASP capable of inducing paracrine senescence in neighboring cells—evidence that senescence signals can propagate within neural microenvironments [226].
These patterns in time and place help refine treatment strategies. Studies show that senolytics given only in the spinal cord bring relief for a short time. This means that clearing senescent cells in the cord is not enough, because the main sources remain outside the cord in the DRG and injured nerves, where senescent nociceptors continue to release harmful SASP factors [218]. By contrast, selectively targeting senescent cells within the nociceptive lineage (source control) or dampening their secretory program (signal control) offers a mechanistic route to durable benefit [227]. From a practical perspective, this points to stage-specific treatment combinations. At first, it may be helpful to allow short-term adaptive senescence, which can protect neurons. Later on, therapy should shift toward removing long-lasting senescent nociceptors through BCL-2–dependent senolysis, blocking harmful SASP signals with JAK/STAT or p38 inhibitors or with IL-6/IL-6R blockade, and restoring balance in redox and DNA-damage pathways that keep cells locked in the senescent state. In sum, neuropathic pain can be reframed as a circuit disorder in which post-mitotic senescence of neurons and glia transitions from protective pause to pathogenic persistence—one that is mechanistically tractable if the right cellular sources and timelines are targeted. Table 3 offers a systematic overview of senescence triggers, pathways, markers, circuit consequences, and practical treatment approaches.

6. Nrf2 as a Central Integrator of Oxidative Stress, Inflammation, Senescence, and Neuropathic Pain

Nuclear factor erythroid 2–related factor 2 (Nrf2) has emerged as a master regulator of cellular defense, positioned at the crossroads of redox signaling, inflammation, senescence, and neuropathic pain [228,229]. Through its transcriptional control over antioxidant response elements (AREs), Nrf2 orchestrates the expression of a broad array of cytoprotective genes involved in maintaining redox homeostasis, mitochondrial function, and anti-inflammatory responses [229]. These adaptive mechanisms are critical in mitigating oxidative damage, sustaining energy metabolism, and preserving the structural and functional integrity of both neuronal and non-neuronal cells exposed to chronic stress [99].
The interplay between oxidative stress, immune activation, and cellular senescence is a defining feature of neuropathic pain and age-associated pathologies [107,230,231]. By controlling this crosstalk, Nrf2 acts as a modulator and sentinel, inhibiting detrimental inflammation, lowering glial reactivity, and postponing the onset of SASP [99,232]. Moreover, Nrf2 brings together signals from inside the cell and from the environment to control how cells respond to stress. It does this through flexible interactions with key metabolic pathways, such as mTOR, and through feedback loops that involve mitochondrial signals and autophagy [233,234,235]. This integrative role makes Nrf2 a promising therapeutic target not only for alleviating chronic pain but also for delaying degenerative processes driven by oxidative and inflammatory damage.

6.1. Nrf2 as a Central Regulator of Redox Homeostasis and Cellular Stress Responses

Nuclear factor erythroid 2–related factor 2 plays a pivotal role in orchestrating the transcriptional response to oxidative and electrophilic stress, acting as a master regulator of genes involved in antioxidant defense, detoxification, and cellular metabolism [229,236]. Under basal conditions, Nrf2 is sequestered in the cytoplasm through its interaction with Kelch-like ECH-associated protein 1 (Keap1), which serves as a redox-sensitive adaptor protein for Cullin 3-based E3 ubiquitin ligase complexes [236,237]. This interaction promotes the ubiquitination and proteasomal degradation of Nrf2, maintaining low steady-state levels within the cell [236,237].
Upon exposure to ROS or electrophilic agents, specific cysteine residues within Keap1 undergo covalent modifications that disrupt its ability to target Nrf2 for degradation [236,237]. As a result, Nrf2 stabilizes, accumulates in the cytoplasm, and translocates into the nucleus, where it binds to AREs in the promoter regions of its target genes. For effective ARE binding, Nrf2 forms heterodimers with small Maf proteins [236,237]. The transcriptional program activated by Nrf2 includes a broad spectrum of cytoprotective genes, notably those encoding heme oxygenase-1 (HO-1), NAD(P)H quinone oxidoreductase 1 (NQO1), glutathione S-transferases, and enzymes involved in glutathione biosynthesis [236]. HO-1, in particular, degrades heme into bioactive molecules such as carbon monoxide (CO) and bilirubin—each with potent anti-inflammatory and antioxidant properties [238]. Additionally, Nrf2 modulates inflammation by suppressing NF-κB signaling through CO-mediated inhibition and by promoting anti-inflammatory cytokine expression [239]. This process is schematically illustrated in Figure 6, which depicts the regulation of Nrf2 under basal and stress conditions, its release from Keap1, nuclear translocation, and transcriptional activation of cytoprotective target genes.
Beyond redox regulation, Nrf2 integrates signals from nutrient, metabolic, and stress pathways [240]. Its stability and activity are further modulated by autophagy-related protein p62, which binds Keap1, thereby interfering with Nrf2 degradation. This p62-mediated Nrf2 activation becomes especially relevant under impaired autophagic flux [241]. Moreover, the phosphoinositide 3-kinase/protein kinase B/glycogen synthase kinase-3 beta (PI3K/Akt/GSK-3β) signaling axis indirectly stabilizes Nrf2 by inhibiting GSK-3β–driven nuclear export and proteasomal degradation [242]. Nrf2 transcriptional activity is also enhanced by its interaction with nuclear receptors such as retinoid X receptor alpha (RXRα), which cooperatively bind to gene promoters to bolster antioxidant gene expression under stress conditions [236].
Although Nrf2 activation is generally considered cytoprotective, its dysregulation can lead to deleterious consequences. Excessive or constitutive Nrf2 activity, often driven by somatic mutations or oncogenic signaling, has been implicated in promoting tumor progression and resistance to chemotherapy by increasing antioxidant capacity, enhancing drug efflux, and suppressing apoptosis [243,244]. Chronic Nrf2 activation has also been linked to fibrosis, autoimmune pathologies, and metabolic dysfunctions, including insulin resistance and steatosis [245,246,247]. In certain cancers, upregulated Nrf2 promotes survival of malignant cells under oxidative and genotoxic stress, diminishing the efficacy of conventional therapies [243,244].

6.2. Nrf2-Mediated Modulation of Oxidative Stress in Neuropathic Pain

The transcription factor Nrf2 plays a central role in counteracting oxidative damage and preserving redox balance, which is essential in modulating nociceptive signaling and protecting neural structures involved in pain transmission [99]. By orchestrating the expression of a battery of antioxidant and cytoprotective genes, Nrf2 reduces the burden of ROS, thereby curbing neuronal hyperexcitability and the sensitization processes underlying neuropathic pain [228].
Within sensory pathways, Nrf2 activation enhances the transcription of enzymes such as HO-1, NQO1, SOD, catalase, and glutathione-related enzymes [236,248]. This upregulation improves the intracellular antioxidant capacity, leading to decreased nociceptor excitability and reduced glutamate-mediated excitotoxicity at synapses [99]. Nrf2 also limits the synthesis of pro-inflammatory mediators, thereby attenuating neuroinflammation—a critical component of chronic pain pathophysiology [249]. In addition to modulating redox status, Nrf2 exerts regulatory control over ion channels and neurotransmitter release, influencing the excitability of both peripheral sensory neurons and central pain circuits [250].
Neuroprotective effects of Nrf2 have been demonstrated in various experimental models. Pharmacological or genetic Nrf2 induction, especially via HO-1, was linked to strong antinociceptive effects in rodent models of inflammatory and neuropathic pain [228,251]. Agents such as cobalt protoporphyrin, curcumin, diosmetin, and sulforaphane restored antioxidant defenses, inhibited glial activation, and suppressed the release of excitatory mediators [252,253,254]. In contrast, Nrf2 inhibition, either directly or through downregulation of its downstream targets, was linked with aggravated pain behaviors and increased ROS levels [255]. Notably, a variety of treatments, such as electroacupuncture, angiotensin receptor blockade, and the administration of flavonoids, have been shown to exert analgesic effects via the Nrf2/HO-1 axis [253,256,257].
Experimental studies further demonstrated that Nrf2 is widely expressed along the nociceptive axis, including the spinal cord, DRG, and peripheral nerves [255,258,259]. Alterations in Nrf2 and HO-1 levels have been reported in various neuropathic pain models, such as those induced by paclitaxel, oxaliplatin, or sciatic nerve constriction [258,259]. These studies suggest that the redox-regulatory capacity of Nrf2 may be context-dependent, varying across disease stages, tissue compartments, and types of injury. For instance, reductions in Nrf2 and its targets have been observed in the DRG during chemotherapeutic neuropathies, whereas spinal upregulation is typical of constriction injuries [258,259].
In addition to neurons, Nrf2 also regulates the activity of glial cells, especially astrocytes and microglia, reducing their role in chronic pain by inhibiting the release of pro-inflammatory cytokines and reactive gliosis [99]. In this way, Nrf2 not only prevents oxidative and inflammatory damage to neurons but also orchestrates a broader neuroimmune regulatory network involved in pain amplification.

6.3. Nrf2 Signaling in Inflammation and Neuropathic Pain Modulation

The transcription factor Nrf2 has emerged as a pivotal modulator of inflammatory processes and a promising therapeutic target in neuropathic pain [99]. Its activation confers antinociceptive effects by downregulating pro-inflammatory signaling and enhancing the cellular antioxidant response [249]. In neuropathic pain models, Nrf2 activation suppresses the production of inflammatory mediators such as TNF-α and IL-1β, dampens microglial activation, and mitigates oxidative stress—factors that are central to the initiation and maintenance of pain signaling [238,239]. By attenuating the inflammatory milieu and oxidative burden within injured tissues, Nrf2 diminishes nociceptor excitability and reduces the transmission of pain impulses, while simultaneously preserving neuronal integrity and function.
Oxidative stress and inflammation are tightly interlinked through feedback amplification loops that sustain pathological pain states. For instance, reactive oxygen species promote NF-κB-mediated transcription of pro-inflammatory cytokines, which in turn exacerbate ROS generation 95]. Nrf2 activation interrupts this cycle, both by upregulating antioxidant enzymes and by inhibiting NF-κB-driven transcription [238,239]. Furthermore, Nrf2 negatively regulates the expression of cell adhesion molecules (CAMs), including vascular cell adhesion molecule 1 (VCAM-1) and E-selectin, partly through its downstream effector HO-1 [63]. This action reduces leukocyte recruitment and extravasation into sites of neural injury, thus curbing local inflammation. In addition, Nrf2 modulates the activity of matrix metalloproteinases (MMPs) and suppresses the expression of iNOS and COX-2, further dampening the inflammatory cascade [260,261].
A growing body of experimental evidence supports the therapeutic potential of targeting Nrf2 in neuropathic pain. Pharmacologic Nrf2 activators such as bardoxolone methyl have been shown to alleviate chemotherapy-induced neuropathy via suppression of inflammatory markers in the dorsal root ganglia [262]. Similarly, omaveloxolone improved mechanical allodynia and reduced neuronal apoptosis and glial activation in chronic constriction injury (CCI) models [263]. Betulinic acid and luteolin, natural compounds with known antioxidant properties, alleviated neuropathic pain by enhancing Nrf2/HO-1 signaling, suppressing glial-mediated inflammation, and modulating cytokine expression in both peripheral and central nervous tissues [264,265].

6.4. Nrf2–mTOR Axis at the Intersection of Oxidative Stress, Cellular Senescence, and Neuropathic Pain

Oxidative stress plays a central role in both initiating and perpetuating cellular senescence, acting upstream of SASP and shaping the inflammatory milieu that characterizes age-related pathologies and chronic pain syndromes [96]. Accumulated ROS trigger persistent DNA damage responses and activate redox-sensitive transcription factors—particularly NF-κB and C/EBPβ—that promote the expression of SASP components, including TNF-α, IL-1β, and IFN-γ [161,266]. These cytokines amplify oxidative stress in adjacent cells, forming a self-reinforcing loop that sustains chronic inflammation and senescence across tissues, a process intimately tied to the development and persistence of neuropathic pain.
Notably, recent findings have elucidated a reciprocal regulatory relationship between Nrf2 and mTOR, forming a redox-sensitive signaling hub that governs both metabolic homeostasis and senescence maintenance [233]. Nrf2 enhances mTORC1 activity through transcriptional upregulation of mTOR and RagD, a lysosome-localized GTPase essential for nutrient-sensing and mTORC1 activation [240]. Conversely, mTOR stabilizes Nrf2 by suppressing its β-TrCP–mediated ubiquitination and degradation, a Keap1-independent pathway that further ensures sustained antioxidant gene expression [240,267]. This mutual reinforcement between Nrf2 and mTOR promotes a prolonged antioxidant state that, while cytoprotective, may paradoxically support senescent cell survival and chronic low-grade inflammation, especially under conditions of persistent mTORC1 activation typical in aging tissues [268].

6.5. Nrf2–HCAR2 Crosstalk in Neuropathic Pain

The antioxidant transcription factor Nrf2 is analgesic in rodent neuropathic pain models, and oral dimethyl fumarate (DMF) reverses hypersensitivity via Nrf2 signaling [269]. Concurrently, the ketone-body receptor HCAR2 —activated by β-hydroxybutyrate and fumarate esters—has emerged as a modulator of neuroinflammation and pain circuitry [269,270]. Monomethyl fumarate (MMF), the active metabolite of DMF, directly engages HCAR2 on microglia/macrophages to dampen inflammatory programs, linking fumarates to this GPCR beyond their canonical Nrf2 effects [271,272]. Genetic and pharmacologic studies indicate that DMF/MMF’s protective actions in neuroinflammation require HCAR2, highlighting a dual mechanism wherein Nrf2 activation intersects with HCAR2 signaling [273].
In neuropathic pain paradigms, activating HCAR2 (via DMF or ketone pathways) alleviates allodynia and hyperalgesia, and these effects are lost in HCAR2-deficient animals, underscoring receptor involvement in nociceptive control [274,275]. Medicinal chemistry and recent reviews further position HCAR2 as a druggable anti-inflammatory/analgesic target, reinforcing the rationale for therapies that co-opt both Nrf2 and HCAR2 axes [276,277].
In the context of neuropathic pain, this crosstalk becomes pathophysiologically relevant. Excess ROS not only exacerbate mitochondrial dysfunction and glial activation but also sustain SASP-driven inflammation in dorsal root ganglia and spinal cord neurons. Nrf2 activation counters this by restoring redox balance, suppressing pro-inflammatory cytokine production, and modulating mTOR-driven metabolic reprogramming. However, in certain contexts, prolonged Nrf2–mTOR co-activation may stabilize senescent-like phenotypes in neural and glial cells, potentially contributing to pain chronification. To synthesize these interactions, we provide a concise overview in Table 4.

7. Conclusions

Neuropathic pain is sustained by an interconnected network of oxidative stress, neuroinflammation, mitochondrial dysfunction, and cellular senescence, in which nerve injury triggers immune activation, ion channel remodeling, and glial reactivity, while persistent redox imbalance amplifies excitability and cytokine-driven inflammation. Mitochondrial impairment fuels reactive species production and energy failure, and post-mitotic senescence of neurons and glia maintains a pro-inflammatory milieu through the senescence-associated secretory phenotype. The transcription factor Nrf2 acts as a central regulator by restoring redox homeostasis, modulating inflammation, preserving mitochondrial function, and counteracting senescence-associated signaling, with experimental evidence showing its activation alleviates neuropathic pain. Given its context-dependent effects, optimal management will require stage-specific, mechanism-based strategies that integrate redox modulation, mitochondrial support, senescence control, and neuroimmune regulation to achieve durable pain relief and improved functional outcomes.

8. Future Perspectives

Despite substantial advances in understanding the interplay between oxidative stress, inflammation, mitochondrial dysfunction, and cellular senescence in neuropathic pain, several critical knowledge gaps remain. First, the precise temporal dynamics and tissue specificity of these processes are incompletely mapped, limiting the ability to identify optimal therapeutic windows. The cellular sources of pathogenic signals—particularly the relative contributions of senescent nociceptors versus glial cells—require clearer delineation. Furthermore, the heterogeneity of neuropathic pain across etiologies and patient populations complicates the translation of preclinical findings into broadly effective interventions. Biomarkers capable of non-invasively detecting redox imbalance, senescence activity, and mitochondrial dysfunction in vivo are still underdeveloped, hindering patient stratification and real-time monitoring of treatment responses.
Promising therapeutic avenues include selective modulation of Nrf2 signaling to restore redox homeostasis while avoiding sustained overactivation that may stabilize senescent phenotypes. Targeted senotherapeutics—either senolytics aimed at clearing maladaptive senescent cells or senomorphics designed to suppress the senescence-associated secretory phenotype (SASP)—offer potential to dismantle the inflammatory–excitatory loop sustaining chronic pain. Mitochondria-directed interventions, such as matrix-targeted antioxidants, quality-control enhancers, and bioenergetic modulators, represent another mechanistically grounded strategy. Combination regimens pairing redox modulation with anti-inflammatory, glia-targeted, or senescence-modifying agents may yield synergistic benefits, particularly when tailored to disease stage and cellular targets.
However, several barriers impede clinical translation. Many preclinical models fail to capture the complexity and chronicity of human neuropathic pain, underscoring the need for predictive and clinically relevant experimental systems. The lack of standardized clinical endpoints for redox- and senescence-targeted therapies complicates trial design and cross-study comparison. Additionally, tissue-specific drug delivery, off-target effects, and compensatory biological responses remain major challenges for sustained efficacy. Overcoming these obstacles will require integrated, multidisciplinary efforts that combine mechanistic discovery with biomarker development, advanced modeling, and carefully designed early-phase clinical studies.

9. Materials and Methods

A systematic literature search was conducted in PubMed, Scopus, and Web of Science to identify peer-reviewed publications examining the interplay of oxidative stress, inflammation, mitochondrial dysfunction, cellular senescence (including SASP), and Nrf2 signaling in the pathophysiology of neuropathic pain. The search covered records from database inception to 1 September 2025.
Search strategies combined Medical Subject Headings (MeSH) and free-text keywords. Representative terms and Boolean logic included: “neuropathic pain” OR allodynia OR hyperalgesia AND (“oxidative stress” OR ROS OR RNS OR “mitochondrial dysfunction” OR mitochondria OR NOX OR “nitric oxide synthase” OR peroxynitrite) AND (senescence OR “cellular senescence” OR SASP OR “senescence-associated” OR Nrf2 OR NRF2 OR Keap1) AND (glia OR microglia OR astrocyte OR “central sensitization” OR TRPA1 OR TRPV1). Database-specific thesauri (e.g., MeSH/Emtree) were mapped to these concepts to optimize sensitivity and specificity. Reference lists of eligible articles and relevant reviews were also hand-searched to capture additional records.
  • Eligibility criteria
We included: (a) original experimental studies in vitro and in vivo (animal models of neuropathic pain), (b) clinical studies in humans (observational or interventional) directly probing oxidative/redox, mitochondrial, inflammatory–glial, senescence/SASP, or Nrf2-Keap1 pathways in neuropathic pain mechanisms, and (c) systematic reviews/meta-analyses that synthesized such mechanistic evidence. Studies were required to investigate at least one mechanistic axis relevant to this review (e.g., NOX/iNOS–derived oxidants, mitochondrial ROS, TRP/NaV redox gating, microglia/astrocyte-driven neuroinflammation, markers of cellular senescence, SASP mediators, or Nrf2-linked antioxidant responses).
We excluded: articles not focused on neuropathic pain; studies centered exclusively on nociception without neuropathic context; papers lacking an oxidative/mitochondrial, inflammatory–glial, senescence/SASP, or Nrf2 component; non-original content (editorials, letters, opinions), conference abstracts without full data; non-English publications; and duplicates or superseded datasets. When multiple reports drew on overlapping cohorts or experiments, the most comprehensive or most recent article was retained.
  • Screening process
Records from all databases were exported and deduplicated. Two reviewers independently screened titles/abstracts, then assessed full texts; disagreements were resolved by discussion with a third reviewer. Data extraction captured model/system, neuropathic-pain etiology, molecular targets (e.g., Nrf2, NF-κB, MAPK, COX-2, mitochondrial pathways, DDR, p16/p21, SASP), cell types (neurons, microglia/astrocytes, immune cells), assays/readouts (oxidative markers, cytokines, electrophysiology/behavior), and directionality of effects. Given heterogeneity, synthesis was thematic across (1) mitochondrial/redox injury and DDR; (2) glial-driven neuroinflammation; (3) senescence/SASP circuits; and (4) bidirectional crosstalk sustaining chronic neuropathic pain.
  • Study selection and PRISMA flow.
The search retrieved 1876 records. After removing 426 duplicates, 1450 records remained for title/abstract screening. 1104 were excluded as off-topic or not meeting inclusion criteria. 346 full-text articles were assessed; 198 were excluded with reasons: 92 lacked mechanistic linkage between the predefined axes, 54 addressed non-neuropathic pain, 28 were non-English or non-eligible article types, and 24 duplicated data. 148 studies met all criteria and were included in the qualitative synthesis. The selection process is depicted in Figure 7 (PRISMA 2020).

Author Contributions

Conceptualization, B.S., I.M.B. and B.S.S.; methodology, M.D.S., I.S. and A.L.; investigation, B.S., I.M.B., M.D.S. and J.N.; validation, B.S.S., T.L. and A.A.A.; resources, M.D.S., B.S.S. and M.M.; writing—original draft preparation, B.S., I.M.B., M.P. (Marko Petrovic), N.M., S.E., M.D.S., J.N. and B.S.S.; writing—review and editing, T.L., D.J., M.M., M.P. (Miodrag Peulic) and N.Z.; visualization, B.S. and M.S.; supervision, T.L., D.J. and N.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

As this is a review article, no new data were created or analyzed. All data supporting the findings of this study are available in the cited literature. Therefore, a data availability statement is not applicable.

Acknowledgments

The authors sincerely thank the Faculty of Medical Sciences, University of Kragujevac, Serbia, for their continuous support.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
4-HNE4-hydroxynonenal
53BP1p53-binding protein 1
AMPKAMP-activated protein kinase
AREsantioxidant response elements
ATMataxia-telangiectasia mutated
ATPadenosine triphosphate
ATPSc-KMTATP synthase c-subunit lysine N-methyltransferase
ATRataxia-telangiectasia and Rad3-related
BCL-2B-cell lymphoma 2
BDNFbrain-derived neurotrophic factor
BrdUbromodeoxyuridine
C/EBPβCCAAT/enhancer-binding protein beta
CCIchronic constriction injury
CCR2C-C chemokine receptor type 2
CCL11C-C motif chemokine ligand 11
CCL2C-C motif chemokine ligand 2
CGRPcalcitonin gene-related peptide
CHK1checkpoint kinase 1
CHK2checkpoint kinase 2
COcarbon monoxide
CAMscell adhesion molecules
CDKscyclin-dependent kinases
COX-2cyclo-oxygenase-2
CSF1colony-stimulating factor 1
CX3CL1C-X3-C motif chemokine ligand 1
CX3CL1–CX3CR1C-X3-C motif chemokine ligand 1–C-X3-C chemokine receptor 1
CXCL1C-X-C motif chemokine ligand 1
DAMPsdanger-associated molecular patterns
DDRDNA damage response
DRGdorsal root ganglia
DRP1dynamin-related protein 1
Duoxdual oxidases
E2FE2 promoter binding factor
EGFepidermal growth factor
EPE-type prostanoid
ERendoplasmic reticulum
ERKextracellular signal-regulated kinase
ERO1–PDIendoplasmic reticulum oxidoreductin 1–protein disulfide isomerase
ETCelectron transport chain
Fe2+ferrous iron
GABAgamma-aminobutyric acid
GADD45Agrowth arrest and DNA damage-inducible alpha
GDF15growth differentiation factor 15
GPR37G protein-coupled receptor 37
GRB2growth factor receptor-bound protein 2
GSHglutathione
GluN1glutamate ionotropic receptor NMDA type subunit 1
GM-CSFgranulocyte-macrophage colony-stimulating factor
gp130glycoprotein 130
H2O2hydrogen peroxide
HCNhyperpolarization-activated cyclic nucleotide-gated
HMGB1high-mobility group box 1
HO-1heme oxygenase-1
HuRhuman antigen R
IASPInternational Association for the Study of Pain
IFN-γinterferon-gamma
IGF-1insulin-like growth factor 1
ILinterleukin
IL-1βinterleukin-1β
IL-4interleukin-4
IL-10interleukin-10
IL-17interleukin-17
IL-33interleukin-33
iNOSinducible nitric oxide synthase
IP3Rinositol 1,4,5-trisphosphate receptor
IRFinterferon regulatory factor
JAK–STATJanus kinase–signal transducer and activator of transcription
JNKc-Jun N-terminal kinase
K+potassium
KCC2potassium-chloride cotransporter 2
Kvvoltage-gated potassium
LCN2lipocalin-2
LPSlipopolysaccharide
MAMsmitochondria-associated membranes
M-CSFmacrophage colony-stimulating factor
MAPKmitogen-activated protein kinase
MDC1mediator of DNA damage checkpoint protein 1
MFN1/2mitofusin-1/2
MiDASMitochondrial dysfunction–associated senescence
MMPsmatrix metalloproteinases
mPTPmitochondrial permeability transition pore
mtDNAMitochondrial DNA
mTORmechanistic target of rapamycin
mTORC1mechanistic target of rapamycin complex 1
MyD88myeloid differentiation primary response 88
NAD+nicotinamide adenine dinucleotide
NCLXsodium/calcium exchanger
NF-κBnuclear factor kappa-light-chain-enhancer of activated B cells
NGFnerve growth factor
NLRP3NOD-, LRR-, and pyrin domain-containing protein 3
NMDAN-methyl-D-aspartate
NOnitric oxide
NOX2NADPH oxidase 2
NPD1neuroprotectin D1
NRF1/2nuclear respiratory factors 1 and 2
Nrf2nuclear factor erythroid 2–related factor 2
NQO1NAD(P)H quinone oxidoreductase 1
O2superoxide anion
ONOOperoxynitrite
OPA1optic atrophy protein 1
P2X4P2X purinoceptor 4
P2X7P2X purinoceptor 7
PCNAproliferating cell nuclear antigen
PGC-1αperoxisome proliferator-activated receptor gamma coactivator 1-alpha
PGE2prostaglandin E2
PI3Kphosphoinositide 3-kinase
PINK1PTEN-induced kinase 1
PoMiCSpost-mitotic cellular senescence
PRRsPattern-recognition receptors
p53protein p53
RAGEreceptor for advanced glycation end products
Rbretinoblastoma
Ref-1redox factor-1
ROS/RNSreactive oxygen and nitrogen species
SA-β-galsenescence-associated β-galactosidase
SAHFsenescence-associated heterochromatic foci
SASPsenescence-associated secretory phenotype
SOD2superoxide dismutase 2
ST2suppression of tumorigenicity 2
TGF-βtransforming growth factor-beta
TFAMmitochondrial transcription factor A
thermo-TRPsThermo-transient receptor potential channels
ThT-helper
TLR4Toll-like receptor 4
TNF-αtumor necrosis factor-alpha
TrkBtropomyosin receptor kinase B
TRIFTIR-domain-containing adapter-inducing interferon-β
TRPA1transient receptor potential ankyrin 1
TRPV1transient receptor potential vanilloid 1
TRPV4transient receptor potential vanilloid 4
TSP1thrombospondin-1
UVUltraviolet
VCAM-1vascular cell adhesion molecule 1
WNTWingless/Integrated
cAMP–PKA/PKCcyclic adenosine monophosphate–protein kinase A/protein kinase C
cGAS–STINGcyclic GMP–AMP synthase–stimulator of interferon genes
NaVvoltage-gated sodium channels

References

  1. Cao, B.; Xu, Q.; Shi, Y.; Zhao, R.; Li, H.; Zheng, J.; Liu, F.; Wan, Y.; Wei, B. Pathology of pain and its implications for therapeutic interventions. Signal Transduct. Target. Ther. 2024, 9, 155. [Google Scholar] [CrossRef]
  2. De Ridder, D.; Adhia, D.; Vanneste, S. The anatomy of pain and suffering in the brain and its clinical implications. Neurosci. Biobehav. Rev. 2021, 130, 125–146. [Google Scholar] [CrossRef]
  3. Khera, T.; Rangasamy, V. Cognition and Pain: A Review. Front. Psychol. 2021, 12, 673962. [Google Scholar] [CrossRef]
  4. Raja, S.N.; Carr, D.B.; Cohen, M.; Finnerup, N.B.; Flor, H.; Gibson, S.; Keefe, F.J.; Mogil, J.S.; Ringkamp, M.; Sluka, K.A.; et al. The revised International Association for the Study of Pain definition of pain: Concepts, challenges, and compromises. PAIN 2020, 161, 1976–1982. [Google Scholar] [CrossRef]
  5. García-Domínguez, M. Injury-Driven Structural and Molecular Modifications in Nociceptors. Biology 2025, 14, 788. [Google Scholar] [CrossRef]
  6. Yam, M.F.; Loh, Y.C.; Tan, C.S.; Khadijah Adam, S.; Abdul Manan, N.; Basir, R. General Pathways of Pain Sensation and the Major Neurotransmitters Involved in Pain Regulation. Int. J. Mol. Sci. 2018, 19, 2164. [Google Scholar] [CrossRef] [PubMed]
  7. Karcz, M.; Abd-Elsayed, A.; Chakravarthy, K.; Aman, M.M.; Strand, N.; Malinowski, M.N.; Latif, U.; Dickerson, D.; Suvar, T.; Lubenow, T.; et al. Pathophysiology of Pain and Mechanisms of Neuromodulation: A Narrative Review (A Neuron Project). J. Pain Res. 2024, 17, 3757–3790. [Google Scholar] [CrossRef] [PubMed]
  8. Suzuki, K.; Suzuki, S.; Shiina, T.; Kobayashi, S.; Hirata, K. Central Sensitization in Migraine: A Narrative Review. J. Pain Res. 2022, 15, 2673–2682. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, J.A.; Yu, J.; Cheung, C.W. Immune Actions on the Peripheral Nervous System in Pain. Int. J. Mol. Sci. 2021, 22, 1448. [Google Scholar] [CrossRef]
  10. Hamdan, A.; Galvez, R.; Katati, M. Shedding light on neuropathic pain: Current and emerging tools for diagnosis, screening, and quantification. SAGE Open Med. 2024, 12, 20503121231218985. [Google Scholar] [CrossRef]
  11. Finnerup, N.B.; Haroutounian, S.; Kamerman, P.; Baron, R.; Bennett, D.L.H.; Bouhassira, D.; Cruccu, G.; Freeman, R.; Hansson, P.; Nurmikko, T.; et al. Neuropathic pain: An updated grading system for research and clinical practice. PAIN 2016, 157, 1599–1606. [Google Scholar] [CrossRef]
  12. Freynhagen, R.; Parada, H.A.; Calderon-Ospina, C.A.; Chen, J.; Rakhmawati Emril, D.; Fernández-Villacorta, F.J.; Franco, H.; Ho, K.Y.; Lara-Solares, A.; Li, C.C.; et al. Current understanding of the mixed pain concept: A brief narrative review. Curr. Med. Res. Opin. 2019, 35, 1011–1018. [Google Scholar] [CrossRef]
  13. Siracusa, R.; Paola, R.D.; Cuzzocrea, S.; Impellizzeri, D. Fibromyalgia: Pathogenesis, Mechanisms, Diagnosis and Treatment Options Update. Int. J. Mol. Sci. 2021, 22, 3891. [Google Scholar] [CrossRef]
  14. Attal, N.; Bouhassira, D.; Colvin, L. Advances and challenges in neuropathic pain: A narrative review and future directions. Br. J. Anaesth. 2023, 131, 79–92. [Google Scholar] [CrossRef] [PubMed]
  15. Jensen, T.S.; Baron, R.; Haanpää, M.; Kalso, E.; Loeser, J.D.; Rice, A.S.C.; Treede, R.-D. A new definition of neuropathic pain. PAIN 2011, 152, 2204–2205. [Google Scholar] [CrossRef] [PubMed]
  16. Jaffal, S.M. Neuroplasticity in chronic pain: Insights into diagnosis and treatment. Korean J. Pain 2025, 38, 89–102. [Google Scholar] [CrossRef]
  17. Costigan, M.; Scholz, J.; Woolf, C.J. Neuropathic pain: A maladaptive response of the nervous system to damage. Annu. Rev. Neurosci. 2009, 32, 1–32. [Google Scholar] [CrossRef] [PubMed]
  18. Finnerup, N.B.; Kuner, R.; Jensen, T.S. Neuropathic Pain: From Mechanisms to Treatment. Physiol. Rev. 2021, 101, 259–301. [Google Scholar] [CrossRef]
  19. Rugnath, R.; Orzechowicz, C.; Newell, C.; Carullo, V.; Rugnath, A. A Literature Review: The Mechanisms and Treatment of Neuropathic Pain—A Brief Discussion. Biomedicines 2024, 12, 204. [Google Scholar] [CrossRef]
  20. Stucky, C.L.; Mikesell, A.R. Cutaneous pain in disorders affecting peripheral nerves. Neurosci. Lett. 2021, 765, 136233. [Google Scholar] [CrossRef]
  21. von Hehn, C.A.; Baron, R.; Woolf, C.J. Deconstructing the neuropathic pain phenotype to reveal neural mechanisms. Neuron 2012, 73, 638–652. [Google Scholar] [CrossRef] [PubMed]
  22. Blyth, F.M. Global burden of neuropathic pain. PAIN 2018, 159, 614–617. [Google Scholar] [CrossRef]
  23. Cavalli, E.; Mammana, S.; Nicoletti, F.; Bramanti, P.; Mazzon, E. The neuropathic pain: An overview of the current treatment and future therapeutic approaches. Int. J. Immunopathol. Pharmacol. 2019, 33, 2058738419838383. [Google Scholar] [CrossRef]
  24. Doneddu, P.E.; Pensato, U.; Iorfida, A.; Alberti, C.; Nobile-Orazio, E.; Fabbri, A.; Voza, A. Neuropathic Pain in the Emergency Setting: Diagnosis and Management. J. Clin. Med. 2023, 12, 6028. [Google Scholar] [CrossRef]
  25. Baskozos, G.; Hébert, H.L.; Pascal, M.M.; Themistocleous, A.C.; Macfarlane, G.J.; Wynick, D.; Bennett, D.L.; Smith, B.H. Epidemiology of neuropathic pain: An analysis of prevalence and associated factors in UK Biobank. Pain Rep. 2023, 8, e1066. [Google Scholar] [CrossRef]
  26. Sawicki, C.M.; Humeidan, M.L.; Sheridan, J.F. Neuroimmune Interactions in Pain and Stress: An Interdisciplinary Approach. Neuroscientist 2021, 27, 113–128. [Google Scholar] [CrossRef] [PubMed]
  27. Malcangio, M. Role of the immune system in neuropathic pain. Scand. J. Pain 2019, 20, 33–37. [Google Scholar] [CrossRef]
  28. Yang, J.-X.; Wang, H.-F.; Chen, J.-Z.; Li, H.-Y.; Hu, J.-C.; Yu, A.-A.; Wen, J.-J.; Chen, S.-J.; Lai, W.-D.; Wang, S.; et al. Potential Neuroimmune Interaction in Chronic Pain: A Review on Immune Cells in Peripheral and Central Sensitization. Front. Pain Res. 2022, 3, 946846. [Google Scholar] [CrossRef]
  29. Boakye, P.A.; Tang, S.-J.; Smith, P.A. Mediators of Neuropathic Pain; Focus on Spinal Microglia, CSF-1, BDNF, CCL21, TNF-α, Wnt Ligands, and Interleukin 1β. Front. Pain Res. 2021, 2, 698157. [Google Scholar] [CrossRef]
  30. Fiorin, F.d.S.; do Espírito Santo, C.C. The Potential Roles of Astrocytes and Microglia in the Spinal Cord and Brain After Spinal Cord Injury. Neuroglia 2025, 6, 12. [Google Scholar] [CrossRef]
  31. Kanda, H.; Kobayashi, K.; Yamanaka, H.; Okubo, M.; Dai, Y.; Noguchi, K. Localization of prostaglandin E2 synthases and E-prostanoid receptors in the spinal cord in a rat model of neuropathic pain. Brain Res. 2021, 1750, 147153. [Google Scholar] [CrossRef]
  32. Fang, X.X.; Zhai, M.N.; Zhu, M.; He, C.; Wang, H.; Wang, J.; Zhang, Z.J. Inflammation in pathogenesis of chronic pain: Foe and friend. Mol. Pain 2023, 19, 17448069231178176. [Google Scholar] [CrossRef]
  33. Seifert, O.; Baerwald, C. Interaction of pain and chronic inflammation. Z. Rheumatol. 2021, 80, 205–213. [Google Scholar] [CrossRef]
  34. Gu, D.; Xia, Y.; Ding, Z.; Qian, J.; Gu, X.; Bai, H.; Jiang, M.; Yao, D. Inflammation in the Peripheral Nervous System after Injury. Biomedicines 2024, 12, 1256. [Google Scholar] [CrossRef] [PubMed]
  35. Ronchetti, S.; Migliorati, G.; Delfino, D.V. Association of inflammatory mediators with pain perception. Biomed. Pharmacother. 2017, 96, 1445–1452. [Google Scholar] [CrossRef] [PubMed]
  36. Dou, X.; Chen, R.; Yang, J.; Dai, M.; Long, J.; Sun, S.; Lin, Y. The potential role of T-cell metabolism-related molecules in chronic neuropathic pain after nerve injury: A narrative review. Front. Immunol. 2023, 14, 1107298. [Google Scholar] [CrossRef]
  37. Catalisano, G.; Campione, G.M.; Spurio, G.; Galvano, A.N.; di Villalba, C.P.; Giarratano, A.; Alongi, A.; Ippolito, M.; Cortegiani, A. Neuropathic pain, antidepressant drugs, and inflammation: A narrative review. J. Anesth. Analg. Crit. Care 2024, 4, 67. [Google Scholar] [CrossRef] [PubMed]
  38. Alles, S.R.A.; Smith, P.A. Peripheral Voltage-Gated Cation Channels in Neuropathic Pain and Their Potential as Therapeutic Targets. Front. Pain Res. 2021, 2, 750583. [Google Scholar] [CrossRef]
  39. Vay, L.; Gu, C.; McNaughton, P.A. The thermo-TRP ion channel family: Properties and therapeutic implications. Br. J. Pharmacol. 2012, 165, 787–801. [Google Scholar] [CrossRef]
  40. Mickle, A.D.; Shepherd, A.J.; Mohapatra, D.P. Nociceptive TRP Channels: Sensory Detectors and Transducers in Multiple Pain Pathologies. Pharmaceuticals 2016, 9, 72. [Google Scholar] [CrossRef]
  41. Zhang, T.; Zhang, M.; Cui, S.; Liang, W.; Jia, Z.; Guo, F.; Ou, W.; Wu, Y.; Zhang, S. The core of maintaining neuropathic pain: Crosstalk between glial cells and neurons (neural cell crosstalk at spinal cord). Brain Behav. 2023, 13, e2868. [Google Scholar] [CrossRef]
  42. Karavis, M.Y.; Siafaka, I.; Vadalouca, A.; Georgoudis, G. Role of Microglia in Neuropathic Pain. Cureus 2023, 15, e43555. [Google Scholar] [CrossRef] [PubMed]
  43. Lu, H.J.; Gao, Y.J. Astrocytes in Chronic Pain: Cellular and Molecular Mechanisms. Neurosci. Bull. 2023, 39, 425–439. [Google Scholar] [CrossRef]
  44. Vergne-Salle, P.; Bertin, P. Chronic pain and neuroinflammation. Jt. Bone Spine 2021, 88, 105222. [Google Scholar] [CrossRef]
  45. Shehab, S.; Hamad, M.I.K.; Emerald, B.S. A novel approach to completely alleviate peripheral neuropathic pain in human patients: Insights from preclinical data. Front. Neuroanat. 2025, 18, 1523095. [Google Scholar] [CrossRef] [PubMed]
  46. Gilron, I.; Baron, R.; Jensen, T. Neuropathic Pain: Principles of Diagnosis and Treatment. Mayo Clin. Proc. 2015, 90, 532–545. [Google Scholar] [CrossRef] [PubMed]
  47. Yeh, T.-Y.; Luo, I.-W.; Hsieh, Y.-L.; Tseng, T.-J.; Chiang, H.; Hsieh, S.-T. Peripheral Neuropathic Pain: From Experimental Models to Potential Therapeutic Targets in Dorsal Root Ganglion Neurons. Cells 2020, 9, 2725. [Google Scholar] [CrossRef]
  48. Cuevas-Diaz Duran, R.; Li, Y.; Garza Carbajal, A.; You, Y.; Dessauer, C.W.; Wu, J.; Walters, E.T. Major Differences in Transcriptional Alterations in Dorsal Root Ganglia Between Spinal Cord Injury and Peripheral Neuropathic Pain Models. J. Neurotrauma 2023, 40, 883–900. [Google Scholar] [CrossRef]
  49. Tong, S.H.; Liu, D.L.; Liao, P.; Zhang, S.Y.; Zhou, J.; Zong, Y.; Zhang, C.Q.; Huang, Y.G.; Gao, J.J. Emerging role of macrophages in neuropathic pain. J. Orthop. Transl. 2025, 51, 227–241. [Google Scholar] [CrossRef]
  50. Kocot-Kępska, M.; Zajączkowska, R.; Mika, J.; Wordliczek, J.; Dobrogowski, J.; Przeklasa-Muszyńska, A. Peripheral Mechanisms of Neuropathic Pain-the Role of Neuronal and Non-Neuronal Interactions and Their Implications for Topical Treatment of Neuropathic Pain. Pharmaceuticals 2021, 14, 77. [Google Scholar] [CrossRef]
  51. Liu, X.; Yang, W.; Zhu, C.; Sun, S.; Wu, S.; Wang, L.; Wang, Y.; Ge, Z. Toll-like receptors and their role in neuropathic pain and migraine. Mol. Brain 2022, 15, 73. [Google Scholar] [CrossRef]
  52. Zhang, Q.; Wang, J.; Duan, M.T.; Han, S.P.; Zeng, X.Y.; Wang, J.Y. NF-κB, ERK, p38 MAPK and JNK contribute to the initiation and/or maintenance of mechanical allodynia induced by tumor necrosis factor-alpha in the red nucleus. Brain Res. Bull. 2013, 99, 132–139. [Google Scholar] [CrossRef] [PubMed]
  53. Sommer, C.; Leinders, M.; Üçeyler, N. Inflammation in the pathophysiology of neuropathic pain. PAIN 2018, 159, 595–602. [Google Scholar] [CrossRef]
  54. Kuffler, D.P. Mechanisms for Reducing Neuropathic Pain. Mol. Neurobiol. 2020, 57, 67–87. [Google Scholar] [CrossRef]
  55. Pinho-Ribeiro, F.A.; Verri, W.A., Jr.; Chiu, I.M. Nociceptor Sensory Neuron-Immune Interactions in Pain and Inflammation. Trends Immunol. 2017, 38, 5–19. [Google Scholar] [CrossRef]
  56. Felix, R.; Corzo-Lopez, A.; Sandoval, A. Voltage-Gated Ion Channels in Neuropathic Pain Signaling. Life 2025, 15, 888. [Google Scholar] [CrossRef]
  57. García-Domínguez, M. The Role of TNF-α in Neuropathic Pain: An Immunotherapeutic Perspective. Life 2025, 15, 785. [Google Scholar] [CrossRef]
  58. Carrasco, C.; Naziroǧlu, M.; Rodríguez, A.B.; Pariente, J.A. Neuropathic Pain: Delving into the Oxidative Origin and the Possible Implication of Transient Receptor Potential Channels. Front. Physiol. 2018, 9, 95. [Google Scholar] [CrossRef]
  59. Wu, X.B.; Zhu, Q.; Gao, Y.J. CCL2/CCR2 Contributes to the Altered Excitatory-inhibitory Synaptic Balance in the Nucleus Accumbens Shell Following Peripheral Nerve Injury-induced Neuropathic Pain. Neurosci. Bull. 2021, 37, 921–933. [Google Scholar] [CrossRef]
  60. Jiang, B.C.; Liu, T.; Gao, Y.J. Chemokines in chronic pain: Cellular and molecular mechanisms and therapeutic potential. Pharmacol. Ther. 2020, 212, 107581. [Google Scholar] [CrossRef]
  61. Volcheck, M.M.; Graham, S.M.; Fleming, K.C.; Mohabbat, A.B.; Luedtke, C.A. Central sensitization, chronic pain, and other symptoms: Better understanding, better management. Clevel. Clin. J. Med. 2023, 90, 245–254. [Google Scholar] [CrossRef]
  62. Smith, P.A. Neuropathic pain; what we know and what we should do about it. Front. Pain Res. 2023, 4, 1220034. [Google Scholar] [CrossRef]
  63. Chen, G.; Zhang, Y.-Q.; Qadri, Y.J.; Serhan, C.N.; Ji, R.-R. Microglia in Pain: Detrimental and Protective Roles in Pathogenesis and Resolution of Pain. Neuron 2018, 100, 1292–1311. [Google Scholar] [CrossRef]
  64. Teixeira-Santos, L.; Veríssimo, E.; Martins, S.; Sousa, T.; Albino-Teixeira, A.; Pinho, D. Effects of NADPH Oxidase Isoform-2 (NOX2) Inhibition on Behavioral Responses and Neuroinflammation in a Mouse Model of Neuropathic Pain. Biomedicines 2023, 11, 416. [Google Scholar] [CrossRef] [PubMed]
  65. Hu, Z.; Yu, X.; Chen, P.; Jin, K.; Zhou, J.; Wang, G.; Yu, J.; Wu, T.; Wang, Y.; Lin, F.; et al. BDNF-TrkB signaling pathway-mediated microglial activation induces neuronal KCC2 downregulation contributing to dynamic allodynia following spared nerve injury. Mol. Pain 2023, 19, 17448069231185439. [Google Scholar] [CrossRef]
  66. Zhang, H.; Zhang, X.; Chai, Y.; Wang, Y.; Zhang, J.; Chen, X. Astrocyte-mediated inflammatory responses in traumatic brain injury: Mechanisms and potential interventions. Front. Immunol. 2025, 16, 1584577. [Google Scholar] [CrossRef]
  67. Grace, P.M.; Gaudet, A.D.; Staikopoulos, V.; Maier, S.F.; Hutchinson, M.R.; Salvemini, D.; Watkins, L.R. Nitroxidative Signaling Mechanisms in Pathological Pain. Trends Neurosci. 2016, 39, 862–879. [Google Scholar] [CrossRef] [PubMed]
  68. Nashtahosseini, Z.; Eslami, M.; Paraandavaji, E.; Haraj, A.; Dowlat, B.F.; Hosseinzadeh, E.; Oksenych, V.; Naderian, R. Cytokine Signaling in Diabetic Neuropathy: A Key Player in Peripheral Nerve Damage. Biomedicines 2025, 13, 589. [Google Scholar] [CrossRef]
  69. Gustin, A.; Kirchmeyer, M.; Koncina, E.; Felten, P.; Losciuto, S.; Heurtaux, T.; Tardivel, A.; Heuschling, P.; Dostert, C. NLRP3 Inflammasome Is Expressed and Functional in Mouse Brain Microglia but Not in Astrocytes. PLoS ONE 2015, 10, e0130624. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, X.; Zhang, Y.; Chen, Y.; Ji, Y.; Lyu, Y.; Miao, Z.; Duan, X.; Liu, X. Unraveling the immune system’s role in peripheral nerve regeneration: A pathway to enhanced healing. Front. Immunol. 2025, 16, 1540199. [Google Scholar] [CrossRef]
  71. Ruan, S.; Jia, R.; Hu, L.; Liu, Y.; Tian, Q.; Jiang, K.; Xia, X.; Tao, X.; Liu, W.T.; Pan, Y.; et al. Ozone promotes macrophage efferocytosis and alleviates neuropathic pain by activating the AMPK/Gas6-MerTK/SOCS3 signaling pathway. Front. Immunol. 2024, 15, 1455771. [Google Scholar] [CrossRef]
  72. Chen, O.; Luo, X.; Ji, R.R. Macrophages and microglia in inflammation and neuroinflammation underlying different pain states. Med. Rev. 2023, 3, 381–407. [Google Scholar] [CrossRef]
  73. Domoto, R.; Sekiguchi, F.; Tsubota, M.; Kawabata, A. Macrophage as a Peripheral Pain Regulator. Cells 2021, 10, 1881. [Google Scholar] [CrossRef]
  74. Matsuda, M.; Huh, Y.; Ji, R.R. Roles of inflammation, neurogenic inflammation, and neuroinflammation in pain. J. Anesth. 2019, 33, 131–139. [Google Scholar] [CrossRef]
  75. Zhu, X.; Cao, S.; Zhu, M.D.; Liu, J.Q.; Chen, J.J.; Gao, Y.J. Contribution of chemokine CCL2/CCR2 signaling in the dorsal root ganglion and spinal cord to the maintenance of neuropathic pain in a rat model of lumbar disc herniation. J. Pain 2014, 15, 516–526. [Google Scholar] [CrossRef] [PubMed]
  76. Baral, P.; Udit, S.; Chiu, I.M. Pain and immunity: Implications for host defence. Nat. Rev. Immunol. 2019, 19, 433–447. [Google Scholar] [CrossRef]
  77. Tsujita, R.; Tsubota, M.; Sekiguchi, F.; Kawabata, A. Role of high-mobility group box 1 and its modulation by thrombomodulin/thrombin axis in neuropathic and inflammatory pain. Br. J. Pharmacol. 2021, 178, 798–812. [Google Scholar] [CrossRef] [PubMed]
  78. Tang, W.; Li, X.; Liu, H.; Xu, C.; Deng, S. The role of macrophages in chronic pain. Cytokine 2025, 185, 156813. [Google Scholar] [CrossRef] [PubMed]
  79. Ji, R.R. Specialized Pro-Resolving Mediators as Resolution Pharmacology for the Control of Pain and Itch. Annu. Rev. Pharmacol. Toxicol. 2023, 63, 273–293. [Google Scholar] [CrossRef]
  80. Bang, S.; Xie, Y.K.; Zhang, Z.J.; Wang, Z.; Xu, Z.Z.; Ji, R.R. GPR37 regulates macrophage phagocytosis and resolution of inflammatory pain. J. Clin. Investig. 2018, 128, 3568–3582. [Google Scholar] [CrossRef]
  81. Ji, R.R.; Nackley, A.; Huh, Y.; Terrando, N.; Maixner, W. Neuroinflammation and Central Sensitization in Chronic and Widespread Pain. Anesthesiology 2018, 129, 343–366. [Google Scholar] [CrossRef]
  82. Zhang, W.J.; Luo, H.L.; Zhu, Z.M. The role of P2X4 receptors in chronic pain: A potential pharmacological target. Biomed. Pharmacother. 2020, 129, 110447. [Google Scholar] [CrossRef]
  83. Sun, S.; Jiang, W.; Yan, X.; Zhang, J.; Gao, L.; Wu, C.; Zhu, B.; Wu, L. Ligand-gated ion channel P2X7 regulates NLRP3/Caspase-1-mediated inflammatory pain caused by pulpitis in the trigeminal ganglion and medullary dorsal horn. Brain Res. Bull. 2023, 192, 1–10. [Google Scholar] [CrossRef]
  84. Furutani, K.; Ji, R.R. Targeting Hv1 proton channel for pain control. Cell Res. 2022, 32, 419–420. [Google Scholar] [CrossRef] [PubMed]
  85. Hu, X.; Du, L.; Liu, S.; Lan, Z.; Zang, K.; Feng, J.; Zhao, Y.; Yang, X.; Xie, Z.; Wang, P.L.; et al. A TRPV4-dependent neuroimmune axis in the spinal cord promotes neuropathic pain. J. Clin. Investig. 2023, 133, e161507. [Google Scholar] [CrossRef]
  86. Dhir, S.; Derue, H.; Ribeiro-da-Silva, A. Temporal changes of spinal microglia in murine models of neuropathic pain: A scoping review. Front. Immunol. 2024, 15, 1460072. [Google Scholar] [CrossRef] [PubMed]
  87. Paolicelli, R.C.; Sierra, A.; Stevens, B.; Tremblay, M.E.; Aguzzi, A.; Ajami, B.; Amit, I.; Audinat, E.; Bechmann, I.; Bennett, M.; et al. Microglia states and nomenclature: A field at its crossroads. Neuron 2022, 110, 3458–3483. [Google Scholar] [CrossRef]
  88. Clark, A.K.; Malcangio, M. Fractalkine/CX3CR1 signaling during neuropathic pain. Front. Cell. Neurosci. 2014, 8, 121. [Google Scholar] [CrossRef]
  89. Clark, A.K.; Yip, P.K.; Malcangio, M. The liberation of fractalkine in the dorsal horn requires microglial cathepsin S. J. Neurosci. 2009, 29, 6945–6954. [Google Scholar] [CrossRef] [PubMed]
  90. Clark, A.K.; Yip, P.K.; Grist, J.; Gentry, C.; Staniland, A.A.; Marchand, F.; Dehvari, M.; Wotherspoon, G.; Winter, J.; Ullah, J.; et al. Inhibition of spinal microglial cathepsin S for the reversal of neuropathic pain. Proc. Natl. Acad. Sci. USA 2007, 104, 10655–10660. [Google Scholar] [CrossRef]
  91. Luongo, L.; Sajic, M.; Grist, J.; Clark, A.K.; Maione, S.; Malcangio, M. Spinal changes associated with mechanical hypersensitivity in a model of Guillain-Barré syndrome. Neurosci. Lett. 2008, 437, 98–102. [Google Scholar] [CrossRef] [PubMed]
  92. Ji, R.R.; Donnelly, C.R.; Nedergaard, M. Astrocytes in chronic pain and itch. Nat. Rev. Neurosci. 2019, 20, 667–685. [Google Scholar] [CrossRef] [PubMed]
  93. Kim, W.; Angulo, M.C. Unraveling the role of oligodendrocytes and myelin in pain. J. Neurochem. 2025, 169, e16206. [Google Scholar] [CrossRef]
  94. Zarpelon, A.C.; Rodrigues, F.C.; Lopes, A.H.; Souza, G.R.; Carvalho, T.T.; Pinto, L.G.; Xu, D.; Ferreira, S.H.; Alves-Filho, J.C.; McInnes, I.B.; et al. Spinal cord oligodendrocyte-derived alarmin IL-33 mediates neuropathic pain. FASEB J. 2016, 30, 54–65. [Google Scholar] [CrossRef]
  95. Bellanti, F.; Coda, A.R.D.; Trecca, M.I.; Lo Buglio, A.; Serviddio, G.; Vendemiale, G. Redox Imbalance in Inflammation: The Interplay of Oxidative and Reductive Stress. Antioxidants 2025, 14, 656. [Google Scholar] [CrossRef] [PubMed]
  96. Stojanovic, B.; Jovanovic, I.; Dimitrijevic Stojanovic, M.; Stojanovic, B.S.; Kovacevic, V.; Radosavljevic, I.; Jovanovic, D.; Miletic Kovacevic, M.; Zornic, N.; Arsic, A.A.; et al. Oxidative Stress-Driven Cellular Senescence: Mechanistic Crosstalk and Therapeutic Horizons. Antioxidants 2025, 14, 987. [Google Scholar] [CrossRef]
  97. Liu, S.; Liu, J.; Wang, Y.; Deng, F.; Deng, Z. Oxidative Stress: Signaling Pathways, Biological Functions, and Disease. MedComm 2025, 6, e70268. [Google Scholar] [CrossRef]
  98. Collin, F. Chemical Basis of Reactive Oxygen Species Reactivity and Involvement in Neurodegenerative Diseases. Int. J. Mol. Sci. 2019, 20, 2407. [Google Scholar] [CrossRef]
  99. Petrikonis, K.; Bernatoniene, J.; Kopustinskiene, D.M.; Casale, R.; Davinelli, S.; Saso, L. The Antinociceptive Role of Nrf2 in Neuropathic Pain: From Mechanisms to Clinical Perspectives. Pharmaceutics 2024, 16, 1068. [Google Scholar] [CrossRef]
  100. Fujii, J.; Homma, T.; Osaki, T. Superoxide Radicals in the Execution of Cell Death. Antioxidants 2022, 11, 501. [Google Scholar] [CrossRef]
  101. Zhang, X.; Gao, Y.; Zhang, S.; Wang, Y.; Pei, X.; Chen, Y.; Zhang, J.; Zhang, Y.; Du, Y.; Hao, S.; et al. Mitochondrial dysfunction in the regulation of aging and aging-related diseases. Cell Commun. Signal. 2025, 23, 290. [Google Scholar] [CrossRef]
  102. Mercola, J. Reductive stress and mitochondrial dysfunction: The hidden link in chronic disease. Free Radic. Biol. Med. 2025, 233, 118–131. [Google Scholar] [CrossRef]
  103. Leto, T.L.; Morand, S.; Hurt, D.; Ueyama, T. Targeting and regulation of reactive oxygen species generation by Nox family NADPH oxidases. Antioxid. Redox Signal. 2009, 11, 2607–2619. [Google Scholar] [CrossRef]
  104. Jomova, K.; Raptova, R.; Alomar, S.Y.; Alwasel, S.H.; Nepovimova, E.; Kuca, K.; Valko, M. Reactive oxygen species, toxicity, oxidative stress, and antioxidants: Chronic diseases and aging. Arch. Toxicol. 2023, 97, 2499–2574. [Google Scholar] [CrossRef]
  105. Vermot, A.; Petit-Härtlein, I.; Smith, S.M.E.; Fieschi, F. NADPH Oxidases (NOX): An Overview from Discovery, Molecular Mechanisms to Physiology and Pathology. Antioxidants 2021, 10, 890. [Google Scholar] [CrossRef] [PubMed]
  106. Zilberter, Y.; Tabuena, D.R.; Zilberter, M. NOX-induced oxidative stress is a primary trigger of major neurodegenerative disorders. Prog. Neurobiol. 2023, 231, 102539. [Google Scholar] [CrossRef]
  107. Dash, U.C.; Bhol, N.K.; Swain, S.K.; Samal, R.R.; Nayak, P.K.; Raina, V.; Panda, S.K.; Kerry, R.G.; Duttaroy, A.K.; Jena, A.B. Oxidative stress and inflammation in the pathogenesis of neurological disorders: Mechanisms and implications. Acta Pharm. Sin. B 2025, 15, 15–34. [Google Scholar] [CrossRef]
  108. Pfeifer, G.P. Mechanisms of UV-induced mutations and skin cancer. Genome Instab. Dis. 2020, 1, 99–113. [Google Scholar] [CrossRef] [PubMed]
  109. Wright, D.; Zampagni, M.; Evangelisti, E.; Conti, S.; D’Adamio, G.; Goti, A.; Becatti, M.; Fiorillo, C.; Taddei, N.; Cecchi, C.; et al. Protective properties of novel S-acyl-glutathione thioesters against ultraviolet-induced oxidative stress. Photochem. Photobiol. 2013, 89, 442–452. [Google Scholar] [CrossRef] [PubMed]
  110. Azzam, E.I.; Jay-Gerin, J.P.; Pain, D. Ionizing radiation-induced metabolic oxidative stress and prolonged cell injury. Cancer Lett. 2012, 327, 48–60. [Google Scholar] [CrossRef]
  111. Yang, P.; Li, J.; Zhang, T.; Ren, Y.; Zhang, Q.; Liu, R.; Li, H.; Hua, J.; Wang, W.A.; Wang, J.; et al. Ionizing radiation-induced mitophagy promotes ferroptosis by increasing intracellular free fatty acids. Cell Death Differ. 2023, 30, 2432–2445. [Google Scholar] [CrossRef]
  112. Fraternale, A.; Zara, C.; De Angelis, M.; Nencioni, L.; Palamara, A.T.; Retini, M.; Di Mambro, T.; Magnani, M.; Crinelli, R. Intracellular Redox-Modulated Pathways as Targets for Effective Approaches in the Treatment of Viral Infection. Int. J. Mol. Sci. 2021, 22, 3603. [Google Scholar] [CrossRef]
  113. Oberacker, T.; Kraft, L.; Schanz, M.; Latus, J.; Schricker, S. The Importance of Thioredoxin-1 in Health and Disease. Antioxidants 2023, 12, 1078. [Google Scholar] [CrossRef]
  114. Mijit, M.; Kpenu, E.; Chowdhury, N.N.; Gampala, S.; Wireman, R.; Liu, S.; Babb, O.; Georgiadis, M.M.; Wan, J.; Fishel, M.L.; et al. In vitro and In vivo evidence demonstrating chronic absence of Ref-1 Cysteine 65 impacts Ref-1 folding configuration, redox signaling, proliferation and metastasis in pancreatic cancer. Redox Biol. 2024, 69, 102977. [Google Scholar] [CrossRef] [PubMed]
  115. Delmastro-Greenwood, M.; Freeman, B.A.; Wendell, S.G. Redox-dependent anti-inflammatory signaling actions of unsaturated fatty acids. Annu. Rev. Physiol. 2014, 76, 79–105. [Google Scholar] [CrossRef]
  116. Chandimali, N.; Bak, S.G.; Park, E.H.; Lim, H.-J.; Won, Y.-S.; Kim, E.-K.; Park, S.-I.; Lee, S.J. Free radicals and their impact on health and antioxidant defenses: A review. Cell Death Discov. 2025, 11, 19. [Google Scholar] [CrossRef] [PubMed]
  117. Kisby, G.E.; Wilson, D.M.; Spencer, P.S. Introducing the Role of Genotoxicity in Neurodegenerative Diseases and Neuropsychiatric Disorders. Int. J. Mol. Sci. 2024, 25, 7221. [Google Scholar] [CrossRef]
  118. Jové, M.; Mota-Martorell, N.; Obis, È.; Sol, J.; Martín-Garí, M.; Ferrer, I.; Portero-Otín, M.; Pamplona, R. Lipid Adaptations against Oxidative Challenge in the Healthy Adult Human Brain. Antioxidants 2023, 12, 177. [Google Scholar] [CrossRef]
  119. Jurcau, M.-C.; Jurcau, A.; Diaconu, R.-G. Oxidative Stress in the Pathogenesis of Neurodegenerative Diseases. Stresses 2024, 4, 827–849. [Google Scholar] [CrossRef]
  120. Ho, T.; Ryan, M.; Holle, J. Redox reactions in chronic pain: Mechanisms and relevance in fibromyalgia. Front. Pain Res. 2025, 6, 1593908. [Google Scholar] [CrossRef]
  121. Muller, F.L.; Liu, Y.; Van Remmen, H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. 2004, 279, 49064–49073. [Google Scholar] [CrossRef]
  122. Palma, F.R.; He, C.; Danes, J.M.; Paviani, V.; Coelho, D.R.; Gantner, B.N.; Bonini, M.G. Mitochondrial Superoxide Dismutase: What the Established, the Intriguing, and the Novel Reveal About a Key Cellular Redox Switch. Antioxid. Redox Signal. 2020, 32, 701–714. [Google Scholar] [CrossRef]
  123. Sil, R.; Chakraborti, A. Major heme proteins hemoglobin and myoglobin with respect to their roles in oxidative stress—A brief review. Front. Chem. 2025, 13, 1543455. [Google Scholar] [CrossRef]
  124. Eid, S.A.; Savelieff, M.G.; Eid, A.A.; Feldman, E.L. Nox, Nox, Are You There? The Role of NADPH Oxidases in the Peripheral Nervous System. Antioxid. Redox Signal. 2022, 37, 613–630. [Google Scholar] [CrossRef] [PubMed]
  125. Kallenborn-Gerhardt, W.; Schröder, K.; Schmidtko, A. NADPH Oxidases in Pain Processing. Antioxidants 2022, 11, 1162. [Google Scholar] [CrossRef] [PubMed]
  126. Cipriano, A.; Viviano, M.; Feoli, A.; Milite, C.; Sarno, G.; Castellano, S.; Sbardella, G. NADPH Oxidases: From Molecular Mechanisms to Current Inhibitors. J. Med. Chem. 2023, 66, 11632–11655. [Google Scholar] [CrossRef] [PubMed]
  127. Förstermann, U.; Sessa, W.C. Nitric oxide synthases: Regulation and function. Eur. Heart J. 2012, 33, 829–837. [Google Scholar] [CrossRef]
  128. Padalko, V.; Posnik, F.; Adamczyk, M. Mitochondrial Aconitase and Its Contribution to the Pathogenesis of Neurodegenerative Diseases. Int. J. Mol. Sci. 2024, 25, 9950. [Google Scholar] [CrossRef]
  129. Martemucci, G.; Costagliola, C.; Mariano, M.; D’andrea, L.; Napolitano, P.; D’Alessandro, A.G. Free Radical Properties, Source and Targets, Antioxidant Consumption and Health. Oxygen 2022, 2, 48–78. [Google Scholar] [CrossRef]
  130. Konno, T.; Melo, E.P.; Chambers, J.E.; Avezov, E. Intracellular Sources of ROS/H(2)O(2) in Health and Neurodegeneration: Spotlight on Endoplasmic Reticulum. Cells 2021, 10, 233. [Google Scholar] [CrossRef]
  131. Li, R.; Yan, X.; Zhao, Y.; Liu, H.; Wang, J.; Yuan, Y.; Li, Q.; Su, J. Oxidative Stress Induced by Nuclear Factor Erythroid 2-Related Factor 2 (NRF2) Dysfunction Aggravates Chronic Inflammation Through the NAD(+)/SIRT3 Axis and Promotes Renal Injury in Diabetes. Antioxidants 2025, 14, 267. [Google Scholar] [CrossRef]
  132. Mahapatra, C.; Thakkar, R.; Kumar, R. Modulatory Impact of Oxidative Stress on Action Potentials in Pathophysiological States: A Comprehensive Review. Antioxidants 2024, 13, 1172. [Google Scholar] [CrossRef]
  133. Hellenthal, K.E.M.; Brabenec, L.; Gross, E.R.; Wagner, N.M. TRP Channels as Sensors of Aldehyde and Oxidative Stress. Biomolecules 2021, 11, 1401. [Google Scholar] [CrossRef]
  134. Choi, S.-I.; Yoo, S.; Lim, J.Y.; Hwang, S.W. Are Sensory TRP Channels Biological Alarms for Lipid Peroxidation? Int. J. Mol. Sci. 2014, 15, 16430–16457. [Google Scholar] [CrossRef]
  135. Eskander, M.A.; Ruparel, S.; Green, D.P.; Chen, P.B.; Por, E.D.; Jeske, N.A.; Gao, X.; Flores, E.R.; Hargreaves, K.M. Persistent Nociception Triggered by Nerve Growth Factor (NGF) Is Mediated by TRPV1 and Oxidative Mechanisms. J. Neurosci. 2015, 35, 8593–8603. [Google Scholar] [CrossRef]
  136. Sahoo, N.; Hoshi, T.; Heinemann, S.H. Oxidative modulation of voltage-gated potassium channels. Antioxid. Redox Signal. 2014, 21, 933–952. [Google Scholar] [CrossRef]
  137. Matuz-Mares, D.; González-Andrade, M.; Araiza-Villanueva, M.G.; Vilchis-Landeros, M.M.; Vázquez-Meza, H. Mitochondrial Calcium: Effects of Its Imbalance in Disease. Antioxidants 2022, 11, 801. [Google Scholar] [CrossRef] [PubMed]
  138. Sidoryk-Węgrzynowicz, M.; Adamiak, K.; Strużyńska, L. Astrocyte-Neuron Interaction via the Glutamate-Glutamine Cycle and Its Dysfunction in Tau-Dependent Neurodegeneration. Int. J. Mol. Sci. 2024, 25, 3050. [Google Scholar] [CrossRef] [PubMed]
  139. Ladagu, A.D.; Olopade, F.E.; Adejare, A.; Olopade, J.O. GluN2A and GluN2B N-Methyl-D-Aspartate Receptor (NMDARs) Subunits: Their Roles and Therapeutic Antagonists in Neurological Diseases. Pharmaceuticals 2023, 16, 1535. [Google Scholar] [CrossRef]
  140. Santos, A.I.; Martínez-Ruiz, A.; Araújo, I.M. S-nitrosation and neuronal plasticity. Br. J. Pharmacol. 2015, 172, 1468–1478. [Google Scholar] [CrossRef] [PubMed]
  141. Silva, Á.J.; de Lavor, M.S. Nitroxidative Stress, Cell—Signaling Pathways, and Manganese Porphyrins: Therapeutic Potential in Neuropathic Pain. Int. J. Mol. Sci. 2025, 26, 2050. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, G.; Wei, H.; Zhao, A.; Yan, X.; Zhang, X.; Gan, J.; Guo, M.; Wang, J.; Zhang, F.; Jiang, Y.; et al. Mitochondrial DNA leakage: Underlying mechanisms and therapeutic implications in neurological disorders. J. Neuroinflamm. 2025, 22, 34. [Google Scholar] [CrossRef]
  143. Xian, H.; Watari, K.; Sanchez-Lopez, E.; Offenberger, J.; Onyuru, J.; Sampath, H.; Ying, W.; Hoffman, H.M.; Shadel, G.S.; Karin, M. Oxidized DNA fragments exit mitochondria via mPTP- and VDAC-dependent channels to activate NLRP3 inflammasome and interferon signaling. Immunity 2022, 55, 1370–1385.e8. [Google Scholar] [CrossRef] [PubMed]
  144. López-Doménech, G.; Kittler, J.T. Mitochondrial regulation of local supply of energy in neurons. Curr. Opin. Neurobiol. 2023, 81, 102747. [Google Scholar] [CrossRef]
  145. Rumpf, S.; Sanal, N.; Marzano, M. Energy metabolic pathways in neuronal development and function. Oxf. Open Neurosci. 2023, 2, kvad004. [Google Scholar] [CrossRef]
  146. Walters, G.C.; Usachev, Y.M. Mitochondrial calcium cycling in neuronal function and neurodegeneration. Front. Cell Dev. Biol. 2023, 11, 1094356. [Google Scholar] [CrossRef]
  147. Tosolini, A.P.; Abatecola, F.; Negro, S.; Sleigh, J.N.; Schiavo, G. The node of Ranvier influences the in vivo axonal transport of mitochondria and signaling endosomes. iScience 2024, 27, 111158. [Google Scholar] [CrossRef]
  148. Doyle, T.M.; Salvemini, D. Mini-Review: Mitochondrial dysfunction and chemotherapy-induced neuropathic pain. Neurosci. Lett. 2021, 760, 136087. [Google Scholar] [CrossRef]
  149. Kowalczyk, P.; Sulejczak, D.; Kleczkowska, P.; Bukowska-Ośko, I.; Kucia, M.; Popiel, M.; Wietrak, E.; Kramkowski, K.; Wrzosek, K.; Kaczyńska, K. Mitochondrial Oxidative Stress-A Causative Factor and Therapeutic Target in Many Diseases. Int. J. Mol. Sci. 2021, 22, 13384. [Google Scholar] [CrossRef]
  150. Palma, F.R.; Gantner, B.N.; Sakiyama, M.J.; Kayzuka, C.; Shukla, S.; Lacchini, R.; Cunniff, B.; Bonini, M.G. ROS production by mitochondria: Function or dysfunction? Oncogene 2024, 43, 295–303. [Google Scholar] [CrossRef] [PubMed]
  151. Paradies, G.; Paradies, V.; Ruggiero, F.M.; Petrosillo, G. Role of Cardiolipin in Mitochondrial Function and Dynamics in Health and Disease: Molecular and Pharmacological Aspects. Cells 2019, 8, 728. [Google Scholar] [CrossRef]
  152. Falabella, M.; Vernon, H.J.; Hanna, M.G.; Claypool, S.M.; Pitceathly, R.D.S. Cardiolipin, Mitochondria, and Neurological Disease. Trends Endocrinol. Metab. 2021, 32, 224–237. [Google Scholar] [CrossRef]
  153. Kasai, S.; Shimizu, S.; Tatara, Y.; Mimura, J.; Itoh, K. Regulation of Nrf2 by Mitochondrial Reactive Oxygen Species in Physiology and Pathology. Biomolecules 2020, 10, 320. [Google Scholar] [CrossRef] [PubMed]
  154. Silva Santos Ribeiro, P.; Willemen, H.; Eijkelkamp, N. Mitochondria and sensory processing in inflammatory and neuropathic pain. Front. Pain Res. 2022, 3, 1013577. [Google Scholar] [CrossRef]
  155. Flatters, S.J. The contribution of mitochondria to sensory processing and pain. Prog. Mol. Biol. Transl. Sci. 2015, 131, 119–146. [Google Scholar] [CrossRef] [PubMed]
  156. Willemen, H.; Santos Ribeiro, P.S.; Broeks, M.; Meijer, N.; Versteeg, S.; Tiggeler, A.; de Boer, T.P.; Małecki, J.M.; Falnes, P.; Jans, J.; et al. Inflammation-induced mitochondrial and metabolic disturbances in sensory neurons control the switch from acute to chronic pain. Cell Rep. Med. 2023, 4, 101265. [Google Scholar] [CrossRef] [PubMed]
  157. Jeyapalan, J.C.; Sedivy, J.M. Cellular senescence and organismal aging. Mech. Ageing Dev. 2008, 129, 467–474. [Google Scholar] [CrossRef]
  158. Di Micco, R.; Krizhanovsky, V.; Baker, D.; d’Adda di Fagagna, F. Cellular senescence in ageing: From mechanisms to therapeutic opportunities. Nat. Rev. Mol. Cell Biol. 2021, 22, 75–95. [Google Scholar] [CrossRef]
  159. Qin, Y.; Liu, H.; Wu, H. Cellular Senescence in Health, Disease, and Lens Aging. Pharmaceuticals 2025, 18, 244. [Google Scholar] [CrossRef]
  160. Song, X.; Liu, X.; Guo, Q.; Xu, H.; Cao, L. Unraveling the nexus between cellular senescence and malignant transformation: A paradigm shift in cancer research. Cancer Biol. Med. 2024, 21, 541–546. [Google Scholar] [CrossRef]
  161. Huang, W.; Hickson, L.J.; Eirin, A.; Kirkland, J.L.; Lerman, L.O. Cellular senescence: The good, the bad and the unknown. Nat. Rev. Nephrol. 2022, 18, 611–627. [Google Scholar] [CrossRef]
  162. Alqahtani, S.; Alqahtani, T.; Venkatesan, K.; Sivadasan, D.; Ahmed, R.; Sirag, N.; Elfadil, H.; Abdullah Mohamed, H.; Haseena, T.A.; Elsayed Ahmed, R.; et al. SASP Modulation for Cellular Rejuvenation and Tissue Homeostasis: Therapeutic Strategies and Molecular Insights. Cells 2025, 14, 608. [Google Scholar] [CrossRef]
  163. Gorgoulis, V.; Adams, P.D.; Alimonti, A.; Bennett, D.C.; Bischof, O.; Bishop, C.; Campisi, J.; Collado, M.; Evangelou, K.; Ferbeyre, G.; et al. Cellular Senescence: Defining a Path Forward. Cell 2019, 179, 813–827. [Google Scholar] [CrossRef]
  164. Kumari, R.; Jat, P. Mechanisms of Cellular Senescence: Cell Cycle Arrest and Senescence Associated Secretory Phenotype. Front. Cell Dev. Biol. 2021, 9, 645593. [Google Scholar] [CrossRef]
  165. Liu, J.; Yu, H.; Xu, Y. Targeting Cellular Senescence: Pathophysiology in Multisystem Age-Related Diseases. Biomedicines 2025, 13, 1727. [Google Scholar] [CrossRef] [PubMed]
  166. Prabhu, K.S.; Kuttikrishnan, S.; Ahmad, N.; Habeeba, U.; Mariyam, Z.; Suleman, M.; Bhat, A.A.; Uddin, S. H2AX: A key player in DNA damage response and a promising target for cancer therapy. Biomed. Pharmacother. 2024, 175, 116663. [Google Scholar] [CrossRef]
  167. Oda, T.; Gotoh, N.; Kasamatsu, T.; Handa, H.; Saitoh, T.; Sasaki, N. DNA damage-induced cellular senescence is regulated by 53BP1 accumulation in the nuclear foci and phase separation. Cell Prolif. 2023, 56, e13398. [Google Scholar] [CrossRef]
  168. Abuetabh, Y.; Wu, H.H.; Chai, C.; Al Yousef, H.; Persad, S.; Sergi, C.M.; Leng, R. DNA damage response revisited: The p53 family and its regulators provide endless cancer therapy opportunities. Exp. Mol. Med. 2022, 54, 1658–1669. [Google Scholar] [CrossRef]
  169. Osipov, A.; Chigasova, A.; Yashkina, E.; Ignatov, M.; Fedotov, Y.; Molodtsova, D.; Vorobyeva, N.; Osipov, A.N. Residual Foci of DNA Damage Response Proteins in Relation to Cellular Senescence and Autophagy in X-Ray Irradiated Fibroblasts. Cells 2023, 12, 1209. [Google Scholar] [CrossRef] [PubMed]
  170. Liu, X.L.; Ding, J.; Meng, L.H. Oncogene-induced senescence: A double edged sword in cancer. Acta Pharmacol. Sin. 2018, 39, 1553–1558. [Google Scholar] [CrossRef] [PubMed]
  171. Saleh, T.; Bloukh, S.; Hasan, M.; Al Shboul, S. Therapy-induced senescence as a component of tumor biology: Evidence from clinical cancer. Biochim. Biophys. Acta (BBA)-Rev. Cancer 2023, 1878, 188994. [Google Scholar] [CrossRef]
  172. Wiley, C.D.; Velarde, M.C.; Lecot, P.; Liu, S.; Sarnoski, E.A.; Freund, A.; Shirakawa, K.; Lim, H.W.; Davis, S.S.; Ramanathan, A.; et al. Mitochondrial Dysfunction Induces Senescence with a Distinct Secretory Phenotype. Cell Metab. 2016, 23, 303–314. [Google Scholar] [CrossRef]
  173. Faust, H.J.; Zhang, H.; Han, J.; Wolf, M.T.; Jeon, O.H.; Sadtler, K.; Peña, A.N.; Chung, L.; Maestas, D.R., Jr.; Tam, A.J.; et al. IL-17 and immunologically induced senescence regulate response to injury in osteoarthritis. J. Clin. Investig. 2020, 130, 5493–5507. [Google Scholar] [CrossRef]
  174. Wagner, V.; Gil, J. Senescence as a therapeutically relevant response to CDK4/6 inhibitors. Oncogene 2020, 39, 5165–5176. [Google Scholar] [CrossRef]
  175. Roger, L.; Tomas, F.; Gire, V. Mechanisms and Regulation of Cellular Senescence. Int. J. Mol. Sci. 2021, 22, 13173. [Google Scholar] [CrossRef] [PubMed]
  176. Zhao, H.; Darzynkiewicz, Z. Biomarkers of cell senescence assessed by imaging cytometry. In Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2012; Volume 965, pp. 83–92. [Google Scholar] [CrossRef]
  177. Song, S.; Lam, E.W.; Tchkonia, T.; Kirkland, J.L.; Sun, Y. Senescent Cells: Emerging Targets for Human Aging and Age-Related Diseases. Trends Biochem. Sci. 2020, 45, 578–592. [Google Scholar] [CrossRef] [PubMed]
  178. Valieva, Y.; Ivanova, E.; Fayzullin, A.; Kurkov, A.; Igrunkova, A. Senescence-Associated β-Galactosidase Detection in Pathology. Diagnostics 2022, 12, 2309. [Google Scholar] [CrossRef]
  179. Kudlova, N.; De Sanctis, J.B.; Hajduch, M. Cellular Senescence: Molecular Targets, Biomarkers, and Senolytic Drugs. Int. J. Mol. Sci. 2022, 23, 4168. [Google Scholar] [CrossRef] [PubMed]
  180. Ogrodnik, M. Cellular aging beyond cellular senescence: Markers of senescence prior to cell cycle arrest in vitro and in vivo. Aging Cell 2021, 20, e13338. [Google Scholar] [CrossRef]
  181. Saul, D.; Kosinsky, R.L.; Atkinson, E.J.; Doolittle, M.L.; Zhang, X.; LeBrasseur, N.K.; Pignolo, R.J.; Robbins, P.D.; Niedernhofer, L.J.; Ikeno, Y.; et al. A new gene set identifies senescent cells and predicts senescence-associated pathways across tissues. Nat. Commun. 2022, 13, 4827. [Google Scholar] [CrossRef]
  182. Xiao, S.; Qin, D.; Hou, X.; Tian, L.; Yu, Y.; Zhang, R.; Lyu, H.; Guo, D.; Chen, X.Z.; Zhou, C.; et al. Cellular senescence: A double-edged sword in cancer therapy. Front. Oncol. 2023, 13, 1189015. [Google Scholar] [CrossRef] [PubMed]
  183. Jena, A.B.; Samal, R.R.; Bhol, N.K.; Duttaroy, A.K. Cellular Red-Ox system in health and disease: The latest update. Biomed. Pharmacother. 2023, 162, 114606. [Google Scholar] [CrossRef] [PubMed]
  184. Hong, Y.; Boiti, A.; Vallone, D.; Foulkes, N.S. Reactive Oxygen Species Signaling and Oxidative Stress: Transcriptional Regulation and Evolution. Antioxidants 2024, 13, 312. [Google Scholar] [CrossRef]
  185. Juan, C.A.; Pérez de la Lastra, J.M.; Plou, F.J.; Pérez-Lebeña, E. The Chemistry of Reactive Oxygen Species (ROS) Revisited: Outlining Their Role in Biological Macromolecules (DNA, Lipids and Proteins) and Induced Pathologies. Int. J. Mol. Sci. 2021, 22, 4642. [Google Scholar] [CrossRef]
  186. Doksani, Y. The Response to DNA Damage at Telomeric Repeats and Its Consequences for Telomere Function. Genes 2019, 10, 318. [Google Scholar] [CrossRef]
  187. Al Bitar, S.; Gali-Muhtasib, H. The Role of the Cyclin Dependent Kinase Inhibitor p21(cip1/waf1) in Targeting Cancer: Molecular Mechanisms and Novel Therapeutics. Cancers 2019, 11, 1475. [Google Scholar] [CrossRef] [PubMed]
  188. Schmidt, A.; Allmann, S.; Schwarzenbach, C.; Snyder, P.; Chen, J.X.; Nagel, G.; Schöneis, A.; Rasenberger, B.; Beli, P.; Loewer, A.; et al. The p21CIP1-CDK4-DREAM axis is a master regulator of genotoxic stress-induced cellular senescence. Nucleic Acids Res. 2024, 52, 6945–6963. [Google Scholar] [CrossRef]
  189. Sławińska, N.; Krupa, R. Molecular Aspects of Senescence and Organismal Ageing-DNA Damage Response, Telomeres, Inflammation and Chromatin. Int. J. Mol. Sci. 2021, 22, 590. [Google Scholar] [CrossRef]
  190. Wen, H.; Deng, H.; Li, B.; Chen, J.; Zhu, J.; Zhang, X.; Yoshida, S.; Zhou, Y. Mitochondrial diseases: From molecular mechanisms to therapeutic advances. Signal Transduct. Target. Ther. 2025, 10, 9. [Google Scholar] [CrossRef]
  191. Martini, H.; Passos, J.F. Cellular senescence: All roads lead to mitochondria. FEBS J. 2023, 290, 1186–1202. [Google Scholar] [CrossRef]
  192. Hwang, I.; Uddin, M.J.; Pak, E.S.; Kang, H.; Jin, E.-J.; Jo, S.; Kang, D.; Lee, H.; Ha, H. The impaired redox balance in peroxisomes of catalase knockout mice accelerates nonalcoholic fatty liver disease through endoplasmic reticulum stress. Free Radic. Biol. Med. 2020, 148, 22–32. [Google Scholar] [CrossRef]
  193. Terlecky, S.R.; Koepke, J.I.; Walton, P.A. Peroxisomes and aging. Biochim. Biophys. Acta (BBA)-Mol. Cell Res. 2006, 1763, 1749–1754. [Google Scholar] [CrossRef] [PubMed]
  194. Summer, R.; Shaghaghi, H.; Schriner, D.; Roque, W.; Sales, D.; Cuevas-Mora, K.; Desai, V.; Bhushan, A.; Ramirez, M.I.; Romero, F. Activation of the mTORC1/PGC-1 axis promotes mitochondrial biogenesis and induces cellular senescence in the lung epithelium. Am. J. Physiol. Lung Cell. Mol. Physiol. 2019, 316, L1049–L1060. [Google Scholar] [CrossRef] [PubMed]
  195. Kim, Y.; Jang, Y.; Kim, M.S.; Kang, C. Metabolic remodeling in cancer and senescence and its therapeutic implications. Trends Endocrinol. Metab. 2024, 35, 732–744. [Google Scholar] [CrossRef] [PubMed]
  196. Hinchy, E.C.; Gruszczyk, A.V.; Willows, R.; Navaratnam, N.; Hall, A.R.; Bates, G.; Bright, T.P.; Krieg, T.; Carling, D.; Murphy, M.P. Mitochondria-derived ROS activate AMP-activated protein kinase (AMPK) indirectly. J. Biol. Chem. 2018, 293, 17208–17217. [Google Scholar] [CrossRef]
  197. Liu, S.; Jiang, X.; Cui, X.; Wang, J.; Liu, S.; Li, H.; Yang, J.; Zhang, C.; Zhang, W. Smooth muscle-specific HuR knockout induces defective autophagy and atherosclerosis. Cell Death Dis. 2021, 12, 385. [Google Scholar] [CrossRef]
  198. Han, Z.; Wang, K.; Ding, S.; Zhang, M. Cross-talk of inflammation and cellular senescence: A new insight into the occurrence and progression of osteoarthritis. Bone Res. 2024, 12, 69. [Google Scholar] [CrossRef]
  199. Gao, X.; Yu, X.; Zhang, C.; Wang, Y.; Sun, Y.; Sun, H.; Zhang, H.; Shi, Y.; He, X. Telomeres and Mitochondrial Metabolism: Implications for Cellular Senescence and Age-related Diseases. Stem Cell Rev. Rep. 2022, 18, 2315–2327. [Google Scholar] [CrossRef]
  200. Kandhaya-Pillai, R.; Miro-Mur, F.; Alijotas-Reig, J.; Tchkonia, T.; Schwartz, S.; Kirkland, J.L.; Oshima, J. Key elements of cellular senescence involve transcriptional repression of mitotic and DNA repair genes through the p53-p16/RB-E2F-DREAM complex. Aging 2023, 15, 4012–4034. [Google Scholar] [CrossRef]
  201. Kale, A.; Rogers, N.M.; Ghimire, K. Thrombospondin-1 CD47 Signalling: From Mechanisms to Medicine. Int. J. Mol. Sci. 2021, 22, 4062. [Google Scholar] [CrossRef]
  202. Singla, B.; Aithbathula, R.V.; Pervaiz, N.; Kathuria, I.; Swanson, M.; Ekuban, F.A.; Ahn, W.; Park, F.; Gyamfi, M.; Cherian-Shaw, M.; et al. CD47 Activation by Thrombospondin-1 in Lymphatic Endothelial Cells Suppresses Lymphangiogenesis and Promotes Atherosclerosis. Arter. Thromb. Vasc. Biol. 2023, 43, 1234–1250. [Google Scholar] [CrossRef]
  203. Pedroza-Diaz, J.; Arroyave-Ospina, J.C.; Serna Salas, S.; Moshage, H. Modulation of Oxidative Stress-Induced Senescence during Non-Alcoholic Fatty Liver Disease. Antioxidants 2022, 11, 975. [Google Scholar] [CrossRef]
  204. Li, X.; Li, C.; Zhang, W.; Wang, Y.; Qian, P.; Huang, H. Inflammation and aging: Signaling pathways and intervention therapies. Signal Transduct. Target. Ther. 2023, 8, 239. [Google Scholar] [CrossRef] [PubMed]
  205. Baechle, J.J.; Chen, N.; Makhijani, P.; Winer, S.; Furman, D.; Winer, D.A. Chronic inflammation and the hallmarks of aging. Mol. Metab. 2023, 74, 101755. [Google Scholar] [CrossRef]
  206. Boccardi, V.; Marano, L. Aging, Cancer, and Inflammation: The Telomerase Connection. Int. J. Mol. Sci. 2024, 25, 8542. [Google Scholar] [CrossRef]
  207. Ren, J.L.; Pan, J.S.; Lu, Y.P.; Sun, P.; Han, J. Inflammatory signaling and cellular senescence. Cell. Signal. 2009, 21, 378–383. [Google Scholar] [CrossRef] [PubMed]
  208. Herbstein, F.; Sapochnik, M.; Attorresi, A.; Pollak, C.; Senin, S.; Gonilski-Pacin, D.; Ciancio Del Giudice, N.; Fiz, M.; Elguero, B.; Fuertes, M.; et al. The SASP factor IL-6 sustains cell-autonomous senescent cells via a cGAS-STING-NFκB intracrine senescent noncanonical pathway. Aging Cell 2024, 23, e14258. [Google Scholar] [CrossRef] [PubMed]
  209. Ortiz-Montero, P.; Londoño-Vallejo, A.; Vernot, J.P. Senescence-associated IL-6 and IL-8 cytokines induce a self- and cross-reinforced senescence/inflammatory milieu strengthening tumorigenic capabilities in the MCF-7 breast cancer cell line. Cell Commun. Signal. 2017, 15, 17. [Google Scholar] [CrossRef]
  210. Hu, Z.; Sui, Q.; Jin, X.; Shan, G.; Huang, Y.; Yi, Y.; Zeng, D.; Zhao, M.; Zhan, C.; Wang, Q.; et al. IL6-STAT3-C/EBPβ-IL6 positive feedback loop in tumor-associated macrophages promotes the EMT and metastasis of lung adenocarcinoma. J. Exp. Clin. Cancer Res. 2024, 43, 63. [Google Scholar] [CrossRef]
  211. Khavinson, V.; Linkova, N.; Dyatlova, A.; Kantemirova, R.; Kozlov, K. Senescence-Associated Secretory Phenotype of Cardiovascular System Cells and Inflammaging: Perspectives of Peptide Regulation. Cells 2022, 12, 106. [Google Scholar] [CrossRef]
  212. Prata, L.; Ovsyannikova, I.G.; Tchkonia, T.; Kirkland, J.L. Senescent cell clearance by the immune system: Emerging therapeutic opportunities. Semin. Immunol. 2018, 40, 101275. [Google Scholar] [CrossRef] [PubMed]
  213. Kale, A.; Sharma, A.; Stolzing, A.; Desprez, P.-Y.; Campisi, J. Role of immune cells in the removal of deleterious senescent cells. Immun. Ageing 2020, 17, 16. [Google Scholar] [CrossRef] [PubMed]
  214. Sah, E.; Krishnamurthy, S.; Ahmidouch, M.Y.; Gillispie, G.J.; Milligan, C.; Orr, M.E. The Cellular Senescence Stress Response in Post-Mitotic Brain Cells: Cell Survival at the Expense of Tissue Degeneration. Life 2021, 11, 229. [Google Scholar] [CrossRef]
  215. Donovan, L.J.; Brewer, C.L.; Bond, S.F.; Laslavic, A.M.; Pena Lopez, A.; Colman, L.; Jordan, C.E.; Hansen, L.H.; González, O.C.; Pujari, A.; et al. Aging and injury drive neuronal senescence in the dorsal root ganglia. Nat. Neurosci. 2025, 28, 985–997. [Google Scholar] [CrossRef]
  216. Herdy, J.R.; Traxler, L.; Agarwal, R.K.; Karbacher, L.; Schlachetzki, J.C.M.; Boehnke, L.; Zangwill, D.; Galasko, D.; Glass, C.K.; Mertens, J.; et al. Increased post-mitotic senescence in aged human neurons is a pathological feature of Alzheimer’s disease. Cell Stem Cell 2022, 29, 1637–1652.e6. [Google Scholar] [CrossRef] [PubMed]
  217. Dehkordi, S.K.; Walker, J.; Sah, E.; Bennett, E.; Atrian, F.; Frost, B.; Woost, B.; Bennett, R.E.; Orr, T.C.; Zhou, Y.; et al. Profiling senescent cells in human brains reveals neurons with CDKN2D/p19 and tau neuropathology. Nat. Aging 2021, 1, 1107–1116. [Google Scholar] [CrossRef]
  218. Muralidharan, A.; Sotocinal, S.G.; Yousefpour, N.; Akkurt, N.; Lima, L.V.; Tansley, S.; Parisien, M.; Wang, C.; Austin, J.S.; Ham, B.; et al. Long-term male-specific chronic pain via telomere- and p53-mediated spinal cord cellular senescence. J. Clin. Investig. 2022, 132, e151817. [Google Scholar] [CrossRef]
  219. Wong, G.C.; Chow, K.H. DNA Damage Response-Associated Cell Cycle Re-Entry and Neuronal Senescence in Brain Aging and Alzheimer’s Disease. J. Alzheimer’s Dis. 2023, 94, S429–S451. [Google Scholar] [CrossRef]
  220. Fielder, E.; von Zglinicki, T.; Jurk, D. The DNA Damage Response in Neurons: Die by Apoptosis or Survive in a Senescence-Like State? J. Alzheimer’s Dis. 2017, 60, S107–S131. [Google Scholar] [CrossRef]
  221. Cui, J.G.; Holmin, S.; Mathiesen, T.; Meyerson, B.A.; Linderoth, B. Possible role of inflammatory mediators in tactile hypersensitivity in rat models of mononeuropathy. PAIN 2000, 88, 239–248. [Google Scholar] [CrossRef]
  222. Heinrich, P.C.; Behrmann, I.; Müller-Newen, G.; Schaper, F.; Graeve, L. Interleukin-6-type cytokine signalling through the gp130/Jak/STAT pathway. Biochem. J. 1998, 334 Pt 2, 297–314. [Google Scholar] [CrossRef] [PubMed]
  223. Kummer, K.K.; Zeidler, M.; Kalpachidou, T.; Kress, M. Role of IL-6 in the regulation of neuronal development, survival and function. Cytokine 2021, 144, 155582. [Google Scholar] [CrossRef]
  224. Fang, D.; Kong, L.Y.; Cai, J.; Li, S.; Liu, X.D.; Han, J.S.; Xing, G.G. Interleukin-6-mediated functional upregulation of TRPV1 receptors in dorsal root ganglion neurons through the activation of JAK/PI3K signaling pathway: Roles in the development of bone cancer pain in a rat model. PAIN 2015, 156, 1124–1144. [Google Scholar] [CrossRef]
  225. Ungerleider, K.; Beck, J.; Lissa, D.; Turnquist, C.; Horikawa, I.; Harris, B.T.; Harris, C.C. Astrocyte senescence and SASP in neurodegeneration: Tau joins the loop. Cell Cycle 2021, 20, 752–764. [Google Scholar] [CrossRef]
  226. Moreno-Blas, D.; Gorostieta-Salas, E.; Pommer-Alba, A.; Muciño-Hernández, G.; Gerónimo-Olvera, C.; Maciel-Barón, L.A.; Konigsberg, M.; Massieu, L.; Castro-Obregón, S. Cortical neurons develop a senescence-like phenotype promoted by dysfunctional autophagy. Aging 2019, 11, 6175–6198. [Google Scholar] [CrossRef]
  227. Paramos-de-Carvalho, D.; Martins, I.; Cristóvão, A.M.; Dias, A.F.; Neves-Silva, D.; Pereira, T.; Chapela, D.; Farinho, A.; Jacinto, A.; Saúde, L. Targeting senescent cells improves functional recovery after spinal cord injury. Cell Rep. 2021, 36, 109334. [Google Scholar] [CrossRef]
  228. Zhou, Y.-Q.; Mei, W.; Tian, X.-B.; Tian, Y.-K.; Liu, D.-Q.; Ye, D.-W. The therapeutic potential of Nrf2 inducers in chronic pain: Evidence from preclinical studies. Pharmacol. Ther. 2021, 225, 107846. [Google Scholar] [CrossRef]
  229. Basu, P.; Averitt, D.L.; Maier, C.; Basu, A. The Effects of Nuclear Factor Erythroid 2 (NFE2)-Related Factor 2 (Nrf2) Activation in Preclinical Models of Peripheral Neuropathic Pain. Antioxidants 2022, 11, 430. [Google Scholar] [CrossRef] [PubMed]
  230. Gil, J. The interplay between senescence, inflammation, and the immune system. Genes Dev. 2025, 39, 923–925. [Google Scholar] [CrossRef] [PubMed]
  231. Varesi, A.; Chirumbolo, S.; Campagnoli, L.I.M.; Pierella, E.; Piccini, G.B.; Carrara, A.; Ricevuti, G.; Scassellati, C.; Bonvicini, C.; Pascale, A. The Role of Antioxidants in the Interplay between Oxidative Stress and Senescence. Antioxidants 2022, 11, 1224. [Google Scholar] [CrossRef]
  232. Baird, L.; Taguchi, K.; Zhang, A.; Takahashi, Y.; Suzuki, T.; Kensler, T.W.; Yamamoto, M. A NRF2-induced secretory phenotype activates immune surveillance to remove irreparably damaged cells. Redox Biol. 2023, 66, 102845. [Google Scholar] [CrossRef]
  233. Bendavit, G.; Aboulkassim, T.; Hilmi, K.; Shah, S.; Batist, G. Nrf2 Transcription Factor Can Directly Regulate mTOR: Linking cytoprotective gene expression to a major metabolic regulator that generates redox activity. J. Biol. Chem. 2016, 291, 25476–25488. [Google Scholar] [CrossRef]
  234. Baiskhanova, D.; Schäfer, H. The Role of Nrf2 in the Regulation of Mitochondrial Function and Ferroptosis in Pancreatic Cancer. Antioxidants 2024, 13, 696. [Google Scholar] [CrossRef] [PubMed]
  235. Ning, B.; Hang, S.; Zhang, W.; Mao, C.; Li, D. An update on the bridging factors connecting autophagy and Nrf2 antioxidant pathway. Front. Cell Dev. Biol. 2023, 11, 1232241. [Google Scholar] [CrossRef] [PubMed]
  236. Ngo, V.; Duennwald, M.L. Nrf2 and Oxidative Stress: A General Overview of Mechanisms and Implications in Human Disease. Antioxidants 2022, 11, 2345. [Google Scholar] [CrossRef] [PubMed]
  237. Baird, L.; Yamamoto, M. The Molecular Mechanisms Regulating the KEAP1-NRF2 Pathway. Mol. Cell. Biol. 2020, 40, e00099-20. [Google Scholar] [CrossRef]
  238. Ahmed, S.M.; Luo, L.; Namani, A.; Wang, X.J.; Tang, X. Nrf2 signaling pathway: Pivotal roles in inflammation. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 585–597. [Google Scholar] [CrossRef]
  239. Saha, S.; Buttari, B.; Panieri, E.; Profumo, E.; Saso, L. An Overview of Nrf2 Signaling Pathway and Its Role in Inflammation. Molecules 2020, 25, 5474. [Google Scholar] [CrossRef]
  240. He, F.; Ru, X.; Wen, T. NRF2, a Transcription Factor for Stress Response and Beyond. Int. J. Mol. Sci. 2020, 21, 4777. [Google Scholar] [CrossRef]
  241. Jiang, T.; Harder, B.; Rojo de la Vega, M.; Wong, P.K.; Chapman, E.; Zhang, D.D. p62 links autophagy and Nrf2 signaling. Free Radic. Biol. Med. 2015, 88, 199–204. [Google Scholar] [CrossRef]
  242. Sotolongo, K.; Ghiso, J.; Rostagno, A. Nrf2 activation through the PI3K/GSK-3 axis protects neuronal cells from Aβ-mediated oxidative and metabolic damage. Alzheimer’s Res. Ther. 2020, 12, 13. [Google Scholar] [CrossRef]
  243. Hasan, S.K.; Jayakumar, S.; Espina Barroso, E.; Jha, A.; Catalano, G.; Sandur, S.K.; Noguera, N.I. Molecular Targets of Oxidative Stress: Focus on Nuclear Factor Erythroid 2–Related Factor 2 Function in Leukemia and Other Cancers. Cells 2025, 14, 713. [Google Scholar] [CrossRef] [PubMed]
  244. Schmidlin, C.J.; Shakya, A.; Dodson, M.; Chapman, E.; Zhang, D.D. The intricacies of NRF2 regulation in cancer. Semin. Cancer Biol. 2021, 76, 110–119. [Google Scholar] [CrossRef]
  245. Hao, W.; Li, M.; Cai, Q.; Wu, S.; Li, X.; He, Q.; Hu, Y. Roles of NRF2 in Fibrotic Diseases: From Mechanisms to Therapeutic Approaches. Front. Physiol. 2022, 13, 889792. [Google Scholar] [CrossRef] [PubMed]
  246. Li, S.; Eguchi, N.; Lau, H.; Ichii, H. The Role of the Nrf2 Signaling in Obesity and Insulin Resistance. Int. J. Mol. Sci. 2020, 21, 6973. [Google Scholar] [CrossRef]
  247. Freeborn, R.A.; Rockwell, C.E. Chapter Two—The role of Nrf2 in autoimmunity and infectious disease: Therapeutic possibilities. In Advances in Pharmacology; Copple, B.L., Rockwell, C.E., Eds.; Academic Press: New York, NY, USA, 2021; Volume 91, pp. 61–110. [Google Scholar]
  248. Teleanu, D.M.; Niculescu, A.G.; Lungu, I.I.; Radu, C.I.; Vladâcenco, O.; Roza, E.; Costăchescu, B.; Grumezescu, A.M.; Teleanu, R.I. An Overview of Oxidative Stress, Neuroinflammation, and Neurodegenerative Diseases. Int. J. Mol. Sci. 2022, 23, 5938. [Google Scholar] [CrossRef] [PubMed]
  249. Chen, Y.; Li, D.; Li, N.; Loh, P.; Guo, Y.; Hu, X.; Zhang, J.; Dou, B.; Wang, L.; Yang, C.; et al. Role of nerve signal transduction and neuroimmune crosstalk in mediating the analgesic effects of acupuncture for neuropathic pain. Front. Neurol. 2023, 14, 1093849. [Google Scholar] [CrossRef]
  250. Orfali, R.; Alwatban, A.Z.; Orfali, R.S.; Lau, L.; Chea, N.; Alotaibi, A.M.; Nam, Y.W.; Zhang, M. Oxidative stress and ion channels in neurodegenerative diseases. Front. Physiol. 2024, 15, 1320086. [Google Scholar] [CrossRef]
  251. Rosa, A.O.; Egea, J.; Lorrio, S.; Rojo, A.I.; Cuadrado, A.; López, M.G. Nrf2-mediated haeme oxygenase-1 up-regulation induced by cobalt protoporphyrin has antinociceptive effects against inflammatory pain in the formalin test in mice. PAIN 2008, 137, 332–339. [Google Scholar] [CrossRef]
  252. Li, W.; Jiang, D.; Li, Q.; Yao, S.; Sun, X.; Yang, Y.; Meng, Z.; Liu, W. Lipopolysaccharide-induced preconditioning protects against traumatic spinal cord injury by upregulating Nrf2 expression in rats. Life Sci. 2016, 162, 14–20. [Google Scholar] [CrossRef]
  253. Zhao, L.; Tao, X.; Wang, Q.; Yu, X.; Dong, D. Diosmetin alleviates neuropathic pain by regulating the Keap1/Nrf2/NF-κB signaling pathway. Biomed. Pharmacother. 2024, 170, 116067. [Google Scholar] [CrossRef] [PubMed]
  254. Zhang, M.W.; Sun, X.; Xu, Y.W.; Meng, W.; Tang, Q.; Gao, H.; Liu, L.; Chen, S.H. Curcumin relieves oxaliplatin-induced neuropathic pain via reducing inflammation and activating antioxidant response. Cell Biol. Int. 2024, 48, 872–882. [Google Scholar] [CrossRef]
  255. Miao, H.; Xu, J.; Xu, D.; Ma, X.; Zhao, X.; Liu, L. Nociceptive behavior induced by chemotherapeutic paclitaxel and beneficial role of antioxidative pathways. Physiol. Res. 2019, 68, 491–500. [Google Scholar] [CrossRef] [PubMed]
  256. Xue, C.; Kui, W.; Huang, A.; Li, Y.; Li, L.; Gu, Z.; Xie, L.; Kong, S.; Yu, J.; Ruan, H.; et al. Electroacupuncture suppresses neuronal ferroptosis to relieve chronic neuropathic pain. J. Cell. Mol. Med. 2024, 28, e18240. [Google Scholar] [CrossRef] [PubMed]
  257. Elsayed, H.R.H.; Ali, E.M.T.; Rabei, M.R.; El Nashar, E.M.; Alghamdi, M.A.; Al-Zahrani, N.S.; Alshehri, S.H.; Aldahhan, R.A.; Morsy, A.I. Angiotensin II Type 1 receptor blockade attenuates the neuropathological changes in the spinal cords of diabetic rats with modulation of nuclear factor erythroid 2-related factor 2/ heme oxygenase 1 system. Tissue Cell 2024, 88, 102420. [Google Scholar] [CrossRef]
  258. Chen, H.; Xie, K.; Chen, Y.; Wang, Y.; Wang, Y.; Lian, N.; Zhang, K.; Yu, Y. Nrf2/HO-1 signaling pathway participated in the protection of hydrogen sulfide on neuropathic pain in rats. Int. Immunopharmacol. 2019, 75, 105746. [Google Scholar] [CrossRef]
  259. Zhou, Y.Q.; Liu, D.Q.; Chen, S.P.; Chen, N.; Sun, J.; Wang, X.M.; Cao, F.; Tian, Y.K.; Ye, D.W. Nrf2 activation ameliorates mechanical allodynia in paclitaxel-induced neuropathic pain. Acta Pharmacol. Sin. 2020, 41, 1041–1048. [Google Scholar] [CrossRef]
  260. Zhang, M.; Hu, W.; Cai, C.; Wu, Y.; Li, J.; Dong, S. Advanced application of stimuli-responsive drug delivery system for inflammatory arthritis treatment. Mater. Today Bio 2022, 14, 100223. [Google Scholar] [CrossRef]
  261. An, J.; Chen, B.; Kang, X.; Zhang, R.; Guo, Y.; Zhao, J.; Yang, H. Neuroprotective effects of natural compounds on LPS-induced inflammatory responses in microglia. Am. J. Transl. Res. 2020, 12, 2353–2378. [Google Scholar]
  262. Kim, H.K.; Wang, Q.; Hwang, S.H.; Dougherty, P.M.; Wang, J.; Abdi, S. Bardoxolone Methyl Ameliorates Chemotherapy-Induced Neuropathic Pain by Activation of Phosphorylated Nuclear Factor Erythroid 2-Related Factor 2 in the Dorsal Root Ganglia. Anesth. Analg. 2024, 138, 664–675. [Google Scholar] [CrossRef]
  263. Lu, Y.Y.; Tsai, H.P.; Tsai, T.H.; Miao, H.C.; Zhang, Z.H.; Wu, C.H. RTA-408 Regulates p-NF-κB/TSLP/STAT5 Signaling to Ameliorate Nociceptive Hypersensitivity in Chronic Constriction Injury Rats. Mol. Neurobiol. 2024, 61, 1714–1725. [Google Scholar] [CrossRef] [PubMed]
  264. Sun, Y.; Yu, F.; Cao, W.; Zhang, W.; Liu, W.; Dai, F. Betulinic acid alleviates neuropathic pain induced by chronic constriction injury of the sciatic nerve in mice. Neurosci. Lett. 2023, 813, 137429. [Google Scholar] [CrossRef] [PubMed]
  265. Mokhtari, T.; Lu, M.; El-Kenawy, A.E. Potential anxiolytic and antidepressant-like effects of luteolin in a chronic constriction injury rat model of neuropathic pain: Role of oxidative stress, neurotrophins, and inflammatory factors. Int. Immunopharmacol. 2023, 122, 110520. [Google Scholar] [CrossRef]
  266. Ray, P.D.; Huang, B.W.; Tsuji, Y. Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling. Cell. Signal. 2012, 24, 981–990. [Google Scholar] [CrossRef]
  267. Panieri, E.; Telkoparan-Akillilar, P.; Suzen, S.; Saso, L. The NRF2/KEAP1 Axis in the Regulation of Tumor Metabolism: Mechanisms and Therapeutic Perspectives. Biomolecules 2020, 10, 791. [Google Scholar] [CrossRef] [PubMed]
  268. Papadopoli, D.; Boulay, K.; Kazak, L.; Pollak, M.; Mallette, F.A.; Topisirovic, I.; Hulea, L. mTOR as a central regulator of lifespan and aging. F1000Research 2019, 8, 998. [Google Scholar] [CrossRef]
  269. Li, J.; Ma, J.; Lacagnina, M.J.; Lorca, S.; Odem, M.A.; Walters, E.T.; Kavelaars, A.; Grace, P.M. Oral Dimethyl Fumarate Reduces Peripheral Neuropathic Pain in Rodents via NFE2L2 Antioxidant Signaling. Anesthesiology 2020, 132, 343–356. [Google Scholar] [CrossRef]
  270. Spigoni, V.; Cinquegrani, G.; Iannozzi, N.T.; Frigeri, G.; Maggiolo, G.; Maggi, M.; Parello, V.; Dei Cas, A. Activation of G protein-coupled receptors by ketone bodies: Clinical implication of the ketogenic diet in metabolic disorders. Front. Endocrinol. 2022, 13, 972890. [Google Scholar] [CrossRef]
  271. Parodi, B.; Rossi, S.; Morando, S.; Cordano, C.; Bragoni, A.; Motta, C.; Usai, C.; Wipke, B.T.; Scannevin, R.H.; Mancardi, G.L.; et al. Fumarates modulate microglia activation through a novel HCAR2 signaling pathway and rescue synaptic dysregulation in inflamed CNS. Acta Neuropathol. 2015, 130, 279–295. [Google Scholar] [CrossRef]
  272. Rosito, M.; Testi, C.; Parisi, G.; Cortese, B.; Baiocco, P.; Di Angelantonio, S. Exploring the Use of Dimethyl Fumarate as Microglia Modulator for Neurodegenerative Diseases Treatment. Antioxidants 2020, 9, 700. [Google Scholar] [CrossRef]
  273. Chen, H.; Assmann, J.C.; Krenz, A.; Rahman, M.; Grimm, M.; Karsten, C.M.; Köhl, J.; Offermanns, S.; Wettschureck, N.; Schwaninger, M. Hydroxycarboxylic acid receptor 2 mediates dimethyl fumarate’s protective effect in EAE. J. Clin. Investig. 2014, 124, 2188–2192. [Google Scholar] [CrossRef]
  274. Boccella, S.; Guida, F.; De Logu, F.; De Gregorio, D.; Mazzitelli, M.; Belardo, C.; Iannotta, M.; Serra, N.; Nassini, R.; de Novellis, V.; et al. Ketones and pain: Unexplored role of hydroxyl carboxylic acid receptor type 2 in the pathophysiology of neuropathic pain. FASEB J. 2019, 33, 1062–1073. [Google Scholar] [CrossRef] [PubMed]
  275. Taing, K.; Chen, L.; Weng, H.-R. Emerging roles of GPR109A in regulation of neuroinflammation in neurological diseases and pain. Neural Regen. Res. 2023, 18, 763–768. [Google Scholar] [PubMed]
  276. Zhao, C.; Wang, H.; Liu, Y.; Cheng, L.; Wang, B.; Tian, X.; Fu, H.; Wu, C.; Li, Z.; Shen, C.; et al. Biased allosteric activation of ketone body receptor HCAR2 suppresses inflammation. Mol. Cell 2023, 83, 3171–3187.e7. [Google Scholar] [CrossRef] [PubMed]
  277. Testai, L.; Guida, F.; Salerno, S.; Brogi, S.; Morace, A.M.; Carbonetti, L.; Ricciardi, F.; Perrone, M.; Palazzo, E.; Calderone, V.; et al. Hydroxyl carboxylic acid receptor-2 (HCAR2) as a potential target in neurometabolic diseases. Pharmacol. Ther. 2025, 274, 108909. [Google Scholar] [CrossRef]
Figure 1. Neuroimmune IL-1β–COX-2–PGE2 pathway in spinal nociceptive sensitization. Immune cells, microglia, and astroglia release IL-1β, which triggers NF-κB and MAPK signaling, leading to COX-2 induction and prostaglandin E2 production. PGE2 engages E-type prostanoid receptors on neurons to increase neuronal firing and synaptic transmission, sustaining central sensitization and chronic pain.
Figure 1. Neuroimmune IL-1β–COX-2–PGE2 pathway in spinal nociceptive sensitization. Immune cells, microglia, and astroglia release IL-1β, which triggers NF-κB and MAPK signaling, leading to COX-2 induction and prostaglandin E2 production. PGE2 engages E-type prostanoid receptors on neurons to increase neuronal firing and synaptic transmission, sustaining central sensitization and chronic pain.
Antioxidants 14 01166 g001
Figure 2. Peripheral immune orchestration of nociceptor sensitization at an injured nerve. Overview of the peripheral tissue environment adjacent to a nociceptor terminal. Neutrophils provide early host defense and opioid-mediated analgesia but, under persistent inflammation, form NETs that aggravate tissue injury. Efferocytosis by macrophages clears apoptotic neutrophils and supports resolution. Monocytes entering from the vasculature differentiate into macrophages that produce TNF, IL-1β, IL-6, PGE2, NGF, IGF-1, CCL2, and CXCL1 in the perineuronal space. These mediators act on the nociceptor ending to increase TRPA1/TRPV1 activity and the availability of Nav1.7–Nav1.9 channels, culminating in increased neuronal excitability and spontaneous activity.
Figure 2. Peripheral immune orchestration of nociceptor sensitization at an injured nerve. Overview of the peripheral tissue environment adjacent to a nociceptor terminal. Neutrophils provide early host defense and opioid-mediated analgesia but, under persistent inflammation, form NETs that aggravate tissue injury. Efferocytosis by macrophages clears apoptotic neutrophils and supports resolution. Monocytes entering from the vasculature differentiate into macrophages that produce TNF, IL-1β, IL-6, PGE2, NGF, IGF-1, CCL2, and CXCL1 in the perineuronal space. These mediators act on the nociceptor ending to increase TRPA1/TRPV1 activity and the availability of Nav1.7–Nav1.9 channels, culminating in increased neuronal excitability and spontaneous activity.
Antioxidants 14 01166 g002
Figure 3. Central glial mechanisms sustaining neuropathic pain. Peripheral nerve injury triggers central inflammation within the dorsal horn of the spinal cord. Microglia are the earliest responders, releasing TNF, IL-1β, PGE2, and BDNF that enhance pain signaling and initiate central sensitization. Astrocytes become reactive slightly later (“astrogliosis”) and maintain sensitization for months through persistent cytokine and prostaglandin production. Oligodendrocytes provide a supportive contribution by releasing IL-33, which activates ST2 receptors on neurons and glia to engage PI3K and MAPK (p38/ERK/JNK) and NF-κB signaling, thereby increasing pain pathway sensitivity. Together, these glial interactions create a self-reinforcing inflammatory network that stabilizes central sensitization and chronic pain.
Figure 3. Central glial mechanisms sustaining neuropathic pain. Peripheral nerve injury triggers central inflammation within the dorsal horn of the spinal cord. Microglia are the earliest responders, releasing TNF, IL-1β, PGE2, and BDNF that enhance pain signaling and initiate central sensitization. Astrocytes become reactive slightly later (“astrogliosis”) and maintain sensitization for months through persistent cytokine and prostaglandin production. Oligodendrocytes provide a supportive contribution by releasing IL-33, which activates ST2 receptors on neurons and glia to engage PI3K and MAPK (p38/ERK/JNK) and NF-κB signaling, thereby increasing pain pathway sensitivity. Together, these glial interactions create a self-reinforcing inflammatory network that stabilizes central sensitization and chronic pain.
Antioxidants 14 01166 g003
Figure 4. Representative reactive oxygen species relevant to oxidative stress. Schematic depiction of major ROS, including molecular oxygen (O2), superoxide anion (O2), peroxide anion (O22−), hydroxyl radical (•OH), and hydroxide ion (OH). Although these species can mediate physiological redox signaling at low levels, their accumulation drives oxidative damage to nucleic acids, proteins, and lipids, contributing to cellular dysfunction and disease progression.
Figure 4. Representative reactive oxygen species relevant to oxidative stress. Schematic depiction of major ROS, including molecular oxygen (O2), superoxide anion (O2), peroxide anion (O22−), hydroxyl radical (•OH), and hydroxide ion (OH). Although these species can mediate physiological redox signaling at low levels, their accumulation drives oxidative damage to nucleic acids, proteins, and lipids, contributing to cellular dysfunction and disease progression.
Antioxidants 14 01166 g004
Figure 5. Major triggers and checkpoint pathways driving cellular senescence. Schematic overview of key senescence inducers—oxidative stress, endoplasmic-reticulum (ER) stress, DNA damage/DDR activation, oncogenic signaling, mitochondrial dysfunction, and telomere shortening—converging on canonical checkpoint modules within a normal cell. p53 stabilization with p21 inhibits CDKs; p16INK4A restrains CDK4/6; CDK2–Cyclin E control is depicted together with the Rb–E2F node. The outcome is durable cell-cycle arrest and a senescent phenotype characterized by enlarged morphology and secretory activity.
Figure 5. Major triggers and checkpoint pathways driving cellular senescence. Schematic overview of key senescence inducers—oxidative stress, endoplasmic-reticulum (ER) stress, DNA damage/DDR activation, oncogenic signaling, mitochondrial dysfunction, and telomere shortening—converging on canonical checkpoint modules within a normal cell. p53 stabilization with p21 inhibits CDKs; p16INK4A restrains CDK4/6; CDK2–Cyclin E control is depicted together with the Rb–E2F node. The outcome is durable cell-cycle arrest and a senescent phenotype characterized by enlarged morphology and secretory activity.
Antioxidants 14 01166 g005
Figure 6. Keap1–Nrf2–ARE signaling pathway and downstream antioxidant gene activation. Under basal conditions, Nrf2 is bound to Keap1, a redox-sensitive adaptor protein that recruits Cullin 3 (Cul3)–based E3 ubiquitin ligase complexes, leading to ubiquitination and proteasomal degradation of Nrf2. Upon exposure to ROS or electrophilic stress, critical cysteine residues of Keap1 undergo covalent modifications, preventing Nrf2 degradation. Stabilized Nrf2 accumulates in the cytoplasm, translocates into the nucleus, and heterodimerizes with small Maf proteins to bind antioxidant response elements (AREs) in DNA. This induces the transcription of a wide range of cytoprotective genes, including glutathione peroxidase 2 (GPx2), peroxiredoxin 1 (Prdx1), thioredoxin 1 (Txn1), sulfiredoxin 1 (Srx1), glutamate–cysteine ligase catalytic subunit (GCLC) and modifier subunit (GCLM), glutathione S-transferases (GSTs), NAD(P)H quinone dehydrogenase 1 (NQO1), cytochrome P450 2A6 (CYP2A6), multidrug resistance–associated protein 2 (MRP2), and heme oxygenase 1 (HO-1). Collectively, these enzymes mitigate oxidative damage, detoxify xenobiotics, and maintain redox homeostasis.
Figure 6. Keap1–Nrf2–ARE signaling pathway and downstream antioxidant gene activation. Under basal conditions, Nrf2 is bound to Keap1, a redox-sensitive adaptor protein that recruits Cullin 3 (Cul3)–based E3 ubiquitin ligase complexes, leading to ubiquitination and proteasomal degradation of Nrf2. Upon exposure to ROS or electrophilic stress, critical cysteine residues of Keap1 undergo covalent modifications, preventing Nrf2 degradation. Stabilized Nrf2 accumulates in the cytoplasm, translocates into the nucleus, and heterodimerizes with small Maf proteins to bind antioxidant response elements (AREs) in DNA. This induces the transcription of a wide range of cytoprotective genes, including glutathione peroxidase 2 (GPx2), peroxiredoxin 1 (Prdx1), thioredoxin 1 (Txn1), sulfiredoxin 1 (Srx1), glutamate–cysteine ligase catalytic subunit (GCLC) and modifier subunit (GCLM), glutathione S-transferases (GSTs), NAD(P)H quinone dehydrogenase 1 (NQO1), cytochrome P450 2A6 (CYP2A6), multidrug resistance–associated protein 2 (MRP2), and heme oxygenase 1 (HO-1). Collectively, these enzymes mitigate oxidative damage, detoxify xenobiotics, and maintain redox homeostasis.
Antioxidants 14 01166 g006
Figure 7. Flowchart of study selection process.
Figure 7. Flowchart of study selection process.
Antioxidants 14 01166 g007
Table 1. Inflammation and Neuropathic Pain—Core Mechanisms and Implications.
Table 1. Inflammation and Neuropathic Pain—Core Mechanisms and Implications.
Mechanistic Node/SiteCore Mediators and SignalsClinical Significance/Therapeutic ImplicationReferences
Peripheral immune activation (lesion/DRG)DAMPs, eicosanoids; endothelial activation → leukocyte extravasation; ROS/RNS; cytokines/chemokinesEarly peripheral anti-inflammatory control to curb nociceptor hyperexcitability and prevent central amplification.[27,47,48,49,50]
Danger sensing and inflammatory transcriptionTLR4–MyD88/TRIF → NF-κB/MAPK; IL-6–gp130/JAK–STATConsider TLR4/NF-κB/JAK–STAT modulators to blunt sustained pro-inflammatory programs and astrocyte reactivity.[51,52,53,54]
Prostaglandin coupling to excitabilityCOX-2 → PGE2 → EP → cAMP–PKA/PKC; ion-channel phosphorylationUse COX-2 inhibitors and/or EP antagonists to decouple cytokine surges from synaptic potentiation and hyperexcitability.[53,54]
Peripheral transduction plasticityTNF/IL-1β ↑ NaV, ↓ K+; PGE2/oxidized lipids → TRPV1 sensitization, TRPA1 activationNaV/TRP-directed strategies plus antioxidant/electrophile-quenching to reduce repetitive firing and thermal/chemical hyperalgesia.[56,57,58]
Central glial-driven sensitizationBarrier compromise; microglia NOX2/p38 → IL-1β/TNF/BDNF → TrkB → KCC2 ↓; astrocytes STAT3/NF-κB → COX-2/PGE2 + chemokinesTime-staged approach: microglia targets (P2X4/P2X7, p38, NOX2) early; astrocyte/COX-2/EP/chemokine blockade for chronic maintenance.[61,62,63,64,65,66,82,83,84,85,86,87,88,89,90,91]
Oxidative–inflammatory loop and resolution leversNOX/iNOS → ROS/RNS; cGAS–STING and NLRP3 → IL-1β maturation; SPMs/NPD1 → GPR37; efferocytosis; M2 programmingCombine NOX/iNOS or cGAS–STING/NLRP3 inhibition with pro-resolving mediators and macrophage reprogramming to terminate inflammation.[32,67,68,69,71,72,78,79,80]
Table 2. Oxidative Stress in Neuropathic Pain—Key Nodes, Effects, and Therapeutic Implications.
Table 2. Oxidative Stress in Neuropathic Pain—Key Nodes, Effects, and Therapeutic Implications.
Mechanism/SourceKey ROS/RNS or NodeEffect on Pain CircuitryClinical Significance/
Therapeutic Implication
References
Mitochondrial ETC leak (Complex I/III; Fenton chemistry)O2 → H2O2 → •OHOxidative injury, energy failure, ectopic firingLimit reverse electron transport; optimize substrates; manage iron; use matrix-targeted antioxidants[99,100,101,120,121,122,123,124]
NOX/iNOS pathwaysNOX1/2/4, DUOX1/2; NO• → ONOOProtein nitration/DNA damage; mitochondrial enzyme inactivationNOX/iNOS modulation; peroxynitrite handling[103,104,105,106,123,124,125,126,127,128]
Redox gating of channels (TRPA1/TRPV1; NaV/Kv/HCN)4-HNE, isoprostanes; S-nitrosylation/oxidationLowered thresholds; repetitive firing; increased gainTRP modulators; NaV-targeted agents; antioxidant strategies[131,132,133,134,135]
Microglial redox signalingNOX2 → p38/JNK/NF-κB; BDNF → TrkB → ↓KCC2Disinhibition; central sensitization/LTP-like potentiationEarly microglia-targeted therapy; TrkB/KCC2 axis (investigational)[63,64,65,138,139,140]
Astrocyte maintenanceSTAT3/NF-κB; COX-2 → PGE2 → EPGlutamate dysregulation; persistent synaptic gainCOX-2 or EP antagonism; astrocyte-focused modulation[137]
Innate immune couplingmtDNA/oxidized DAMPs → cGAS–STING; NLRP3 inflammasomeIL-1β maturation; sustained cytokine excessTarget cGAS–STING/NLRP3 (investigational)[141,142]
Mitochondrial integrityCardiolipin oxidation; impaired biogenesis/mitophagySupercomplex destabilization; ROS-rich, depolarized mitochondriaCardiolipin-stabilizing peptides; support mitophagy/biogenesis[147,148,149,150,151]
Table 3. Cellular Senescence in Neuropathic Pain—Triggers, Pathways, Markers, Consequences, and Therapeutic Implications.
Table 3. Cellular Senescence in Neuropathic Pain—Triggers, Pathways, Markers, Consequences, and Therapeutic Implications.
ThemeTriggers/DriversCore Pathways and MarkersKey Effects in Pain BiologyClinical Significance/Therapeutic ImplicationReferences
Definition and typesReplicative, oncogene-, therapy-, stress-induced; MiDAS; immune-drivenp53–p21, p16–Rb; SA-β-gal; SASPContext-dependent: repair vs. chronic inflammationUse multi-marker panels; recognize MiDAS/SASP heterogeneity[95,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172]
DDR gatekeepingTelomere attrition; DNA double-strand breaksγH2AX, MDC1, 53BP1; ATM/ATR → CHK1/2 → p53 → p21; SAHFIrreversible arrest; checkpoint enforcementDDR biomarkers for identification; potential checkpoint modulation[163,164,165,166,167,168,169,173,174,201]
Oxidative stress–mitochondria loopROS/RNS overload; mtDNA damage; ETC leak; peroxisomal insufficiencyComplex I/III ROS; catalase decline; AMPK; ATM–Akt–mTORC1–PGC-1α; glycolytic shiftROS amplification, energy failure, Ca2+ dysregulationAntioxidants/ROS scavenging; mito-support (biogenesis, mitophagy, fission–fusion balance)[95,148,184,185,186,187,188,189,190,191,192,193,194,195,196]
SASP control and propagationPersistent stress/DDR; redox crosstalkNF-κB, p38 MAPK, mTOR, C/EBPβ; cGAS–STING; IL-1β/IL-6/IL-8, chemokines, MMPs, EVsParacrine senescence; tissue remodeling; chronic inflammationTarget SASP nodes (NF-κB/p38/mTOR; cGAS–STING); time-sensitive modulation[95,161,165,197,198]
Neural and glial senescence (PoMiCS)Axotomy, chemo, oxidative/genotoxic stress; agingNeurons/glia with p21/p16, SA-β-gal, γH2AX; TRPV1+ DRG IL-6 SASP; astrocyte/microglia SASPAltered excitability, disinhibition, central sensitizationStage-aware therapy: nociceptor source control + glial signal control (e.g., IL-6/JAK–STAT, COX-2/EP, chemokines)[213,214,215,216,217,218,219,220,221,222,223,224,225]
Table 4. Nrf2 as a Central Integrator of Oxidative Stress, Inflammation, Senescence, and Neuropathic Pain.
Table 4. Nrf2 as a Central Integrator of Oxidative Stress, Inflammation, Senescence, and Neuropathic Pain.
Module/AxisCore Mechanisms and EvidenceClinical Significance/Therapeutic ImplicationReferences
Keap1–Nrf2–ARE coreOxidant/electrophile modification of Keap1 cysteines → Nrf2 stabilization → nuclear translocation; sMaf heterodimers bind ARE; induction of HO-1, NQO1, GSTs, GSH biosynthesis enzymes; HO-1 generates CO/bilirubin (anti-inflammatory/antioxidant).Boosts cellular redox buffering, preserves mitochondria, limits excitotoxicity; HO-1/CO axis provides analgesic anti-inflammatory effects.[228,235,236,237]
Anti-inflammatory crosstalkNrf2 restrains NF-κB; lowers CAMs (VCAM-1, E-selectin); suppresses iNOS, COX-2, MMPs.Reduces leukocyte recruitment and neuroinflammation; complements standard analgesics; candidate for disease-modifying strategies.[63,237,238,259,260]
Neuro-glial modulationNrf2 decreases microglial activation and astrocyte reactivity; dampens TNF/IL-1β; limits glutamate-driven excitability; influences ion channels/neurotransmission.Attenuates central sensitization; pairs well with glia-targeted and synaptic-modulating therapies.[98,238,239,240,241,242,243,244,245,246,247,248,249]
Evidence in pain models and activatorsNrf2/HO-1 induction yields antinociception; agents: cobalt protoporphyrin, curcumin, diosmetin, sulforaphane; bardoxolone, omaveloxolone; electroacupuncture/ARBs/flavonoids act via Nrf2; Nrf2 inhibition worsens pain/ROS.Pharmacologic Nrf2 activation is analgesic across models; translational potential for chemotherapy- and injury-induced neuropathies.[227,250,251,252,253,254,255,256,261,262,263,264]
Integrative signaling nodesp62 competes for Keap1 (autophagy link); PI3K/Akt inhibits GSK-3β → stabilizes Nrf2; cooperation with RXRα.Opportunity to combine Nrf2 activation with autophagy support or PI3K/Akt tuning; context-aware dosing.[235,240,241]
Nrf2–mTOR and senescenceReciprocal Nrf2 ↔ mTOR reinforcement (Nrf2 ↑ mTOR/RagD; mTOR ↓ β-TrCP-mediated Nrf2 degradation); impacts SASP and persistence of senescent phenotypes.Stage-specific use: activate Nrf2 to quell ROS/inflammation, but avoid chronic overactivation that may stabilize senescence; consider mTOR modulation.[232,239,266,267]
Localization and dynamics in pain axisNrf2/HO-1 expressed in spinal cord, DRG, peripheral nerves; chemo-neuropathy vs. CCI show compartment- and stage-dependent shifts.Target tissues/stages selectively; consider biomarkers of Nrf2 activity for patient stratification.[63,254,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277]
Risks of chronic/constitutive activationTumor promotion, therapy resistance; fibrosis, autoimmunity; metabolic dysregulation (insulin resistance, steatosis).Careful patient selection/monitoring; caution in active malignancy; balance benefit vs. risk.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Stojanovic, B.; Milivojcevic Bevc, I.; Dimitrijevic Stojanovic, M.; Stojanovic, B.S.; Lazarevic, T.; Spasic, M.; Petrovic, M.; Stefanovic, I.; Markovic, M.; Nesic, J.; et al. Oxidative Stress, Inflammation, and Cellular Senescence in Neuropathic Pain: Mechanistic Crosstalk. Antioxidants 2025, 14, 1166. https://doi.org/10.3390/antiox14101166

AMA Style

Stojanovic B, Milivojcevic Bevc I, Dimitrijevic Stojanovic M, Stojanovic BS, Lazarevic T, Spasic M, Petrovic M, Stefanovic I, Markovic M, Nesic J, et al. Oxidative Stress, Inflammation, and Cellular Senescence in Neuropathic Pain: Mechanistic Crosstalk. Antioxidants. 2025; 14(10):1166. https://doi.org/10.3390/antiox14101166

Chicago/Turabian Style

Stojanovic, Bojan, Ivana Milivojcevic Bevc, Milica Dimitrijevic Stojanovic, Bojana S. Stojanovic, Tatjana Lazarevic, Marko Spasic, Marko Petrovic, Ivana Stefanovic, Marina Markovic, Jelena Nesic, and et al. 2025. "Oxidative Stress, Inflammation, and Cellular Senescence in Neuropathic Pain: Mechanistic Crosstalk" Antioxidants 14, no. 10: 1166. https://doi.org/10.3390/antiox14101166

APA Style

Stojanovic, B., Milivojcevic Bevc, I., Dimitrijevic Stojanovic, M., Stojanovic, B. S., Lazarevic, T., Spasic, M., Petrovic, M., Stefanovic, I., Markovic, M., Nesic, J., Jovanovic, D., Peulic, M., Azanjac Arsic, A., Lukovic, A., Mirkovic, N., Eric, S., & Zornic, N. (2025). Oxidative Stress, Inflammation, and Cellular Senescence in Neuropathic Pain: Mechanistic Crosstalk. Antioxidants, 14(10), 1166. https://doi.org/10.3390/antiox14101166

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop