Next Article in Journal
Personalized Management of Patients with Proliferative Diabetic Vitreoretinopathy
Previous Article in Journal
Congenital Pulmonary Airway Malformation in Preterm Infants: A Case Report and Review of the Literature
Previous Article in Special Issue
Pilot Study of Home-Based Virtual Reality Fitness Training in Post-Discharge Rehabilitation for Patients with Spinal Cord Injury: A Randomized Double-Blind Multicenter Trial
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chemotherapy-Induced Peripheral Neuropathy: A Recent Update on Pathophysiology and Treatment

1
Department of Pharmaceutical and Administrative Sciences, Presbyterian College School of Pharmacy, Clinton, SC 29325, USA
2
Department of Pharmaceutical and Graduate Life Sciences, College of Health Sciences, Nursing, and Pharmacy, Manchester University, Fort Wayne, IN 46845, USA
3
Department of Pharmacy Practice, College of Health Sciences, Nursing, and Pharmacy, Manchester University, Fort Wayne, IN 46845, USA
4
Department of Biology, Presbyterian College, Clinton, SC 29325, USA
5
Department of Pharmaceutical and Clinical Sciences, College of Pharmacy and Health Sciences, Campbell University, Buies Creek, NC 27506, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to the paper.
Life 2024, 14(8), 991; https://doi.org/10.3390/life14080991
Submission received: 20 June 2024 / Revised: 29 July 2024 / Accepted: 7 August 2024 / Published: 9 August 2024
(This article belongs to the Special Issue Conservative Management of Chronic Disease)

Abstract

:
Chemotherapy-induced peripheral neuropathy (CIPN) is a major long-lasting side effect of some chemotherapy drugs, which threatens cancer survival rate. CIPN mostly affects sensory neurons and occasionally motor neurons, causing numbness, tingling, discomfort, and burning pain in the upper and lower extremities. The pathophysiology of CIPN is not completely understood; however, it is believed that chemotherapies induce peripheral neuropathy via directly damaging mitochondria, impairing the function of ion channels, triggering immunological mechanisms, and disrupting microtubules. The treatment of CIPN is a medical challenge, and there are no approved pharmacological options. Currently, duloxetine and other antidepressants, antioxidant, anti-inflammatory, and ion-channel targeted therapies are commonly used in clinics to relieve the symptoms of CIPN. Several other types of drugs, such as cannabinoids, sigma−1 receptor antagonists, and nicotinamides ribose, are being evaluated in preclinical and clinical studies. This paper summarizes the information related to the physiology of CIPN and medicines that could be used for treating this condition.

1. Introduction

Chemotherapy-induced peripheral neuropathy (CIPN) is defined as an injury of the somatosensory nervous system after treatment with chemotherapy [1]. There are two types of sensory abnormalities contributing to the development of motor and autonomic changes in CIPN: these include abnormal sensory discrimination of touch, vibration, and thermal information and a tingling or burning sensation also referred to as mechanical allodynia [2]. The severity and the duration of CIPN depends on, treatment agent, chemotherapy dose, duration, a combination of neurotoxic drugs, and the presence of preexisting diseases such as diabetes or chronic kidney disease [3]. In most cases, the sensation happens in the hands and feet known as “gloves and stocking” distribution [4]. These symptoms can happen with just one high single dose or after accumulation of chemotherapy drugs [1]. Symptoms usually stabilize during treatment and relieve with time after termination. However, about one-third of patients experience symptoms after 6 months of treatment cessation [5]. In some cases, the chemotherapy drugs can persist in the nerve axon after treatment and worsen the neuropathic symptoms, and this is known as the “Coasting Phenomena” [3,6].
CIPN is a serious problem that ranks as a major long-lasting adverse effect related to chemotherapy medications, such as platinum, ixabepilone, and thalidomide [7,8]. It is estimated that 30–40% of patients undergoing chemotherapy will experience CIPN each year and can lead to either premature treatment termination or chemotherapy dose reduction that influences the treatment efficacy and survival rate [9,10]. The prevalence range of CIPN is from 19% to more than 85%. The severity of CIPN depends on the length and agent used with some chemotherapeutic agents having the highest rates such as platinum agents, taxanes, vinca alkaloids, proteasome inhibitors, and immunomodulatory drugs [11]. Seventy percent of patients who received paclitaxel and 90% of those who received oxaliplatin develop the symptoms of CIPN [12].
The treatment of CIPN has been a medical challenge. The foremost challenge is the lack of agents recommended for CIPN prevention and management in the American Society for Clinical Oncology clinical practice (ASCO) guidelines [13]. Current recommendations suggest a focus on symptoms affecting quality of life by using pharmacological interventions for pain and physiotherapy solutions like exercise, acupuncture, massage therapy for functional impairment, and in some cases provision relief [14]. However, despite current efforts, CIPN is still not well managed in the majority of patients and has a negative impact on quality of life [13]. Duloxetine is the only off-label FDA-approved drug that has shown efficacy in treating CIPN when compared to antidepressants and anticonvulsants [13,15]. In addition, CIPN also increases the economic burden, while decreasing quality of life (QOL) [13]. A study carried out where both physical and mental health were evaluated to measure QOL showed that a higher symptom burden is associated with a lower QOL [16]. On cost, a 2022 review of studies carried out on CIPN treatment indicated that the monthly drug cost of treating CIPN has a range of USD 15 to USD 1425 [17]. Further, the symptoms of CIPN reduce work efficiency resulting in loss of job, which adds to the economic burden [2].
Treating CIPN is a challenge due in part to the prioritization of life-threatening chemotherapy side effects over CIPN, discussions of side effects are more centered around acute issues like nausea, vomiting, hair loss, and infections [18]. Even though there are advances in understanding the biological mechanisms of chemotherapy-induced peripheral neuropathy, few prevention or treatment options exist [10,15]. Therefore, discovering efficient approaches for CIPN management is highly demanding.
In this paper, we will discuss the pathology of CIPN and summarize the current pharmacological therapy utilized in CIPN. Further, the medicines under clinical and preclinical investigation will be discussed. We also summarize the recent findings of the mechanism of action, safety, and efficacy of drugs for treating CIPN. In addition, this paper will list some limitations and future directions related to these options based on data from clinical trials.

2. Pathophysiology of Chemotherapy-Induced Peripheral Neuropathy

Chemotherapy medications induce peripheral neuropathy via different mechanisms: changes in neuronal cytoskeleton, directly damaging neuron/DRG mitochondria, impairing the function of ion channels, triggering immunological mechanisms via glial cell activation, and disrupting microtubules [4].

2.1. Cytoskeleton Changes

A functional cytoskeleton is required to maintain the function and integrity of the nervous system [19]. Chemotherapy medications promote major modifications in the neuronal cytoskeleton, resulting in the fragmentation of axons in sensory neurons leading to changes in the microtubule network [19,20]. These changes compromise the structural integrity of sensory neurons, resulting in impaired axonal transport of mitochondria and mRNA [21]. For example, paclitaxel and vincristine can cross the blood-nerve-barrier and bind to beta-tubulin of sensory nerve fibers where they affect the structure of the cytoskeleton of healthy peripheral neurons, such as the primary afferent neurons of the dorsal root ganglia [19,20].

2.2. Oxidative Stress

Mitochondria are membrane-bound cell organelles responsible for producing energy through the oxidative phosphorylation process [22]. Chemotherapy can cause mitochondrial dysfunction through different mechanisms [23]. Platinum binds to mitochondrial DNA and impairs replication, transcription, and protein synthesis, resulting in respiratory chain damage, reduction of cellular metabolism, and increased oxidative stress which leads to damage to peripheral nerves [24,25,26,27,28,29,30,31]. Taxanes can alter mitochondrial membrane permeability pores leading to the release of calcium, which further activates calpain that catalyzes protein degeneration including neuronal calcium sensor 1 (NCS1). These changes lead to mitochondrial depolarization, impaired energy production, and neuronal dysfunction which plays a role in the induction of neuropathic pain [32].

2.3. Ion Channel

Altered expression of the voltage-gated sodium channel (NaV), voltage-gated potassium channel (Kv), and voltage-gated calcium channel (CaV) leads to a change in neuronal excitability [33,34]. NaV is essential to initiate and propagate action potential in neurons [25]. Chemotherapeutic drugs can increase the expression of NaV in nociceptive neurons increasing action potential [34]. The amplification of action potential in nociceptive neurons results in hyperexcitability and spontaneous firing of neurons which results in increased pain sensitivity and neuropathic pain [35]. Treatment with paclitaxel increases the expression of NaV in DRG causing increased action potential firing resulting in hyperexcitability [36,37]. Significant downregulation of voltage-gated potassium channels has been found in patients after oxaliplatin treatment, leading to increased membrane excitability in neurons [38,39]. Chemotherapy can also upregulate voltage-gated calcium channels that change peripheral neurons’ action potential threshold, causing hyperexcitability [40,41,42].

2.4. Triggering Immunological Responses

Previous studies have demonstrated a strong correlation between CIPN and increased levels of pro-inflammatory cytokines. This is because cytokine signaling, crucial for neuroinflammation and sensitization in the sensory nervous system, plays a key role in chemotherapy-induced peripheral neuropathy (CIPN) [33]. CIPN is characterized by increased levels of pro-inflammatory cytokines (TNF-α, IL-1β, IL-6) and reduced levels of the anti-inflammatory cytokine IL-10 [36,43]. Oxaliplatin can activate astrocytes leading to the activation of the pro-inflammatory cascade and production of pro-inflammatory cytokines such as IL-Iβ, TNF-α, and IL-6 [44,45,46]. Taxanes can bind to toll-like receptor 4 (TLR-4) triggering activation of NF-κB [42].

2.5. Disruption of Microtubules

Taxanes and Vinca Alkaloids also induce peripheral neuropathy by interrupting tubulin polymerization. After binding to tubulin, Taxanes and Vinca Alkaloids cause microtubule dysfunction leading to alteration of axonal synaptic vesicle transportation as well as interfering with the regeneration and remodeling of axons [47,48,49,50].

3. Current Approach for CIPN Management

Based on the most up to date ASCO guidelines, there is no agent recommended for the prevention of CIPN [13]. However, the ASCO does recommend that patients with CIPN should receive reduced or delayed doses and substitute or terminate the chemo agent [13]. Duloxetine (brand name Cymbalta), an agent off-label used to treat CIPN, is the only agent with moderate efficacy [13,51]. Duloxetine inhibits serotonin and norepinephrine reuptakes, which increases the availability of the key neurotransmitters to activate the descending pathway [52]. Duloxetine could also reduce inflammation and nerve injury by inhibiting the activation of p38 and NF-kB [53]. Duloxetine has been tested in many clinical trials and has been approved to reduce pain compared with placebo and other classes of medications [29,41].
Several complementary and alternative medicines have been tested for treating CIPN. Nonpharmacological approaches, including acupuncture, exercise, mindfulness practices, yoga, meditation, and touch therapies like acupressure, reflexology, and massage, have been found to reduce chemotherapy-related symptoms and improve quality of life [54]. In addition, some nutrients and Chinese herbal medicines have shown potential therapeutic effects in patients with neuropathic pain [14].
There are major limitations in the current treatment of CIPN. A major limitation of duloxetine use is that insurance companies recommend using either pregabalin or gabapentin before the usage of duloxetine even though that is contradictory to the recommendation of ASCO guidelines [13]. By far, none of the complementary and alternative approaches has shown promising results for treating CIPN [54]. In addition, there is a lack of guidelines for using nutrients and herbs for the treatment of CIPN in clinics.

4. Medications under Clinical Trials

With no effective therapies to treat CIPN, several clinical trials have been carried out and continue to be carried out to find a clinical solution (Table 1). Therapies used for other disease states have been studied to ascertain their effectiveness in alleviating CIPN. However, we only include the clinical trials of which the results have been published.

4.1. Serotonin and Norepinephrine Reuptake Inhibitors

Venlafaxine is a second-generation tricyclic antidepressant (TCA) that works by inhibiting neuronal serotonin activity, and norepinephrine and serotonin reuptake [55]. It is a good alternative for patients who cannot afford duloxetine [56]. A clinical trial conducted in 2018 found that venlafaxine effectively reduced the grade of cranial, motor, sensory, and neuropathic pain in the patients who had CIPN, but with less efficacy compared to Duloxetine [57]. In a phase III randomized double-blinded control trial, venlafaxine was effective in relieving acute neurotoxicity and improving functional status in patients who suffered from oxaliplatin-induced neuropathy [58]. However, in a pilot trial, Venlafaxine was not effective in preventing either acute or chronic neuropathy symptoms, such as throat discomfort and discomfort swallowing cold liquids, induced by oxaliplatin treatment [59]. Original guideline commentary for venlafaxine has shown its use as a preventative agent, however, longer follow-up data do not support this use [58].
Amitriptyline is a tricyclic TCA medication being tested in clinical trials [60]. One study showed that systemic treatment with amitriptyline was unable to relieve CIPN symptoms, including tingling, numbness, impaired sensory function, and pain in hands and feet [61]. In contrast, another clinical trial found that topical amitriptyline alone or with baclofen and ketamine effectively reduced the neuropathic pain associated with CIPN in patients who had received vinca alkaloids, oxaliplatin, cisplatin, taxanes, thalidomide or other neurotoxic agents [62,63]. Both amitriptyline and venlafaxine are not FDA approved nor have definite doses for treating CIPN [13,58,62]. In addition, they have more side effects compared to duloxetine. Thus, more trials need to be carried out to provide guidance on optimal doses with fewer side effects to treat CIPN [45].

4.2. Ion Channel Targeted Therapy

Studies have shown that ion channel targeted therapies, such as lidocaine, gabapentin, and pregabalin, have also been successful in reducing CIPN [9]. Lidocaine is an antiarrhythmic drug that can non-selectively block sodium, potassium, and calcium channels [64]. Systemic treatment with lidocaine via intravenous injection was effective in pain reduction with an analgesic effect that persisted for an average of 23 days in patients [65]. This is further supported by a case where neuropathy symptoms were completely relieved by intravenous lidocaine infusion [66]. A recent randomized controlled study also found that lidocaine infusion showed a similar effect as duloxetine on reducing the incidence and severity of taxane-induced peripheral neuropathy [67].
Gabapentin and pregabalin are anticonvulsant agents that block presynaptic voltage-gated calcium channels and down-regulate excitatory [68]. Several clinical trials have evaluated the efficacy of gabapentin and pregabalin in CIPN prevention and treatment. A phase 3 randomized controlled trial found that treatment with gabapentin for 6 weeks failed to relieve the CIPN symptoms [69]. Another randomized trial also had similar results. Gabapentin (20 mg/kg/day) increased the pain scores and opioid consumption in pediatric patients with vincristine-induced neuropathy [70]. Treatment with pregabalin (75 mg) twice daily showed a trend to reduce numbness, but not any other paclitaxel-induced neuropathy symptoms [71]. More trials are needed to definitively provide indications of these drugs’ use in CIPN treatment [69].

4.3. Anti-Inflammatory Medications

Typically, these drugs are used in other disease states but also exhibit anti-inflammatory effects that have shown efficacy in decreasing CIPN. Metformin, a drug commonly used for treating diabetes, has shown anti-inflammatory effects on endothelial cells by inhibiting the production of IL-6 and TNF-α thus reducing the phosphorylation of mitogen-activated protein kinase (MAPK) [68]. Additionally, Metformin activates adenosine monophosphate-activated protein kinase (AMPK) which could inhibit the mammalian target of the rapamycin (mTOR) pathway causing nociception blockage [72]. A study was conducted using cisplatin and Paclitaxel, which are drugs known to induce mechanical allodynia. One arm had co-administration of Metformin with cisplatin and the other arm administered Metformin with Paclitaxel [45,72,73]. In a randomized controlled study, metformin 500 mg three times a day was given to patients receiving oxaliplatin in the FOLFOX-4 regimen for colorectal cancer. Metformin offered protection against oxaliplatin-induced chronic peripheral sensory neuropathy [74].
Minocycline is an antibiotic that inhibits the production of pro-inflammatory cytokines in monocytes and microglia to reduce mechanical sensitivity [75]. In the Academic and Community Cancer Research United (ACCRU) pilot study, Minocycline reduced the daily average pain score but no prevention of CIPN was found. However, when dosed based on weight, minocycline prevented chemotherapy-induced neuropathy in mice receiving paclitaxel [76,77]. In contrast, in a phase II randomized clinical trial where an oxaliplatin chemotherapy regimen was used, minocycline did not relieve fatigue or numbness/tingling symptoms and failed to reduce pro-inflammatory cytokines compared to placebo [78].

4.4. Antioxidant Medications

It has been shown that the antineoplastic agent oxaliplatin does induce peripheral neurotoxicity. One of the ways suggested on how this neurotoxicity occurs is oxidative stress, where there’s an increase of reactive oxygen species and reactive nitrogen species creating an imbalance between their production and removal. Mitochondrial superoxide dismutase (MnSOD) is what keeps the balance on these species. Mangafodipir, a contrast agent, was discovered to have mitochondrial superoxide dismutase mimetic activity and was therefore tested as a cytoprotectant [79].
In the PLIANT phase II placebo-controlled study, patients treated with oxaliplatin were given calmangafodipir (Pledox) which is derived from mangafodipir [80,81,82,83,84]. It was observed that there was a delay of onset and a reduction in intensity of the CIPN symptoms, in particular the cold allodynia symptoms, with calamangafodipir at a dose of 5 µmol/kg [83].
Amifostine is a potent free radical scavenger pro-drug and has been shown to protect healthy tissues during chemotherapy and radiation [85]. The active metabolite (WR-1065) demonstrated the prevention of oxaliplatin-induced neurotoxicity [86]. This protection is selective for non-tumor tissues [10,13]. However, in a phase II clinical trial, where cisplatin and 3-hour paclitaxel were the chemotherapy agents used [87]. Amifostine had a diminished effect on preventing CIPN [83]. These two trials demonstrate that amifostine is a strategy to be used specifically in oxaliplatin-induced neurotoxicity, however, it was not so in other platinum compounds like cisplatin and paclitaxel [83,87].

4.5. Sigma-1 Receptors (S1R)

Sigma-1 receptors (S1R) play an important role in modulating several types of ion channels such as NaV 1.2, NaV 1.4, KV 1.2, KV 1.3, and KV 1.4 [88,89]. They also can modulate intramitochondrial calcium homeostasis and trafficking of functional transient receptor potential ankyrin 1 (TRPA1), which play a critical role in the development of CIPN [90,91]. MR309 is a selective S1R antagonist that has shown efficacy in reducing allodynia and cold hypersensitivity [92]. A phase 1 trial found that MR309 has shown good safety and tolerability in healthy people [93]. In two other clinical studies, MR309 showed a significant reduction in cold allodynia and hyperexcitability motor symptoms compared to placebo [88,89].

4.6. Cannabinoids

The endocannabinoid system is an endogenous system responsible for pain perception modulation and has attracted interest in the management of CIPN [94]. It consists of two types of receptors, CB1 dominant in the brain and CB2 dominant in peripheral immune systems and central nervous systems [95]. Both CB1 and CB2 are G-protein coupled receptors and can modulate the activities of numerous intracellular signaling pathways, such as MAPK, phosphoinositide 3-kinase (PI3K)/Akt, Nrf2, and Ca2+-regulated signaling cascades [96]. Endocannabinoid system homeostasis is maintained via transporters that control the transport of anandamide (AEA) and 2-arachidonoylglycerol (2-AG) from the cells to synapse and vice versa [97]. The activity of the endocannabinoid system is also controlled by enzymes, such as fatty acid amino hydrolase, cyclooxygenase-2 (COX-2), and N-acylethanolamine-hydrolyzing acid amidase, which are involved in the degradation of the endocannabinoid [98].
The safety, efficacy, and pharmacokinetics of cannabinoids have been studied in several clinical trials [99,100,101,102,103,104,105]. A double-blind placebo-controlled study determined that oral mucosal spray containing cannabinoids had no significant difference in treating CIPN [106]. However, the five “responders” had 2.6 of 11-point NRS-PI lower than the placebo group [103]. Treatment with CBD oil (300 mg/daily) had lower scores on cold sensitivity to touch, discomfort swallowing cold liquids, and throat discomfort in patients with CIPN [105]. A case series reported that topical use of CBD alone or with THC reduced neuropathic pain in patients with CIPN [107]. Numerous clinical trials are ongoing to test the effect of cannabinoids on treating and preventing CIPN.
Table 1. Medications under clinical trials for the treatment of CIPN.
Table 1. Medications under clinical trials for the treatment of CIPN.
InterventionChemotherapySample SizeDoseResultsReference
Amitriptyline 44 1 g of 10% amitriptyline cream twice a day Topical amitriptyline was effective in reducing VAS pain score[56]
AmitriptylineTaxane, vinca alkaloid, or platinum114 25 mg daily up to 100 mg daily Amitriptyline was not effective in CIPN prevention[61]
Amitriptyline and ketamineTaxane462 Apply 4 g of 2% ketamine plus 4% amitriptyline (KA) cream BID No difference in 6-week pain score, numbness, and tingling[63]
Amitriptyline Baclofen
ketamine
208 baclofen 10 mg, amitriptyline HCL 40 mg, and ketamine 20 mg BAK gel was statistically significant in motor neuropathy improvement but not statistically significant in sensory neuropathy[62]
AmifostineCisplatin or paclitaxel27 amifostine 740 mg/m2Not effective in protection against CIPN[87]
CalmangafodipirOxaliplatin173 calmangafodipir 2 µmol/kg or 5 µmol/kg Calmangafodipir can prevent CIPN during the 1st 6 months[83]
CalmangafodipirOxaliplatin592 Calmangafodipir 2 µmol/kg, or 5 µmol/kg, or placebo Month 9 of treatment data shows 54% with moderate-severe CIPN and higher percentage(s) of hypersensitivity reactions[84]
CannabidiolCarboplatin and paclitaxel, or
capecitabine and oxaliplatin
54 150 mg CBD oil twice daily (300 mg/daily) for 8 days Reduce cold sensitivity to touch, discomfort swallowing cold liquids, and throat discomfort[105]
Cannabidiol 40 Topical use for 2 weeks No improvement on painful established CIPN[108]
Cannabinoid extract (Nabiximols)Paclitaxel, vincristine, or cisplatin16 Two to twelve sprays per day for a week An average decrease of 2.6 on an 11-point NRS-PI in five “responders”[106]
CapsaicinOxaliplatin18High dose 8% patchEffective in treating pain associated with CIPN [109]
l-carnosineOxaliplatin65L-carnosine 500 mg dailySignificant reduction in NF-κB and TNFα; Neuroprotective[110]
Duloxetine 231 30 mg QD 7D, then 30 mg BID Use for 5 weeks resulted in greater overall pain reduction[51]
DuloxetineTaxane4730 mg daily in the first week following the injection of paclitaxel and 60 mg during the second week Reduces the scores of neuropathies, but not paresthesia, numbness, cold sensitivity, and other nerve conduction velocity (NCV) values[111]
GabapentinTaxane,
platinum, or
vinca alkaloid
115300 mg up to 2700 mg dailyNot effective in controlling CIPN symptoms[69]
GabapentinVincristine4920 mg/kg/day Gabapentin was associated with more opioid consumption and high pain score compared to placebo[70]
GlutamineVincristine56Glutamine 6 g/m2 BIDWell tolerated
Sensory function and QOL improvement
[112]
Glutathione Paclitaxel and carboplatin185 1.5 g/m2 IV over 15 min immediately before chemotherapy No significant difference in peripheral neuropathy [113]
Glutathione and Mecobalamin 158 2.4 g of glutathione IV once daily 2–3 days before chemotherapy, plus 500 µg mecobalamin IV once every other day Significant reduction in incidence and severity of CIPN [114]
Huangqi Guizhi Wuwu Decoction (HGWD)Paclitaxel92Washing limbs with HGWD for 20 min twice a day for consecutive 14 daysLargely reduces the CIPN sensory scores, but not the autonomic scores [115]
LafutidineCarboplatin or paclitaxel1810 mg twice dailyShows a trend to reduce neurotoxicity[116]
Lamotrigine 131Target dose 300 mg/dayLamotrigine is not effective as CIPN treatment[117]
Lidocaine 91.5 mg/kg for 10 min followed by 1.5 mg/kg/h over 5 hPain reduction that persists for 23 days average[65]
LidocaineTaxane602 mg/kg with saline infusion 40 min prior to taxane therapyDecreased the incidence and severity of taxane-induced peripheral neuropathy[67]
LithiumTaxane 37300 mg daily for 5 daysNot effective in CIPN prevention[118]
Melissa officinalis (MO)Cisplatin, oxaliplatin), vincristine, bortezomib, or taxanes80500 mg 2 times a day for 3 monthsMO reduces pain and diarrhea based on the scores of EORTC QLQ-C30 (Integrated System for Quality of Life Assessment)[119]
MetforminOxaliplatin55Metformin 500 mg TIDLess incidence of grade 2/3 neuropathy
Significantly low pain score
Low serum level of malondialdehyde and neurotensin
High Ntx-12 score
[74]
MinocyclinePaclitaxel47200 mg on day 1 followed by 100 mg BIDNot effective in CIPN prevention, however, it is effective in p-APS score and fatigue reduction[76]
MinocyclinePaclitaxel66 100 mg BID No significant reduction in fatigue/numbness symptoms with no difference in serum pro-inflammatory markers[120]
MirogabalinOxaliplatin or taxane58 Between 5 and 15 mg twice daily Reduce numeric rating scale (NRS) score [121]
MirogabalinTaxane43 5–30 mg daily Relieve the CIPN symptoms[122]
N-acetyl cysteine Paclitaxel751200 mg daily or 1200 BIDSignificant reduction in peripheral neuropathy in high-dose group
QOL, mTNS were significantly improved.
Significant elevation of NGF serum level
[123]
Omega-3 fatty acidsPaclitaxel604 gNo benefit[124]
Pregabalin (children) 30150–300 mgPregabalin improved pain symptoms significantly [125]
Pregabalin Paclitaxel4675 mg BIDPregabalin not effective in reducing tingling pain, or EORTC QLQ-CIPN20 subscale scores[71]
PregabalinPaclitaxel, docetaxel, or oxaliplatin2675 mg BID for 3 days, 150 mg BID for 3 days, then 300 mg BID until and including day 28Pregabalin was not effective in reduction of average or worst pain compared to placebo [126]
Pregabalin vs. duloxetineTaxane82Pregabalin 150 mg once duloxetine 60 mgBoth were efficacious in decreasing CIPN[127]
Renin-angiotensin-aldosterone system inhibitorsPaclitaxel5886 Decreases the incidence of paclitaxel-induced CIPN in patients with lung cancer[128]
Sigma-1 receptor antagonistOxaliplatin124400 mg MR309 daily dose during the first 5 days of each chemotherapy cycleMR309 reduces sensory and motor hyperexcitability
Patients able to get higher dose of oxaliplatin
[89]
Silybum marianum (SM)Cisplatin60 140 mg three times daily for 90 days SM reduces the scores of DN4 (Douleur neuropathique 4 questions) and CIPNAT (chemotherapy-induced peripheral neuropathy assessment tool)[129]
Venlafaxine Oxaliplatin5037.5 mg XR BIDNot effective in prevention of either acute or chronic neuropathy [59]
VenlafaxineOxaliplatin4850 mg 1 h prior oxaliplatin infusion then 37.5 mg BID from day 21 to 1Effective in relieving acute neurotoxicity as well as improving functional status[58]
Vitamin ETaxane140Vit E 400 mg BIDVit E is not effective in CIPN prevention; however, it helps with shortening of neuropathy[130]

5. Preclinical Studies

Discovering novel treatments for CIPN has attracted a lot of interest since most of the medications tested in clinical trials are unsatisfactory. A meta-analysis conducted in 2015 determined that no drug showed promising effects in terms of preventing or treating CIPN [131]. Therefore, developing novel approaches is in high demand for improving the treatment of CIPN. Numerous compounds and targets have been tested in different animal models, even though these findings may not always be able to translate into patient care in clinics (Table 2).

5.1. Sphingosine-1-Phosphate (S1P) Receptor 1 Antagonist

Sphingosine-1-phosphate is a lipid that alters numerous cellular functions [132]. S1P binds to a group of G protein-coupled receptors (S1PR1-5), and increased activity of S1PR1 could contribute to the development of CIPN [133,134]. Fingolimod is a functional antagonist of S1PR1 and has been approved by the FDA for treating multiple sclerosis [134,135]. It downregulates the expression of the S1P1R receptor to inhibit the NF-kB pathway, which can alleviate [136,137]. Treatment with bortezomib alters S1PRI receptor activity in astrocytes leading to diminishing neuropathic pain [133]. Intrathecal administration of S1PR1 agonist leads to mechanical allodynia in wild-type mice and knocking out S1PR1 receptors in astrocytes does [138]. However, the effects of S1P1R inhibition on preventing and treating CIPN may not work in females [134]. More studies are required to further confirm the therapeutic effect of S1P1R antagonists.

5.2. Cannabinoids

Several animal studies have shown the correlation between endocannabinoid systems and peripheral neuropathy induced by several chemotherapy reagents. Both cannabinoid agonists and cannabinoid enzyme degradation inhibitors (fatty acid amide hydrolase and monoacylglycerol lipase) were efficacious in controlling CIPN. Treatment with cannabigerol, a bioactive compound from Cannabis, diminished the mechanical hypersensitivity by activating α2-adrenergic, CB1, and CB2 receptors in mice receiving cisplatin [139,140]. Administration of cannabidiol and anandamide prevents the development of neurotoxicity induced by paclitaxel via activating the serotonin 1A receptor and reducing the expression of toll-like receptor (TLR)-4 and Iba1 [141,142,143]. Some synthetic cannabinoids, GAT229, WIN55212, and PrNMI, can activate the CB1 receptors to allodynia induced by cisplatin [144,145,146]. Seminally, treatment with synthetic CB2 agonists, such as LY2828360, AM1710, JWH133, and MDA7, attenuates paclitaxel-induced mechanical and cold hypersensitivity by reducing the expression of TNF-α, brain-derived neurotrophic factor (BDNF), and MCP-1 in the lumbar spinal cord [146,147,148,149,150,151,152].
The transporters and enzymes, such as fatty acid amide hydrolase (FAAH), which are involved in cannabinoid receptor function, have been targets for discovering drugs for preventing and treating CIPN [97,98]. A few FAAH inhibitors, including ST4070, URB597, URB937, JZL184, and MJN110, have been tested in animal models of CIPN. Administration of these inhibitors reduces the mechanical hypersensitivity and spontaneous pain behavior evoked by cisplatin [153,154,155,156,157,158,159].
The analgesic effect of crude extract of Cannabis and some minor cannabinoids produced by Cannabis has also been studied in CIPN [160,161,162]. The supercritical fluid carbon dioxide extract of Cannabis sativa L. (Hemp) reduced the mechanical and thermal hypersensitivity by altering the neuroactive ligand–receptor interaction pathway, PPAR signaling pathway, and cAMP signaling pathway [162]. However, some minor cannabinoids, such as cannabinol (CBN), cannabidivarin (CBDV), cannabigerol (CBG), Δ8-tetrahydrocannabinol (Δ8-THC), and Δ9-tetrahydrocannabivarin (THCV), separated from Cannabis had a different analgesic effect. Only CBN was able to relieve the CIPN symptoms in a particular study [161].
Cannabinoids may enhance the analgesic effect of drugs for relieving the CIPN symptoms. Haddad et al. (2023) found that co-treatment with WIN55212 exhibited a great inhibitory effect on capsaicin-induced calcium responses which contributed to the development of nociception [163]. A combination of HU-210 (a CB1R agonist) and SNC80 (DOR agonist) reduced the allodynia in mice treated with paclitaxel [164]. CB2 agonist LY2828360 not only delays the development of neuropathy symptoms but also inhibits morphine tolerance in the CIPN model [152,165]. The combination of cannabidiol and mitragynine significantly reduces thermal hypersensitivity in mice treated with paclitaxel [166]. Some other compounds, including mitragynine, ART26.12, and hyperbaric oxygen, also prevent the CIPN symptoms by activating the cannabinoid receptors [167,168,169].

5.3. Histone Deacetylase 6 (HDAC6) Inhibition

Histone deacetylase 6 (HDAC6) is an enzyme that regulates multiple intracellular functions via deacetylate non-histone proteins [170]. Studies have found that HDAC6 inhibitors, such as ACY-1215, ACY-1083, and ACY738, can prevent and reverse the development of CIPN [40,171,172,173,174]. Treatment with ACY-1215 or ACY-1083 blocks the cisplatin-induced mechanical allodynia by preventing the loss of intraepidermal nerve fibers and mitochondrial damage in peripheral nerves and dorsal root ganglia neurons [40,171]. ACY-1215 and ACY-1083 can also increase the population of M2-macrophages to promote the production of IL-10 leading to a neuroprotective effect in the DRG neurons [172]. In addition, ACY-1083 enhances the tonic enkephalin-DOR (delta opioid receptor) signaling sensory neurons to prevent CIPN [173]. Those findings indicate that HDAC6 inhibitors could be promising drugs to prevent and treat CIPN.

5.4. Interleukin (IL)-10

Interleukin 10 is a classic anti-inflammatory cytokine produced by different types of immune cells [175]. It is a negative feedback regulator that regulates the inflammatory responses to different physiological and pathological conditions [176]. In the last few years, increased IL-10 has been found in different approaches to prevent and treat CIPN. Chemotherapy drugs, such as cisplatin and paclitaxel, can activate CD8+ T cells and CD4+ T cells or trigger receptors expressed on the myeloid cells 2 (TREM2)/DNAX-activating protein of 12 kDa (DAP12), signaling to increase production of IL-10 and leading to the resolution of CIPN [177,178,179,180,181]. Treatment with trimetazidine, inducible co-stimulatory molecule (ICOS) agonist antibody, or adenosine receptor (A3AR) agonist increases the production of IL-10 in the DRG to reduce the mechanical hypersensitivity induced by paclitaxel [44,182,183]. Interestingly, nasal administration of mesenchymal stem cells increases the production of IL-10 to reverse paclitaxel-induced neuropathy [184].

5.5. Transient Receptor Potential Vanilloid 1 and Ankyrin 1 (TRPV1 and TRPA1)

TRPV1 and TRPA1 are non-selective cation channels and are well-known nociceptive receptors for sensing thermal stimulation [185]. The GG-genotype of TRPV1 is associated with severe CIPN in patients with non-small cell lung cancer [186]. Administration of chemotherapy drugs can activate the TRPV1 and TRPA1 channels to induce neuropathy [187,188,189,190]. Studies have found that some potential treatments can reduce the CIPN by inhibiting the activities of TRPV1 and TRPA1. Antioxidants, such as henyl N-tert-butylnitrone and gergenin, reduce the activation of TRP channels to relieve the mechanical hypersensitivity in the animals receiving paclitaxel [187,191]. Intrathecal administration of D-series resolvin 5 (RvD5) reduces the mechanical allodynia induced by paclitaxel only in male mice, but not female ones [192]. Treatment with LPP1 and pregabalin attenuates thermal hypersensitivity, which is partially contributed by the inhibition of TRPV1 and TRPA1 [193]. Apolipoprotein A-I binding protein (AIBP) attenuates the toll-like receptor 4 (TLR4)- and TRPV1- induced pathways to prevent the development of CIPN [194].

5.6. Sterile Alpha and TIR Motif Containing 1 (SARM1) Inhibition

The sterile alpha and TIR motif containing 1 (SARM1) is an enzyme that degenerates axons by increasing intra-axonal calcium flux [195,196]. Pharmacological inhibition or genetic knockout SARM1 has shown protective effects on axons, blocking the development of CIPN. Several studies reported that irreversible SARM1 inhibitors or knockout SARM1 prevented chemotherapy drug (i.e., paclitaxel, vincristine, and bortezomib)-induced intraepidermal nerve fiber loss to maintain axonal function [197,198,199,200]. Knockout SARM1 can also reduce the changes in gene expression in the DRG to block the mechanical and cold hypersensitivity induced by oxaliplatin [201].

5.7. Herbal Medicines

Herbal medicines provide great resources for drug discovery [202]. Numerous herbs have been tested in preventing and treating CIPN. The decoction of Divya-Peedantak-Kwath (DPK) reduced inflammation and oxidative stress to prevent the allodynia and hyperalgesia induced by paclitaxel [203]. Goshajinkigan (GJG), a Japanese herbal medicine, prevented paclitaxel-induced neuropathy by inhibiting the activation of astrocytes in the primary sensory cortex [204,205,206]. Oral administration of ginger water extract increased the expression of 5-HT1A receptors, blocking oxaliplatin-induced cold and mechanical allodynia [207]. Danshen and its bioactive compounds, tanshinone IIA and cryptotanshinone, dose-dependently diminished oxaliplatin-induced CIPN [208]. Commiphora myrrha (CM) resin extract upregulates TRPV1 expression in the spinal cord to prevent the pathogenesis of paclitaxel-induced thermal hyperalgesia and mechanical allodynia [209]. Herbal prescriptions normally include several types of herbs. Some studies reveal that herbal prescriptions, such as SH003 and Siwei Jianbu, show neuroprotective effects, preventing CIPN symptoms [210,211,212].

5.8. Others

Stroke Homing peptide (SHp)-guided deoxyribonuclease 1 (DNase1) and inhibition of myeloperoxidase (MPO) or peptidyl arginine deiminase-4 (PAD4) can reduce chemotherapy-induced mechanical hyperalgesia and prevent the development of CIPN in mice model [213]. Inhibiting the CXCR1/2 signaling pathway using reparixin or ruxolitinib attenuated the development of allodynia induced by oxaliplatin, but not vincristine [214]. Nicotinamide Riboside (NR) is a vitamin B3 precursor of NAD+ which shows efficacy in suppressing tactile and cold hypersensitivity induced by paclitaxel in rats [215]. The NR blunts the loss of the intraepidermal nerve fiber. A high NAD+ level also helps with the deacetylation of alpha-tubulin in DRG via normalizing sirtuin Z [216]. Intraperitoneal injection with magnolin can reduce cold allodynia by inhibiting the activation of extracellular signal-regulated kinase (ERK) in mice with CIPN [217]. Oral treatment with antioxidants, such as N-acetylcysteine, α-lipoic-acid, Vitamin C, and Vitamin E, can dramatically inhibit the neuropathy and neuroinflammation induced by increased production of reactive oxygen species associated with oxaliplatin [218,219].
Numerous other potential targets have been discovered for treating CIPN. Peroxisome proliferator-activated receptor gamma (PPARγ) coactivator 1α (PGC1α) is important for maintaining the function of mitochondria and can reduce oxidative stress, leading to the prevention of CIPN [220]. Treatment with PPARγ activators, such as TZD-A1 and Ursolic acid, decreases paclitaxel-induced mechanical and thermal hypersensitivity [221,222]. Activation of adenosine monophosphate (AMP)-activated protein kinase (AMPK) can promote the expression of macrophage scavenger receptor A1 (SR-A1) or suppressors of cytokine signaling (SOCS) 3, which reduces CIPN in the mice model [223,224]. Vargas-Aliaga A. and colleagues have found that the development of CIPN could be contributed by the genetic polymorphisms of ATP binding cassette subfamily B member 1 (ABCB1, C1236T, and C3435T), but not Glutathione-S Transferase (GST) genes, in patients who received chemotherapy drugs [225,226]. Some other targets, such as the spinal neuronal microRNA-124 (miR-124), miR-3184-5p, Kinin B1 and B2 Receptors, and major histocompatibility complex II (MHCII), have shown activities that regulate the development of CIPN [227,228,229,230,231].
Table 2. Medications in preclinical evaluations to treat CIPN.
Table 2. Medications in preclinical evaluations to treat CIPN.
InterventionAnimal ModelChemotherapyResultsReference
Amifostine MouseoxaliplatinAmifostine prevents mechanical hyperplasia and thermal allodynia development as well as protecting from neural hyperplasia and damage.[232]
Anandamide RatcisplatinAnandamide can prevent mechanical allodynia induced by cisplatin [143]
αO-Conotoxin GeXIVA[1,2]MouseoxaliplatinGeXIVA[1,2] can attenuate the development of oxaliplatin-induced CIPN[233]
C781 Mousepaclitaxel Reverse the mechanical allodynia [234]
Calmangafodibir MouseoxaliplatinCalmangafidipir can protect against mechanical allodynia and thermal hyperplasia induced by oxaliplatin[79]
Cannabinoid RatcisplatinSignificant reduction of mechanical allodynia threshold in rats treated with cannabinoids administered intraperitoneally or locally[146]
Cannabigerol (CBG) and Cannabidiol (CBD) MouseoxaliplatinCBG, CBD, or a combination of both can reverse the mechanical hypersensitivity [235]
CB2 agonistMousepaclitaxelSuppress allodynia induced by paclitaxel[147]
Cannabigerol (CBG)MousecisplatinAttenuates mechanical hypersensitivity[139]
Cannabinoid enzyme degradation inhibitor MousepaclitaxelSuppressed mechanical and cold allodynia[156]
RatvincristineSuppressed mechanical allodynia[159]
MousecisplatinSuppress mechanical allodynia, prevent mechanical hyperplasia[158]
CGRP monoclonal antibody (ZR8 mAb)Mouse & RatcisplatinZR8 mAb reduces mechanical hypersensitivity and thermal nociceptive sensitization induced by cisplatin[236]
Glutathione MouseoxaliplatinGSH relieves neuropathic pain by chelating aluminum[237]
Glycyrrhizic Acid MousepaclitaxelGA exhibits neuroprotective activity and attenuates the development of CIPN[238]
MinocyclineMousevincristineMechanical hypersensitivity reduction with no change in thermal threshold [239]
MinocyclineMousepaclitaxelMinocycline can prevent hypoesthesia and hyperesthesia induced by paclitaxel[77]
NaringeninMousepaclitaxelAttenuate the neuropathic pain induced by paclitaxel [240]
Nicotinamide riboside RatpaclitaxelNR can suppress tactile and cold hypersensitivity without altering tumor growth[215]
Phlomidis radix (P. radix) ethanol extractMousepaclitaxelP. radix and its bioactive compound, sesamoside, diminish the cold and mechanical pain [241]
Phosphosulindac Mousepaclitaxel, vincristine, or oxaliplatinPS dose-dependently relieves the allodynia in mice with CIPN[242]
Polyvalent immunoglobulins Mousevincristine or oxaliplatinAttenuate tactile/cold hypersensitivity and nerve injuries[243]
Prazosin + duloxetine RatoxaliplatinCombination of those drugs reduces paw withdrawal, but no improvement in allodynia or hyperalgesia[244]
PNA6 MouseoxaliplatinPNA6 relieves mechanical hypersensitivity[245]
SS-20 MousepaclitaxelInhibit the development of CIPN[246]
Synthetic peripherally restricted cannabinoids RatcisplatinDose-dependently suppress mechanical and cold allodynia in local and systemic administration[144]
Trimethoxyflavanone (Y3) MousepaclitaxelY3 reduces the activities of DRG neurons to reduce the CIPN.[247]

6. Discussion

In this paper, we have summarized the current information related to the pathophysiology and treatment of CIPN. We also discuss the clinical and preclinical trials that test potential approaches for CIPN management.
The neuropathic pain induced by chemotherapy drugs has attracted a lot of interest in drug discovery. Cancer patients are treated with at least one type of chemotherapy reagent after surgically removing the tumor [248]. Thus, CIPN is predictable in cancer survivors and should be managed with standardized guidelines. Developing an individualized protocol should be considered to prevent and treat CIPN based on the chemotherapy agent(s) to be used.
Several drugs, including Duloxetine, TCA medications, and ion channel modifiers, are used for CIPN treatment, even without FDA approval. Ion channel modifiers have also been shown to be successful in reducing CIPN. Even drugs exhibiting promising therapeutic effects still lack sufficient data to support their use for CIPN treatment. Several potential approaches have been tested in clinical trials. Cannabinoids have risen as a medication to treat emesis, nausea, and vomiting in cancer patients [249]. According to the findings from several clinical studies, the National Academies of Sciences, Engineering, and Medicine have confirmed the promising results of cannabinoids and cannabis in treating CIPN [42]. Non-opioid analgesics, metformin, minocycline, and Calmangafodipir appear to be effective in delaying the onset and reducing the intensity of CIPN symptoms in the acute and chronic phases of neuropathy. However, more clinical trials are required to confirm their therapeutic effects. The potential adverse effects of those reagents are also unclear.
As discussed above, discovering drugs for preventing and treating CIPN has been a hot topic. A few directions could be considered. First, prevention and prophylaxis should be the focus and should be considered in order to reduce the burden on cancer patients who are already suffering. CIPN is an irreversible side effect [2]; thus, approaches for preventing the development of CIPN should be prioritized. Appropriate experimental design and large sample size are required to determine novel approaches for the prevention of CIPN [9]. Second, developing safer chemotherapy drugs with less risk of nerve injury should be focused on. Third, calcium and magnesium channels could be targets for discovering approaches to preventing CIPN [80]. Finally, the combination of pharmacological and nonpharmacological approaches should be applied to patients who have suffered CIPN to increase the efficacy of treatment and improve the quality of life.
In conclusion, the information summarized in this paper provides a better understanding of the approaches and potential medications for treating CIPN. However, more studies are required to develop an efficient plan for the management of CIPN. Further, discovering approaches to preventing the pathogenesis of CIPN should be considered.

Author Contributions

Writing—original draft preparation, M.M., F.U. and M.A.H.; writing—review and editing, M.W.W., R.T., J.S.P., X.C. and W.L.; supervision, W.L. 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

No new data were created in this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zajączkowska, R.; Kocot-Kępska, M.; Leppert, W.; Wrzosek, A.; Mika, J.; Wordliczek, J. Mechanisms of Chemotherapy-Induced Peripheral Neuropathy. Int. J. Mol. Sci. 2019, 20, 1451. [Google Scholar] [CrossRef] [PubMed]
  2. Ibrahim, E.Y.; Ehrlich, B.E. Prevention of chemotherapy-induced peripheral neuropathy: A review of recent findings. Crit. Rev. Oncol. Hematol. 2020, 145, 102831. [Google Scholar] [CrossRef] [PubMed]
  3. Molassiotis, A.; Cheng, H.L.; Leung, K.T.; Li, Y.C.; Wong, K.H.; Au, J.S.K.; Sundar, R.; Chan, A.; Ng, T.R.; Suen, L.K.P.; et al. Risk factors for chemotherapy-induced peripheral neuropathy in patients receiving taxane- and platinum-based chemotherapy. Brain Behav. 2019, 9, e01312. [Google Scholar] [CrossRef] [PubMed]
  4. Gordon-Williams, R.; Farquhar-Smith, P. Recent advances in understanding chemotherapy-induced peripheral neuropathy. F1000Research 2020, 9. [Google Scholar] [CrossRef] [PubMed]
  5. Seretny, M.; Currie, G.L.; Sena, E.S.; Ramnarine, S.; Grant, R.; MacLeod, M.R.; Colvin, L.A.; Fallon, M. Incidence, prevalence, and predictors of chemotherapy-induced peripheral neuropathy: A systematic review and meta-analysis. Pain 2014, 155, 2461–2470. [Google Scholar] [CrossRef] [PubMed]
  6. Burgess, J.; Ferdousi, M.; Gosal, D.; Boon, C.; Matsumoto, K.; Marshall, A.; Mak, T.; Marshall, A.; Frank, B.; Malik, R.A.; et al. Chemotherapy-Induced Peripheral Neuropathy: Epidemiology, Pathomechanisms and Treatment. Oncol. Ther. 2021, 9, 385–450. [Google Scholar] [CrossRef]
  7. Gaskin, D.J.; Richard, P. The economic costs of pain in the United States. J. Pain 2012, 13, 715–724. [Google Scholar] [CrossRef] [PubMed]
  8. Sphar, B.G.; Bowe, C.; Dains, J.E. The Impact of Peripheral Cooling on Chemotherapy-Induced Peripheral Neuropathy: An Integrative Review. J. Adv. Pract. Oncol. 2020, 11, 845–857. [Google Scholar]
  9. Colvin, L.A. Chemotherapy-induced peripheral neuropathy: Where are we now? Pain 2019, 160 (Suppl. S1), S1–S10. [Google Scholar] [CrossRef]
  10. Staff, N.P.; Grisold, A.; Grisold, W.; Windebank, A.J. Chemotherapy-induced peripheral neuropathy: A current review. Ann. Neurol. 2017, 81, 772–781. [Google Scholar] [CrossRef]
  11. Yoon, S.Y.; Oh, J. Neuropathic cancer pain: Prevalence, pathophysiology, and management. Korean J. Intern. Med. 2018, 33, 1058–1069. [Google Scholar] [CrossRef]
  12. Tamburin, S.; Park, S.B.; Schenone, A.; Mantovani, E.; Hamedani, M.; Alberti, P.; Yildiz-Kabak, V.; Kleckner, I.R.; Kolb, N.; Mazzucchelli, M.; et al. Rehabilitation, exercise, and related non-pharmacological interventions for chemotherapy-induced peripheral neurotoxicity: Systematic review and evidence-based recommendations. Crit. Rev. Oncol. Hematol. 2022, 171, 103575. [Google Scholar] [CrossRef] [PubMed]
  13. Loprinzi, C.L.; Lacchetti, C.; Bleeker, J.; Cavaletti, G.; Chauhan, C.; Hertz, D.L.; Kelley, M.R.; Lavino, A.; Lustberg, M.B.; Paice, J.A.; et al. Prevention and Management of Chemotherapy-Induced Peripheral Neuropathy in Survivors of Adult Cancers: ASCO Guideline Update. J. Clin. Oncol. 2020, 38, 3325–3348. [Google Scholar] [CrossRef]
  14. Klafke, N.; Bossert, J.; Kröger, B.; Neuberger, P.; Heyder, U.; Layer, M.; Winkler, M.; Idler, C.; Kaschdailewitsch, E.; Heine, R.; et al. Prevention and Treatment of Chemotherapy-Induced Peripheral Neuropathy (CIPN) with Non-Pharmacological Interventions: Clinical Recommendations from a Systematic Scoping Review and an Expert Consensus Process. Med. Sci. 2023, 11, 15. [Google Scholar] [CrossRef]
  15. Maihöfner, C.; Diel, I.; Tesch, H.; Quandel, T.; Baron, R. Chemotherapy-induced peripheral neuropathy (CIPN): Current therapies and topical treatment option with high-concentration capsaicin. Support. Care Cancer 2021, 29, 4223–4238. [Google Scholar] [CrossRef] [PubMed]
  16. Miaskowski, C.; Paul, S.M.; Mastick, J.; Abrams, G.; Topp, K.; Smoot, B.; Kober, K.M.; Chesney, M.; Mazor, M.; Mausisa, G.; et al. Associations between Perceived Stress and Chemotherapy-Induced Peripheral Neuropathy and Otoxicity in Adult Cancer Survivors. J. Pain Symptom Manag. 2018, 56, 88–97. [Google Scholar] [CrossRef]
  17. Wang, C.; Chen, S.; Jiang, W. Treatment for chemotherapy-induced peripheral neuropathy: A systematic review of randomized control trials. Front. Pharmacol. 2022, 13, 1080888. [Google Scholar] [CrossRef] [PubMed]
  18. Tanay, M.A.L.; Robert, G.; Rafferty, A.M.; Moss-Morris, R.; Armes, J. Clinician and patient experiences when providing and receiving information and support for managing chemotherapy-induced peripheral neuropathy: A qualitative multiple methods study. Eur. J. Cancer Care 2022, 31, e13517. [Google Scholar] [CrossRef]
  19. Malacrida, A.; Meregalli, C.; Rodriguez-Menendez, V.; Nicolini, G. Chemotherapy-Induced Peripheral Neuropathy and Changes in Cytoskeleton. Int. J. Mol. Sci. 2019, 20, 2287. [Google Scholar] [CrossRef]
  20. Kober, K.M.; Schumacher, M.; Conley, Y.P.; Topp, K.; Mazor, M.; Hammer, M.J.; Paul, S.M.; Levine, J.D.; Miaskowski, C. Signaling pathways and gene co-expression modules associated with cytoskeleton and axon morphology in breast cancer survivors with chronic paclitaxel-induced peripheral neuropathy. Mol. Pain 2019, 15, 1744806919878088. [Google Scholar] [CrossRef]
  21. Chen, X.; Gan, Y.; Au, N.P.B.; Ma, C.H.E. Current understanding of the molecular mechanisms of chemotherapy-induced peripheral neuropathy. Front. Mol. Neurosci. 2024, 17, 1345811. [Google Scholar] [CrossRef] [PubMed]
  22. Angelova, P.R.; Abramov, A.Y. Functional role of mitochondrial reactive oxygen species in physiology. Free Radic. Biol. Med. 2016, 100, 81–85. [Google Scholar] [CrossRef] [PubMed]
  23. Klein, I.; Lehmann, H.C. Pathomechanisms of Paclitaxel-Induced Peripheral Neuropathy. Toxics 2021, 9, 229. [Google Scholar] [CrossRef] [PubMed]
  24. Podratz, J.L.; Knight, A.M.; Ta, L.E.; Staff, N.P.; Gass, J.M.; Genelin, K.; Schlattau, A.; Lathroum, L.; Windebank, A.J. Cisplatin induced mitochondrial DNA damage in dorsal root ganglion neurons. Neurobiol. Dis. 2011, 41, 661–668. [Google Scholar] [CrossRef] [PubMed]
  25. Boyette-Davis, J.; Xin, W.; Zhang, H.; Dougherty, P.M. Intraepidermal nerve fiber loss corresponds to the development of taxol-induced hyperalgesia and can be prevented by treatment with minocycline. Pain 2011, 152, 308–313. [Google Scholar] [CrossRef] [PubMed]
  26. Di Cesare Mannelli, L.; Zanardelli, M.; Failli, P.; Ghelardini, C. Oxaliplatin-induced neuropathy: Oxidative stress as pathological mechanism. Protective effect of silibinin. J. Pain 2012, 13, 276–284. [Google Scholar] [CrossRef] [PubMed]
  27. Ma, J.; Kavelaars, A.; Dougherty, P.M.; Heijnen, C.J. Beyond symptomatic relief for chemotherapy-induced peripheral neuropathy: Targeting the source. Cancer 2018, 124, 2289–2298. [Google Scholar] [CrossRef]
  28. Sałat, K. Chemotherapy-induced peripheral neuropathy-part 2: Focus on the prevention of oxaliplatin-induced neurotoxicity. Pharmacol. Rep. 2020, 72, 508–527. [Google Scholar] [CrossRef] [PubMed]
  29. Han, Y.; Smith, M.T. Pathobiology of cancer chemotherapy-induced peripheral neuropathy (CIPN). Front. Pharmacol. 2013, 4, 156. [Google Scholar] [CrossRef]
  30. Flatters, S.J.L.; Dougherty, P.M.; Colvin, L.A. Clinical and preclinical perspectives on Chemotherapy-Induced Peripheral Neuropathy (CIPN): A narrative review. Br. J. Anaesth. 2017, 119, 737–749. [Google Scholar] [CrossRef]
  31. Yamamoto, S.; Egashira, N. Drug Repositioning for the Prevention and Treatment of Chemotherapy-Induced Peripheral Neuropathy: A Mechanism- and Screening-Based Strategy. Front. Pharmacol. 2020, 11, 607780. [Google Scholar] [CrossRef]
  32. Addington, J.; Freimer, M. Chemotherapy-induced peripheral neuropathy: An update on the current understanding. F1000Research 2016, 5. [Google Scholar] [CrossRef]
  33. Aromolaran, K.A.; Goldstein, P.A. Ion channels and neuronal hyperexcitability in chemotherapy-induced peripheral neuropathy; cause and effect? Mol. Pain 2017, 13, 1744806917714693. [Google Scholar] [CrossRef] [PubMed]
  34. Starobova, H.; Vetter, I. Pathophysiology of Chemotherapy-Induced Peripheral Neuropathy. Front. Mol. Neurosci. 2017, 10, 174. [Google Scholar] [CrossRef] [PubMed]
  35. Bennett, D.L.; Clark, A.J.; Huang, J.; Waxman, S.G.; Dib-Hajj, S.D. The Role of Voltage-Gated Sodium Channels in Pain Signaling. Physiol. Rev. 2019, 99, 1079–1151. [Google Scholar] [CrossRef]
  36. Calhoun, E.A.; Welshman, E.E.; Chang, C.H.; Lurain, J.R.; Fishman, D.A.; Hunt, T.L.; Cella, D. Psychometric evaluation of the Functional Assessment of Cancer Therapy/Gynecologic Oncology Group-Neurotoxicity (Fact/GOG-Ntx) questionnaire for patients receiving systemic chemotherapy. Int. J. Gynecol. Cancer 2003, 13, 741–748. [Google Scholar] [CrossRef]
  37. Li, Y.; Zhang, H.; Zhang, H.; Kosturakis, A.K.; Jawad, A.B.; Dougherty, P.M. Toll-like receptor 4 signaling contributes to Paclitaxel-induced peripheral neuropathy. J. Pain 2014, 15, 712–725. [Google Scholar] [CrossRef] [PubMed]
  38. Viatchenko-Karpinski, V.; Ling, J.; Gu, J.G. Down-regulation of Kv4.3 channels and a-type K(+) currents in V2 trigeminal ganglion neurons of rats following oxaliplatin treatment. Mol. Pain 2018, 14, 1744806917750995. [Google Scholar] [CrossRef]
  39. Zhang, H.; Yoon, S.Y.; Zhang, H.; Dougherty, P.M. Evidence that spinal astrocytes but not microglia contribute to the pathogenesis of Paclitaxel-induced painful neuropathy. J. Pain 2012, 13, 293–303. [Google Scholar] [CrossRef]
  40. Krukowski, K.; Ma, J.; Golonzhka, O.; Laumet, G.O.; Gutti, T.; van Duzer, J.H.; Mazitschek, R.; Jarpe, M.B.; Heijnen, C.J.; Kavelaars, A. HDAC6 inhibition effectively reverses chemotherapy-induced peripheral neuropathy. Pain 2017, 158, 1126–1137. [Google Scholar] [CrossRef]
  41. Flatters, S.J.; Bennett, G.J. Ethosuximide reverses paclitaxel- and vincristine-induced painful peripheral neuropathy. Pain 2004, 109, 150–161. [Google Scholar] [CrossRef] [PubMed]
  42. Li, Y.; North, R.Y.; Rhines, L.D.; Tatsui, C.E.; Rao, G.; Edwards, D.D.; Cassidy, R.M.; Harrison, D.S.; Johansson, C.A.; Zhang, H.; et al. DRG Voltage-Gated Sodium Channel 1.7 Is Upregulated in Paclitaxel-Induced Neuropathy in Rats and in Humans with Neuropathic Pain. J. Neurosci. 2018, 38, 1124–1136. [Google Scholar] [CrossRef] [PubMed]
  43. Lees, J.G.; Makker, P.G.; Tonkin, R.S.; Abdulla, M.; Park, S.B.; Goldstein, D.; Moalem-Taylor, G. Immune-mediated processes implicated in chemotherapy-induced peripheral neuropathy. Eur. J. Cancer 2017, 73, 22–29. [Google Scholar] [CrossRef] [PubMed]
  44. Wahlman, C.; Doyle, T.M.; Little, J.W.; Luongo, L.; Janes, K.; Chen, Z.; Esposito, E.; Tosh, D.K.; Cuzzocrea, S.; Jacobson, K.A.; et al. Chemotherapy-induced pain is promoted by enhanced spinal adenosine kinase levels through astrocyte-dependent mechanisms. Pain 2018, 159, 1025–1034. [Google Scholar] [CrossRef]
  45. Hu, L.-Y.; Mi, W.-L.; Wu, G.-C.; Wang, Y.-Q.; Mao-Ying, Q.-L. Prevention and treatment for chemotherapy-induced peripheral neuropathy: Therapies based on CIPN mechanisms. Curr. Neuropharmacol. 2019, 17, 184–196. [Google Scholar] [CrossRef] [PubMed]
  46. Robinson, C.R.; Zhang, H.; Dougherty, P.M. Astrocytes, but not microglia, are activated in oxaliplatin and bortezomib-induced peripheral neuropathy in the rat. Neuroscience 2014, 274, 308–317. [Google Scholar] [CrossRef] [PubMed]
  47. Gornstein, E.L.; Schwarz, T.L. Neurotoxic mechanisms of paclitaxel are local to the distal axon and independent of transport defects. Exp. Neurol. 2017, 288, 153–166. [Google Scholar] [CrossRef] [PubMed]
  48. Bober, B.G.; Shah, S.B. Paclitaxel alters sensory nerve biomechanical properties. J. Biomech. 2015, 48, 3559–3567. [Google Scholar] [CrossRef] [PubMed]
  49. LaPointe, N.E.; Morfini, G.; Brady, S.T.; Feinstein, S.C.; Wilson, L.; Jordan, M.A. Effects of eribulin, vincristine, paclitaxel and ixabepilone on fast axonal transport and kinesin-1 driven microtubule gliding: Implications for chemotherapy-induced peripheral neuropathy. Neurotoxicology 2013, 37, 231–239. [Google Scholar] [CrossRef]
  50. Shemesh, O.A.; Spira, M.E. Paclitaxel induces axonal microtubules polar reconfiguration and impaired organelle transport: Implications for the pathogenesis of paclitaxel-induced polyneuropathy. Acta Neuropathol. 2010, 119, 235–248. [Google Scholar] [CrossRef]
  51. Smith, E.M.; Pang, H.; Cirrincione, C.; Fleishman, S.; Paskett, E.D.; Ahles, T.; Bressler, L.R.; Fadul, C.E.; Knox, C.; Le-Lindqwister, N.; et al. Effect of duloxetine on pain, function, and quality of life among patients with chemotherapy-induced painful peripheral neuropathy: A randomized clinical trial. JAMA 2013, 309, 1359–1367. [Google Scholar] [CrossRef] [PubMed]
  52. Schuessler, B. What do we know about duloxetine’s mode of action? Evidence from animals to humans. BJOG 2006, 113 (Suppl. S1), 5–9. [Google Scholar] [CrossRef] [PubMed]
  53. Meng, J.; Zhang, Q.; Yang, C.; Xiao, L.; Xue, Z.; Zhu, J. Duloxetine, a Balanced Serotonin-Norepinephrine Reuptake Inhibitor, Improves Painful Chemotherapy-Induced Peripheral Neuropathy by Inhibiting Activation of p38 MAPK and NF-κB. Front. Pharmacol. 2019, 10, 365. [Google Scholar] [CrossRef] [PubMed]
  54. Gu, J.; Zhang, H.; Hu, M.; Liu, L.; Chen, C.; Wang, J.; Zhu, F.; Wei, G.; Huo, J. Complementary and alternative medicine in relation to chemotherapy-induced peripheral neuropathy: A narrative review. Explore 2024, 20, 181–187. [Google Scholar] [CrossRef] [PubMed]
  55. Desforges, A.D.; Hebert, C.M.; Spence, A.L.; Reid, B.; Dhaibar, H.A.; Cruz-Topete, D.; Cornett, E.M.; Kaye, A.D.; Urits, I.; Viswanath, O. Treatment and diagnosis of chemotherapy-induced peripheral neuropathy: An update. Biomed. Pharmacother. 2022, 147, 112671. [Google Scholar] [CrossRef] [PubMed]
  56. Rossignol, J.; Cozzi, B.; Liebaert, F.; Hatton, S.; Viallard, M.-L.; Hermine, O.; Greco, C. High concentration of topical amitriptyline for treating chemotherapy-induced neuropathies. Support. Care Cancer 2019, 27, 3053–3059. [Google Scholar] [CrossRef] [PubMed]
  57. Farshchian, N.; Alavi, A.; Heydarheydari, S.; Moradian, N. Comparative study of the effects of venlafaxine and duloxetine on chemotherapy-induced peripheral neuropathy. Cancer Chemother. Pharmacol. 2018, 82, 787–793. [Google Scholar] [CrossRef] [PubMed]
  58. Durand, J.P.; Deplanque, G.; Montheil, V.; Gornet, J.M.; Scotte, F.; Mir, O.; Cessot, A.; Coriat, R.; Raymond, E.; Mitry, E.; et al. Efficacy of venlafaxine for the prevention and relief of oxaliplatin-induced acute neurotoxicity: Results of EFFOX, a randomized, double-blind, placebo-controlled phase III trial. Ann. Oncol. 2012, 23, 200–205. [Google Scholar] [CrossRef] [PubMed]
  59. Zimmerman, C.; Atherton, P.J.; Pachman, D.; Seisler, D.; Wagner-Johnston, N.; Dakhil, S.; Lafky, J.M.; Qin, R.; Grothey, A.; Loprinzi, C.L. MC11C4: A pilot randomized, placebo-controlled, double-blind study of venlafaxine to prevent oxaliplatin-induced neuropathy. Support. Care Cancer 2016, 24, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  60. Kamp, C.B.; Petersen, J.J.; Faltermeier, P.; Juul, S.; Siddiqui, F.; Barbateskovic, M.; Kristensen, A.T.; Moncrieff, J.; Horowitz, M.A.; Hengartner, M.P.; et al. Beneficial and harmful effects of tricyclic antidepressants for adults with major depressive disorder: A systematic review with meta-analysis and trial sequential analysis. BMJ Ment. Health 2024, 27, e300730. [Google Scholar] [CrossRef]
  61. Kautio, A.L.; Haanpää, M.; Kautiainen, H.; Kalso, E.; Saarto, T. Burden of chemotherapy-induced neuropathy—A cross-sectional study. Support. Care Cancer 2011, 19, 1991–1996. [Google Scholar] [CrossRef] [PubMed]
  62. Barton, D.L.; Wos, E.J.; Qin, R.; Mattar, B.I.; Green, N.B.; Lanier, K.S.; Bearden, J.D., 3rd; Kugler, J.W.; Hoff, K.L.; Reddy, P.S.; et al. A double-blind, placebo-controlled trial of a topical treatment for chemotherapy-induced peripheral neuropathy: NCCTG trial N06CA. Support. Care Cancer 2011, 19, 833–841. [Google Scholar] [CrossRef] [PubMed]
  63. Gewandter, J.S.; Mohile, S.G.; Heckler, C.E.; Ryan, J.L.; Kirshner, J.J.; Flynn, P.J.; Hopkins, J.O.; Morrow, G.R. A phase III randomized, placebo-controlled study of topical amitriptyline and ketamine for chemotherapy-induced peripheral neuropathy (CIPN): A University of Rochester CCOP study of 462 cancer survivors. Support. Care Cancer 2014, 22, 1807–1814. [Google Scholar] [CrossRef] [PubMed]
  64. Güler, S.; Könemann, H.; Wolfes, J.; Güner, F.; Ellermann, C.; Rath, B.; Frommeyer, G.; Lange, P.S.; Köbe, J.; Reinke, F.; et al. Lidocaine as an anti-arrhythmic drug: Are there any indications left? Clin. Transl. Sci. 2023, 16, 2429–2437. [Google Scholar] [CrossRef] [PubMed]
  65. van den Heuvel, S.A.S.; van der Wal, S.E.I.; Smedes, L.A.; Radema, S.A.; van Alfen, N.; Vissers, K.C.P.; Steegers, M.A.H. Intravenous Lidocaine: Old-School Drug, New Purpose-Reduction of Intractable Pain in Patients with Chemotherapy Induced Peripheral Neuropathy. Pain Res. Manag. 2017, 2017, 8053474. [Google Scholar] [CrossRef]
  66. Papapetrou, P.; Kumar, A.J.; Muppuri, R.; Chakrabortty, S. Intravenous Lidocaine Infusion to Treat Chemotherapy-Induced Peripheral Neuropathy. A&A Case Rep. 2015, 5, 154–155. [Google Scholar] [CrossRef]
  67. Abouelmagd, G.M.T.; El-Karadawy, S.A.; Abo-Ollo, M.M.; Elwany, Y.N.; Mohamed, E.R.; El-Amrawy, W.Z. Lidocaine Infusion Versus Duloxetine for Prevention and Management of Taxane-Induced Peripheral Neuropathy among Breast Cancer Patients—A Randomized Controlled Study. Pain Physician 2023, 26, E497–E507. [Google Scholar]
  68. Huang, N.L.; Chiang, S.H.; Hsueh, C.H.; Liang, Y.J.; Chen, Y.J.; Lai, L.P. Metformin inhibits TNF-alpha-induced IkappaB kinase phosphorylation, IkappaB-alpha degradation and IL-6 production in endothelial cells through PI3K-dependent AMPK phosphorylation. Int. J. Cardiol. 2009, 134, 169–175. [Google Scholar] [CrossRef] [PubMed]
  69. Rao, R.D.; Michalak, J.C.; Sloan, J.A.; Loprinzi, C.L.; Soori, G.S.; Nikcevich, D.A.; Warner, D.O.; Novotny, P.; Kutteh, L.A.; Wong, G.Y.; et al. Efficacy of gabapentin in the management of chemotherapy-induced peripheral neuropathy. Cancer 2007, 110, 2110–2118. [Google Scholar] [CrossRef]
  70. Anghelescu, D.L.; Tesney, J.M.; Jeha, S.; Wright, B.B.; Trujillo, L.; Sandlund, J.T.; Pauley, J.; Cheng, C.; Pei, D.; Pui, C.H. Prospective randomized trial of interventions for vincristine-related neuropathic pain. Pediatr. Blood Cancer 2020, 67, e28539. [Google Scholar] [CrossRef]
  71. Shinde, S.S.; Seisler, D.; Soori, G.; Atherton, P.J.; Pachman, D.R.; Lafky, J.; Ruddy, K.J.; Loprinzi, C.L. Can pregabalin prevent paclitaxel-associated neuropathy?—An ACCRU pilot trial. Support. Care Cancer 2016, 24, 547–553. [Google Scholar] [CrossRef] [PubMed]
  72. Taylor, A.; Westveld, A.H.; Szkudlinska, M.; Guruguri, P.; Annabi, E.; Patwardhan, A.; Price, T.J.; Yassine, H.N. The use of metformin is associated with decreased lumbar radiculopathy pain. J. Pain Res. 2013, 6, 755–763. [Google Scholar] [CrossRef] [PubMed]
  73. Mao-Ying, Q.L.; Kavelaars, A.; Krukowski, K.; Huo, X.J.; Zhou, W.; Price, T.J.; Cleeland, C.; Heijnen, C.J. The anti-diabetic drug metformin protects against chemotherapy-induced peripheral neuropathy in a mouse model. PLoS ONE 2014, 9, e100701. [Google Scholar] [CrossRef] [PubMed]
  74. El-Fatatry, B.M.; Ibrahim, O.M.; Hussien, F.Z.; Mostafa, T.M. Role of metformin in oxaliplatin-induced peripheral neuropathy in patients with stage III colorectal cancer: Randomized, controlled study. Int. J. Color. Dis. 2018, 33, 1675–1683. [Google Scholar] [CrossRef] [PubMed]
  75. Grothey, A.; Sobrero, A.F.; Shields, A.F.; Yoshino, T.; Paul, J.; Taieb, J.; Souglakos, J.; Shi, Q.; Kerr, R.; Labianca, R.; et al. Duration of Adjuvant Chemotherapy for Stage III Colon Cancer. N. Engl. J. Med. 2018, 378, 1177–1188. [Google Scholar] [CrossRef] [PubMed]
  76. Pachman, D.R.; Dockter, T.; Zekan, P.J.; Fruth, B.; Ruddy, K.J.; Ta, L.E.; Lafky, J.M.; Dentchev, T.; Le-Lindqwister, N.A.; Sikov, W.M.; et al. A pilot study of minocycline for the prevention of paclitaxel-associated neuropathy: ACCRU study RU221408I. Support. Care Cancer 2017, 25, 3407–3416. [Google Scholar] [CrossRef]
  77. Masocha, W. Paclitaxel-induced hyposensitivity to nociceptive chemical stimulation in mice can be prevented by treatment with minocycline. Sci. Rep. 2014, 4, 6719. [Google Scholar] [CrossRef]
  78. Wang, C.H.; Cherng, W.J.; Yang, N.I.; Hsu, C.M.; Yeh, C.H.; Lan, Y.J.; Wang, J.S.; Verma, S. Cyclosporine increases ischemia-induced endothelial progenitor cell mobilization through manipulation of the CD26 system. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008, 294, R811–R818. [Google Scholar] [CrossRef] [PubMed]
  79. Canta, A.; Chiorazzi, A.; Pozzi, E.; Fumagalli, G.; Monza, L.; Meregalli, C.; Carozzi, V.A.; Rodriguez-Menendez, V.; Oggioni, N.; Näsström, J.; et al. Calmangafodipir Reduces Sensory Alterations and Prevents Intraepidermal Nerve Fibers Loss in a Mouse Model of Oxaliplatin Induced Peripheral Neurotoxicity. Antioxidants 2020, 9, 594. [Google Scholar] [CrossRef]
  80. Guillaumot, M.A.; Cerles, O.; Bertrand, H.C.; Benoit, E.; Nicco, C.; Chouzenoux, S.; Schmitt, A.; Batteux, F.; Policar, C.; Coriat, R. Oxaliplatin-induced neuropathy: The preventive effect of a new super-oxide dismutase modulator. Oncotarget 2019, 10, 6418–6431. [Google Scholar] [CrossRef]
  81. Pachman, D.R.; Qin, R.; Seisler, D.K.; Smith, E.M.; Beutler, A.S.; Ta, L.E.; Lafky, J.M.; Wagner-Johnston, N.D.; Ruddy, K.J.; Dakhil, S.; et al. Clinical Course of Oxaliplatin-Induced Neuropathy: Results From the Randomized Phase III Trial N08CB (Alliance). J. Clin. Oncol. 2015, 33, 3416–3422. [Google Scholar] [CrossRef] [PubMed]
  82. Canta, A.; Pozzi, E.; Carozzi, V.A. Mitochondrial Dysfunction in Chemotherapy-Induced Peripheral Neuropathy (CIPN). Toxics 2015, 3, 198–223. [Google Scholar] [CrossRef] [PubMed]
  83. Glimelius, B.; Manojlovic, N.; Pfeiffer, P.; Mosidze, B.; Kurteva, G.; Karlberg, M.; Mahalingam, D.; Buhl Jensen, P.; Kowalski, J.; Bengtson, M.; et al. Persistent prevention of oxaliplatin-induced peripheral neuropathy using calmangafodipir (PledOx(®)): A placebo-controlled randomised phase II study (PLIANT). Acta Oncol. 2018, 57, 393–402. [Google Scholar] [CrossRef]
  84. Pfeiffer, P.; Lustberg, M.; Näsström, J.; Carlsson, S.; Persson, A.; Nagahama, F.; Cavaletti, G.; Glimelius, B.; Muro, K. Calmangafodipir for Prevention of Oxaliplatin-Induced Peripheral Neuropathy: Two Placebo-Controlled, Randomized Phase 3 Studies (POLAR-A/POLAR-M). JNCI Cancer Spectr. 2022, 6, pkac075. [Google Scholar] [CrossRef]
  85. Stankiewicz, A.; Skrzydlewska, E. Amifostine-antioxidant drug in anticancer therapy. Toxicol. Mech. Methods 2006, 16, 181–188. [Google Scholar] [CrossRef]
  86. Verstappen, C.C.; Geldof, A.A.; Postma, T.J.; Heimans, J.J. In vitro protection from cisplatin-induced neurotoxicity by amifostine and its metabolite WR1065. J. Neuro-Oncol. 1999, 44, 1–5. [Google Scholar] [CrossRef]
  87. Moore, D.H.; Donnelly, J.; McGuire, W.P.; Almadrones, L.; Cella, D.F.; Herzog, T.J.; Waggoner, S.E. Limited access trial using amifostine for protection against cisplatin- and three-hour paclitaxel-induced neurotoxicity: A phase II study of the Gynecologic Oncology Group. J. Clin. Oncol. 2003, 21, 4207–4213. [Google Scholar] [CrossRef] [PubMed]
  88. Dorsey, S.G.; Kleckner, I.R.; Barton, D.; Mustian, K.; O’Mara, A.; St Germain, D.; Cavaletti, G.; Danhauer, S.C.; Hershman, D.L.; Hohmann, A.G.; et al. The National Cancer Institute Clinical Trials Planning Meeting for Prevention and Treatment of Chemotherapy-Induced Peripheral Neuropathy. J. Natl. Cancer Inst. 2019, 111, 531–537. [Google Scholar] [CrossRef]
  89. Bruna, J.; Videla, S.; Argyriou, A.A.; Velasco, R.; Villoria, J.; Santos, C.; Nadal, C.; Cavaletti, G.; Alberti, P.; Briani, C.; et al. Efficacy of a Novel Sigma-1 Receptor Antagonist for Oxaliplatin-Induced Neuropathy: A Randomized, Double-Blind, Placebo-Controlled Phase IIa Clinical Trial. Neurotherapeutics 2018, 15, 178–189. [Google Scholar] [CrossRef]
  90. Su, T.P.; Hayashi, T.; Maurice, T.; Buch, S.; Ruoho, A.E. The sigma-1 receptor chaperone as an inter-organelle signaling modulator. Trends Pharmacol. Sci. 2010, 31, 557–566. [Google Scholar] [CrossRef]
  91. Marcotti, A.; Fernández-Trillo, J.; González, A.; Vizcaíno-Escoto, M.; Ros-Arlanzón, P.; Romero, L.; Vela, J.M.; Gomis, A.; Viana, F.; de la Peña, E. TRPA1 modulation by Sigma-1 receptor prevents oxaliplatin-induced painful peripheral neuropathy. Brain 2023, 146, 475–491. [Google Scholar] [CrossRef] [PubMed]
  92. Nieto, F.R.; Cendán, C.M.; Cañizares, F.J.; Cubero, M.A.; Vela, J.M.; Fernández-Segura, E.; Baeyens, J.M. Genetic inactivation and pharmacological blockade of sigma-1 receptors prevent paclitaxel-induced sensory-nerve mitochondrial abnormalities and neuropathic pain in mice. Mol. Pain 2014, 10, 11. [Google Scholar] [CrossRef] [PubMed]
  93. Díaz, A.F.; Polo, S.; Gallardo, N.; Leánez, S.; Pol, O. Analgesic and antidepressant effects of oltipraz on neuropathic pain in mice by modulating microglial activation. J. Clin. Med. 2019, 8, 890. [Google Scholar] [CrossRef] [PubMed]
  94. Woodhams, S.G.; Sagar, D.R.; Burston, J.J.; Chapman, V. The role of the endocannabinoid system in pain. Handb. Exp. Pharmacol. 2015, 227, 119–143. [Google Scholar] [CrossRef] [PubMed]
  95. Zou, S.; Kumar, U. Cannabinoid Receptors and the Endocannabinoid System: Signaling and Function in the Central Nervous System. Int. J. Mol. Sci. 2018, 19, 833. [Google Scholar] [CrossRef] [PubMed]
  96. Tadijan, A.; Vlašić, I.; Vlainić, J.; Đikić, D.; Oršolić, N.; Jazvinšćak Jembrek, M. Intracellular Molecular Targets and Signaling Pathways Involved in Antioxidative and Neuroprotective Effects of Cannabinoids in Neurodegenerative Conditions. Antioxidants 2022, 11, 2049. [Google Scholar] [CrossRef] [PubMed]
  97. Kasatkina, L.A.; Rittchen, S.; Sturm, E.M. Neuroprotective and Immunomodulatory Action of the Endocannabinoid System under Neuroinflammation. Int. J. Mol. Sci. 2021, 22, 5431. [Google Scholar] [CrossRef] [PubMed]
  98. Cuddihey, H.; MacNaughton, W.K.; Sharkey, K.A. Role of the Endocannabinoid System in the Regulation of Intestinal Homeostasis. Cell. Mol. Gastroenterol. Hepatol. 2022, 14, 947–963. [Google Scholar] [CrossRef] [PubMed]
  99. Ellis, R.J.; Toperoff, W.; Vaida, F.; van den Brande, G.; Gonzales, J.; Gouaux, B.; Bentley, H.; Atkinson, J.H. Smoked medicinal cannabis for neuropathic pain in HIV: A randomized, crossover clinical trial. Neuropsychopharmacology 2009, 34, 672–680. [Google Scholar] [CrossRef]
  100. Ware, M.A.; Wang, T.; Shapiro, S.; Robinson, A.; Ducruet, T.; Huynh, T.; Gamsa, A.; Bennett, G.J.; Collet, J.P. Smoked cannabis for chronic neuropathic pain: A randomized controlled trial. CMAJ 2010, 182, E694–E701. [Google Scholar] [CrossRef]
  101. Eisenberg, E.; Ogintz, M.; Almog, S. The pharmacokinetics, efficacy, safety, and ease of use of a novel portable metered-dose cannabis inhaler in patients with chronic neuropathic pain: A phase 1a study. J. Pain Palliat. Care Pharmacother. 2014, 28, 216–225. [Google Scholar] [CrossRef]
  102. Hansen, J.S.; Boix, F.; Hasselstrøm, J.B.; Sørensen, L.K.; Kjolby, M.; Gustavsen, S.; Hansen, R.M.; Petersen, T.; Sellebjerg, F.; Kasch, H.; et al. Pharmacokinetics and pharmacodynamics of cannabis-based medicine in a patient population included in a randomized, placebo-controlled, clinical trial. Clin. Transl. Sci. 2024, 17, e13685. [Google Scholar] [CrossRef]
  103. Weizman, L.; Sharon, H.; Dayan, L.; Espaniol, J.; Brill, S.; Nahman-Averbuch, H.; Hendler, T.; Jacob, G. Oral Delta-9-Tetrahydrocannabinol (THC) Increases Parasympathetic Activity and Supraspinal Conditioned Pain Modulation in Chronic Neuropathic Pain Male Patients: A Crossover, Double-Blind, Placebo-Controlled Trial. CNS Drugs 2024, 38, 375–385. [Google Scholar] [CrossRef] [PubMed]
  104. Langford, R.M.; Mares, J.; Novotna, A.; Vachova, M.; Novakova, I.; Notcutt, W.; Ratcliffe, S. A double-blind, randomized, placebo-controlled, parallel-group study of THC/CBD oromucosal spray in combination with the existing treatment regimen, in the relief of central neuropathic pain in patients with multiple sclerosis. J. Neurol. 2013, 260, 984–997. [Google Scholar] [CrossRef] [PubMed]
  105. Nielsen, S.W.; Hasselsteen, S.D.; Dominiak, H.S.H.; Labudovic, D.; Reiter, L.; Dalton, S.O.; Herrstedt, J. Oral cannabidiol for prevention of acute and transient chemotherapy-induced peripheral neuropathy. Support. Care Cancer 2022, 30, 9441–9451. [Google Scholar] [CrossRef]
  106. Lynch, M.E.; Cesar-Rittenberg, P.; Hohmann, A.G. A double-blind, placebo-controlled, crossover pilot trial with extension using an oral mucosal cannabinoid extract for treatment of chemotherapy-induced neuropathic pain. J. Pain Symptom Manag. 2014, 47, 166–173. [Google Scholar] [CrossRef] [PubMed]
  107. D’Andre, S.; McAllister, S.; Nagi, J.; Giridhar, K.V.; Ruiz-Macias, E.; Loprinzi, C. Topical Cannabinoids for Treating Chemotherapy-Induced Neuropathy: A Case Series. Integr. Cancer Ther. 2021, 20, 15347354211061739. [Google Scholar] [CrossRef]
  108. D’Andre, S.; Novotny, P.; Walters, C.; Lewis-Peters, S.; Thomé, S.; Tofthagen, C.S.; Giridhar, K.V.; Loprinzi, C. Topical Cannabidiol for Established Chemotherapy-Induced Neuropathy: A Pilot Randomized Placebo-Controlled Trial. Cannabis Cannabinoid Res. 2024; ahead of print. [Google Scholar] [CrossRef]
  109. Filipczak-Bryniarska, I.; Krzyzewski, R.M.; Kucharz, J.; Michalowska-Kaczmarczyk, A.; Kleja, J.; Woron, J.; Strzepek, K.; Kazior, L.; Wordliczek, J.; Grodzicki, T.; et al. High-dose 8% capsaicin patch in treatment of chemotherapy-induced peripheral neuropathy: Single-center experience. Med. Oncol. 2017, 34, 162. [Google Scholar] [CrossRef]
  110. Yehia, R.; Saleh, S.; El Abhar, H.; Saad, A.S.; Schaalan, M. L-Carnosine protects against Oxaliplatin-induced peripheral neuropathy in colorectal cancer patients: A perspective on targeting Nrf-2 and NF-κB pathways. Toxicol. Appl. Pharmacol. 2019, 365, 41–50. [Google Scholar] [CrossRef]
  111. Aghili, M.; Taherioun, M.; Jafari, F.; Azadvari, M.; Lashkari, M.; Kolahdouzan, K.; Ghalehtaki, R.; Abdshah, A. Duloxetine to prevent neuropathy in breast cancer patients under paclitaxel chemotherapy (a double-blind randomized trial). Support. Care Cancer 2024, 32, 493. [Google Scholar] [CrossRef]
  112. Sands, S.; Ladas, E.J.; Kelly, K.M.; Weiner, M.; Lin, M.; Ndao, D.H.; Dave, A.; Vahdat, L.T.; Bender, J.G. Glutamine for the treatment of vincristine-induced neuropathy in children and adolescents with cancer. Support. Care Cancer 2017, 25, 701–708. [Google Scholar] [CrossRef]
  113. Leal, A.D.; Qin, R.; Atherton, P.J.; Haluska, P.; Behrens, R.J.; Tiber, C.H.; Watanaboonyakhet, P.; Weiss, M.; Adams, P.T.; Dockter, T.J.; et al. North Central Cancer Treatment Group/Alliance trial N08CA-the use of glutathione for prevention of paclitaxel/carboplatin-induced peripheral neuropathy: A phase 3 randomized, double-blind, placebo-controlled study. Cancer 2014, 120, 1890–1897. [Google Scholar] [CrossRef] [PubMed]
  114. Huang, H.; Zhao, M.; Liu, X.; Song, J.; Liu, L.; Liu, Y.; Xiang, P.; Wang, Y.; Fang, B. Glutathione combined with mecobalamin in the treatment of chemotherapy-induced peripheral neuropathy in multiple myeloma: A retrospective clinical study. Ann. Palliat. Med. 2021, 10, 12335–12346. [Google Scholar] [CrossRef]
  115. Chai, Y.; Zhao, F.; Ye, P.; Ma, F.; Wang, J.; Zhang, P.; Li, Q.; Wang, J.; Wang, W.; Li, Q.; et al. A Prospective, Randomized, Placebo-Controlled Study Assessing the Efficacy of Chinese Herbal Medicine (Huangqi Guizhi Wuwu Decoction) in the Treatment of Albumin-Bound Paclitaxel-Induced Peripheral Neuropathy. J. Clin. Med. 2023, 12, 505. [Google Scholar] [CrossRef] [PubMed]
  116. Cho, K.; Saikawa, H.; Hashimoto, T.; Katagiri, H.; Owada, Y.; Yakuwa, K.; Fujimura, I.; Utsumi, Y.; Akiyama, M.; Nagashima, H.; et al. A randomized trial to evaluate the preventive effect of lafutidine on chemotherapy-induced peripheral neuropathy in patients treated with carboplatin and paclitaxel for lung cancer. Ann. Palliat. Med. 2023, 12, 1136–1145. [Google Scholar] [CrossRef] [PubMed]
  117. Rao, R.D.; Flynn, P.J.; Sloan, J.A.; Wong, G.Y.; Novotny, P.; Johnson, D.B.; Gross, H.M.; Renno, S.I.; Nashawaty, M.; Loprinzi, C.L. Efficacy of lamotrigine in the management of chemotherapy-induced peripheral neuropathy: A phase 3 randomized, double-blind, placebo-controlled trial, N01C3. Cancer 2008, 112, 2802–2808. [Google Scholar] [CrossRef]
  118. Najafi, S.; Heidarali, Z.; Rajabi, M.; Omidi, Z.; Zayeri, F.; Salehi, M.; Haghighat, S. Lithium and preventing chemotherapy-induced peripheral neuropathy in breast cancer patients: A placebo-controlled randomized clinical trial. Trials 2021, 22, 835. [Google Scholar] [CrossRef]
  119. Ehsani, Z.; Salehifar, E.; Habibi, E.; Alizadeh-Navaei, R.; Moosazadeh, M.; Tabrizi, N.; Zaboli, E.; Omrani-Nava, V.; Shekarriz, R. Effect of Melissa officinalis on Chemotherapy-Induced Peripheral Neuropathy in Cancer Patients: A Randomized Trial. Int. J. Hematol. Oncol. Stem Cell Res. 2024, 18, 165–173. [Google Scholar] [CrossRef]
  120. Wang, X.S.; Shi, Q.; Bhadkamkar, N.A.; Cleeland, C.S.; Garcia-Gonzalez, A.; Aguilar, J.R.; Heijnen, C.; Eng, C. Minocycline for Symptom Reduction during Oxaliplatin-Based Chemotherapy for Colorectal Cancer: A Phase II Randomized Clinical Trial. J. Pain Symptom. Manag. 2019, 58, 662–671. [Google Scholar] [CrossRef]
  121. Misawa, S.; Denda, T.; Kodama, S.; Suzuki, T.; Naito, Y.; Kogawa, T.; Takada, M.; Suichi, T.; Shiosakai, K.; Kuwabara, S. Efficacy and safety of mirogabalin for chemotherapy-induced peripheral neuropathy: A prospective single-arm trial (MiroCIP study). BMC Cancer 2023, 23, 1098. [Google Scholar] [CrossRef]
  122. Fujii, T.; Honda, C.; Tanabe, K.; Ogino, M.; Obayashi, S.; Shirabe, K. Efficacy of Mirogabalin for Taxane-associated Chemotherapy-induced Peripheral Neuropathy in Perioperative Chemotherapy for Early Breast Cancer. In Vivo 2024, 38, 1935–1938. [Google Scholar] [CrossRef]
  123. Khalefa, H.G.; Shawki, M.A.; Aboelhassan, R.; El Wakeel, L.M. Evaluation of the effect of N-acetylcysteine on the prevention and amelioration of paclitaxel-induced peripheral neuropathy in breast cancer patients: A randomized controlled study. Breast Cancer Res. Treat. 2020, 183, 117–125. [Google Scholar] [CrossRef]
  124. Tawfik, B.; Dayao, Z.R.; Brown-Glaberman, U.A.; Pankratz, V.S.; Lafky, J.M.; Loprinzi, C.L.; Barton, D.L. A pilot randomized, placebo-controlled, double-blind study of omega-3 fatty acids to prevent paclitaxel-associated acute pain syndrome in breast cancer patients: Alliance A22_Pilot2. Support. Care Cancer 2023, 31, 637. [Google Scholar] [CrossRef]
  125. Vondracek, P.; Oslejskova, H.; Kepak, T.; Mazanek, P.; Sterba, J.; Rysava, M.; Gal, P. Efficacy of pregabalin in neuropathic pain in paediatric oncological patients. Eur. J. Paediatr. Neurol. 2009, 13, 332–336. [Google Scholar] [CrossRef] [PubMed]
  126. Hincker, A.; Frey, K.; Rao, L.; Wagner-Johnston, N.; Ben Abdallah, A.; Tan, B.; Amin, M.; Wildes, T.; Shah, R.; Karlsson, P.; et al. Somatosensory predictors of response to pregabalin in painful chemotherapy-induced peripheral neuropathy: A randomized, placebo-controlled, crossover study. Pain 2019, 160, 1835–1846. [Google Scholar] [CrossRef] [PubMed]
  127. Salehifar, E.; Janbabaei, G.; Hendouei, N.; Alipour, A.; Tabrizi, N.; Avan, R. Comparison of the Efficacy and Safety of Pregabalin and Duloxetine in Taxane-Induced Sensory Neuropathy: A Randomized Controlled Trial. Clin. Drug Investig. 2020, 40, 249–257. [Google Scholar] [CrossRef] [PubMed]
  128. Ihara, Y.; Sawa, K.; Imai, T.; Kimura, T.; Otani, M.; Kawai, R.; Takatori, S.; Shintani, A. Renin-angiotensin-aldosterone system inhibitors are associated with improved paclitaxel-induced peripheral neuropathy in lung cancer: A study using administrative claims data. Support. Care Cancer 2023, 31, 730. [Google Scholar] [CrossRef]
  129. Gholami, A.H.; Ansari, H.; Dadkhah, A. Effect of Silybum Marianum on Reduction of Chemotherapy-Induced Peripheral Neurotoxicity with Cisplatin. Adv. Biomed. Res. 2024, 13, 21. [Google Scholar] [CrossRef] [PubMed]
  130. Heiba, M.A.; Ismail, S.S.; Sabry, M.; Bayoumy, W.A.E.; Kamal, K.A.-A. The use of vitamin E in preventing taxane-induced peripheral neuropathy. Cancer Chemother. Pharmacol. 2021, 88, 931–939. [Google Scholar] [CrossRef]
  131. Poupon, L.; Kerckhove, N.; Vein, J.; Lamoine, S.; Authier, N.; Busserolles, J.; Balayssac, D. Minimizing chemotherapy-induced peripheral neuropathy: Preclinical and clinical development of new perspectives. Expert Opin. Drug Saf. 2015, 14, 1269–1282. [Google Scholar] [CrossRef]
  132. Bryan, A.M.; Del Poeta, M. Sphingosine-1-phosphate receptors and innate immunity. Cell. Microbiol. 2018, 20, e12836. [Google Scholar] [CrossRef]
  133. Stockstill, K.; Doyle, T.M.; Yan, X.; Chen, Z.; Janes, K.; Little, J.W.; Braden, K.; Lauro, F.; Giancotti, L.A.; Harada, C.M.; et al. Dysregulation of sphingolipid metabolism contributes to bortezomib-induced neuropathic pain. J. Exp. Med. 2018, 215, 1301–1313. [Google Scholar] [CrossRef]
  134. Stockstill, K.; Wahlman, C.; Braden, K.; Chen, Z.; Yosten, G.L.; Tosh, D.K.; Jacobson, K.A.; Doyle, T.M.; Samson, W.K.; Salvemini, D. Sexually dimorphic therapeutic response in bortezomib-induced neuropathic pain reveals altered pain physiology in female rodents. Pain 2020, 161, 177–184. [Google Scholar] [CrossRef]
  135. Groves, A.; Kihara, Y.; Chun, J. Fingolimod: Direct CNS effects of sphingosine 1-phosphate (S1P) receptor modulation and implications in multiple sclerosis therapy. J. Neurol. Sci. 2013, 328, 9–18. [Google Scholar] [CrossRef]
  136. Janes, K.; Esposito, E.; Doyle, T.; Cuzzocrea, S.; Tosh, D.K.; Jacobson, K.A.; Salvemini, D. A3 adenosine receptor agonist prevents the development of paclitaxel-induced neuropathic pain by modulating spinal glial-restricted redox-dependent signaling pathways. Pain 2014, 155, 2560–2567. [Google Scholar] [CrossRef]
  137. Huwiler, A.; Zangemeister-Wittke, U. The sphingosine 1-phosphate receptor modulator fingolimod as a therapeutic agent: Recent findings and new perspectives. Pharmacol. Ther. 2018, 185, 34–49. [Google Scholar] [CrossRef]
  138. Jaggi, A.S.; Singh, N. Mechanisms in cancer-chemotherapeutic drugs-induced peripheral neuropathy. Toxicology 2012, 291, 1–9. [Google Scholar] [CrossRef] [PubMed]
  139. Sepulveda, D.E.; Morris, D.P.; Raup-Konsavage, W.M.; Sun, D.; Vrana, K.E.; Graziane, N.M. Cannabigerol (CBG) attenuates mechanical hypersensitivity elicited by chemotherapy-induced peripheral neuropathy. Eur. J. Pain 2022, 26, 1950–1966. [Google Scholar] [CrossRef]
  140. Li, H.; Ward, S.J. Paclitaxel-Associated Mechanical Sensitivity and Neuroinflammation Are Sex-, Time-, and Site-Specific and Prevented through Cannabigerol Administration in C57Bl/6 Mice. Int. J. Mol. Sci. 2024, 25, 4277. [Google Scholar] [CrossRef]
  141. Dos Santos, R.; Veras, F.; Netto, G.; Elisei, L.; Sorgi, C.; Faccioli, L.; Galdino, G. Cannabidiol prevents chemotherapy-induced neuropathic pain by modulating spinal TLR4 via endocannabinoid system activation. J. Pharm. Pharmacol. 2023, 75, 655–665. [Google Scholar] [CrossRef]
  142. Ward, S.J.; McAllister, S.D.; Kawamura, R.; Murase, R.; Neelakantan, H.; Walker, E.A. Cannabidiol inhibits paclitaxel-induced neuropathic pain through 5-HT(1A) receptors without diminishing nervous system function or chemotherapy efficacy. Br. J. Pharmacol. 2014, 171, 636–645. [Google Scholar] [CrossRef]
  143. Çengelli Ünel, Ç.; Ayaz, B.; Aydın, Ş.; Ulupınar, E.; Özatik, O.; Kaygısız, B.; Yıldırım, E.; Erol, K. Protective effects of anandamide against cisplatin-induced peripheral neuropathy in rats. Turk. J. Med. Sci. 2021, 51, 3098–3107. [Google Scholar] [CrossRef]
  144. Mulpuri, Y.; Marty, V.N.; Munier, J.J.; Mackie, K.; Schmidt, B.L.; Seltzman, H.H.; Spigelman, I. Synthetic peripherally-restricted cannabinoid suppresses chemotherapy-induced peripheral neuropathy pain symptoms by CB1 receptor activation. Neuropharmacology 2018, 139, 85–97. [Google Scholar] [CrossRef]
  145. Bagher, A.M.; Binmahfouz, L.S.; Shaik, R.A.; Eid, B.G. Cannabinoid receptor 1 positive allosteric modulator (GAT229) attenuates cisplatin-induced neuropathic pain in mice. Saudi Pharm. J. 2023, 31, 255–264. [Google Scholar] [CrossRef]
  146. Vera, G.; Cabezos, P.A.; Martín, M.I.; Abalo, R. Characterization of cannabinoid-induced relief of neuropathic pain in a rat model of cisplatin-induced neuropathy. Pharmacol. Biochem. Behav. 2013, 105, 205–212. [Google Scholar] [CrossRef]
  147. Lin, X.; Xu, Z.; Carey, L.; Romero, J.; Makriyannis, A.; Hillard, C.J.; Ruggiero, E.; Dockum, M.; Houk, G.; Mackie, K.; et al. A peripheral CB2 cannabinoid receptor mechanism suppresses chemotherapy-induced peripheral neuropathy: Evidence from a CB2 reporter mouse. Pain 2022, 163, 834–851. [Google Scholar] [CrossRef]
  148. Deng, L.; Guindon, J.; Cornett, B.L.; Makriyannis, A.; Mackie, K.; Hohmann, A.G. Chronic cannabinoid receptor 2 activation reverses paclitaxel neuropathy without tolerance or cannabinoid receptor 1-dependent withdrawal. Biol. Psychiatry 2015, 77, 475–487. [Google Scholar] [CrossRef]
  149. Naguib, M.; Xu, J.J.; Diaz, P.; Brown, D.L.; Cogdell, D.; Bie, B.; Hu, J.; Craig, S.; Hittelman, W.N. Prevention of paclitaxel-induced neuropathy through activation of the central cannabinoid type 2 receptor system. Anesth. Analg. 2012, 114, 1104–1120. [Google Scholar] [CrossRef]
  150. Wu, J.; Hocevar, M.; Bie, B.; Foss, J.F.; Naguib, M. Cannabinoid Type 2 Receptor System Modulates Paclitaxel-Induced Microglial Dysregulation and Central Sensitization in Rats. J. Pain 2019, 20, 501–514. [Google Scholar] [CrossRef]
  151. Rahn, E.J.; Deng, L.; Thakur, G.A.; Vemuri, K.; Zvonok, A.M.; Lai, Y.Y.; Makriyannis, A.; Hohmann, A.G. Prophylactic cannabinoid administration blocks the development of paclitaxel-induced neuropathic nociception during analgesic treatment and following cessation of drug delivery. Mol. Pain 2014, 10, 27. [Google Scholar] [CrossRef]
  152. Guenther, K.G.; Lin, X.; Xu, Z.; Makriyannis, A.; Romero, J.; Hillard, C.J.; Mackie, K.; Hohmann, A.G. Cannabinoid CB (2) receptors in primary sensory neurons are implicated in CB (2) agonist-mediated suppression of paclitaxel-induced neuropathic nociception and sexually-dimorphic sparing of morphine tolerance. bioRxiv 2024. [Google Scholar] [CrossRef]
  153. Uhelski, M.L.; Khasabova, I.A.; Simone, D.A. Inhibition of anandamide hydrolysis attenuates nociceptor sensitization in a murine model of chemotherapy-induced peripheral neuropathy. J. Neurophysiol. 2015, 113, 1501–1510. [Google Scholar] [CrossRef]
  154. Curry, Z.A.; Wilkerson, J.L.; Bagdas, D.; Kyte, S.L.; Patel, N.; Donvito, G.; Mustafa, M.A.; Poklis, J.L.; Niphakis, M.J.; Hsu, K.L.; et al. Monoacylglycerol Lipase Inhibitors Reverse Paclitaxel-Induced Nociceptive Behavior and Proinflammatory Markers in a Mouse Model of Chemotherapy-Induced Neuropathy. J. Pharmacol. Exp. Ther. 2018, 366, 169–183. [Google Scholar] [CrossRef] [PubMed]
  155. Thomas, A.; Okine, B.N.; Finn, D.P.; Masocha, W. Peripheral deficiency and antiallodynic effects of 2-arachidonoyl glycerol in a mouse model of paclitaxel-induced neuropathic pain. Biomed. Pharmacother. 2020, 129, 110456. [Google Scholar] [CrossRef]
  156. Slivicki, R.A.; Xu, Z.; Kulkarni, P.M.; Pertwee, R.G.; Mackie, K.; Thakur, G.A.; Hohmann, A.G. Positive Allosteric Modulation of Cannabinoid Receptor Type 1 Suppresses Pathological Pain without Producing Tolerance or Dependence. Biol. Psychiatry 2018, 84, 722–733. [Google Scholar] [CrossRef]
  157. Slivicki, R.A.; Xu, Z.; Mali, S.S.; Hohmann, A.G. Brain permeant and impermeant inhibitors of fatty-acid amide hydrolase suppress the development and maintenance of paclitaxel-induced neuropathic pain without producing tolerance or physical dependence in vivo and synergize with paclitaxel to reduce tumor cell line viability in vitro. Pharmacol. Res. 2019, 142, 267–282. [Google Scholar] [CrossRef]
  158. Khasabova, I.A.; Yao, X.; Paz, J.; Lewandowski, C.T.; Lindberg, A.E.; Coicou, L.; Burlakova, N.; Simone, D.A.; Seybold, V.S. JZL184 is anti-hyperalgesic in a murine model of cisplatin-induced peripheral neuropathy. Pharmacol. Res. 2014, 90, 67–75. [Google Scholar] [CrossRef]
  159. Caprioli, A.; Coccurello, R.; Rapino, C.; Serio, S.D.; Tommaso, M.D.; Vertechy, M.; Vacca, V.; Battista, N.; Pavone, F.; Maccarrone, M.; et al. The Novel Reversible Fatty Acid Amide Hydrolase Inhibitor ST4070 Increases Endocannabinoid Brain Levels and Counteracts Neuropathic Pain in Different Animal Models. J. Pharmacol. Exp. Ther. 2012, 342, 188–195. [Google Scholar] [CrossRef] [PubMed]
  160. Schwarz, A.M.; Keresztes, A.; Bui, T.; Hecksel, R.; Peña, A.; Lent, B.; Gao, Z.G.; Gamez-Rivera, M.; Seekins, C.A.; Chou, K.; et al. Terpenes from Cannabis sativa induce antinociception in a mouse model of chronic neuropathic pain via activation of adenosine A2A receptors. Pain 2024. [Google Scholar] [CrossRef]
  161. Schwarz, A.M.; Kobeci, D.; Mancuso, J.; Moreno-Rodriguez, V.; Seekins, C.; Bui, T.; Welborn, A.; Carr, J.; Streicher, J.M. Select Minor Cannabinoids from Cannabis sativa are Cannabimimetic and Antinociceptive in a Mouse Model of Chronic Neuropathic Pain. J. Pharmacol. Exp. Ther. 2024. [Google Scholar] [CrossRef]
  162. Xu, Y.; Yao, L.; Guo, Y.; Shi, C.; Zhou, J.; Hua, M. The Potential Antinociceptive Effect and Mechanism of Cannabis sativa L. Extract on Paclitaxel-Induced Neuropathic Pain in Rats Uncovered by Multi-Omics Analysis. Molecules 2024, 29, 1958. [Google Scholar] [CrossRef] [PubMed]
  163. Haddad, M.; Alsalem, M.; Saleh, T.; Jaffal, S.M.; Barakat, N.A.; El-Salem, K. Interaction of the synthetic cannabinoid WIN55212 with tramadol on nociceptive thresholds and core body temperature in a chemotherapy-induced peripheral neuropathy pain model. Neuroreport 2023, 34, 441–448. [Google Scholar] [CrossRef] [PubMed]
  164. Sierra, S.; Gupta, A.; Gomes, I.; Fowkes, M.; Ram, A.; Bobeck, E.N.; Devi, L.A. Targeting Cannabinoid 1 and Delta Opioid Receptor Heteromers Alleviates Chemotherapy-Induced Neuropathic Pain. ACS Pharmacol. Transl. Sci. 2019, 2, 219–229. [Google Scholar] [CrossRef] [PubMed]
  165. Iyer, G.; Morris, M.J.; Rathkopf, D.; Slovin, S.F.; Steers, M.; Larson, S.M.; Schwartz, L.H.; Curley, T.; DeLaCruz, A.; Ye, Q.; et al. A phase I trial of docetaxel and pulse-dose 17-allylamino-17-demethoxygeldanamycin in adult patients with solid tumors. Cancer Chemother. Pharmacol. 2012, 69, 1089–1097. [Google Scholar] [CrossRef]
  166. Ortiz, Y.T.; Bilbrey, J.A.; Felix, J.S.; Kienegger, E.A.; Mottinelli, M.; Mukhopadhyay, S.; McCurdy, C.R.; McMahon, L.R.; Wilkerson, J.L. Cannabidiol and mitragynine exhibit differential interactive effects in the attenuation of paclitaxel-induced mechanical allodynia, acute antinociception, and schedule-controlled responding in mice. Pharmacol. Rep. 2023, 75, 937–950. [Google Scholar] [CrossRef]
  167. Warren, G.; Osborn, M.; Tsantoulas, C.; David-Pereira, A.; Cohn, D.; Duffy, P.; Ruston, L.; Johnson, C.; Bradshaw, H.; Kaczocha, M.; et al. Discovery and Preclinical Evaluation of a Novel Inhibitor of FABP5, ART26.12, Effective in Oxaliplatin-Induced Peripheral Neuropathy. J. Pain 2024, 25, 104470. [Google Scholar] [CrossRef]
  168. Meng, X.; Sun, Q.; Miao, T.; Liu, Q.; Ren, H.; Feng, Y. Involvement of spinal cannabinoid receptor type 2 in the analgesia effect of hyperbaric oxygen treatment on paclitaxel-induced neuropathic pain. Undersea Hyperb. Med. 2022, 49, 65–75. [Google Scholar] [CrossRef]
  169. Farkas, D.J.; Inan, S.; Heydari, L.N.; Johnson, C.T.; Zhao, P.; Bradshaw, H.B.; Ward, S.J.; Rawls, S.M. Cannabinoid mechanisms contribute to the therapeutic efficacy of the kratom alkaloid mitragynine against neuropathic, but not inflammatory pain. Life Sci. 2023, 328, 121878. [Google Scholar] [CrossRef]
  170. Yang, W.-B.; Wu, A.-C.; Hsu, T.-I.; Liou, J.-P.; Lo, W.-L.; Chang, K.-Y.; Chen, P.-Y.; Kikkawa, U.; Yang, S.-T.; Kao, T.-J.; et al. Histone deacetylase 6 acts upstream of DNA damage response activation to support the survival of glioblastoma cells. Cell Death Dis. 2021, 12, 884. [Google Scholar] [CrossRef]
  171. Ma, J.; Trinh, R.T.; Mahant, I.D.; Peng, B.; Matthias, P.; Heijnen, C.J.; Kavelaars, A. Cell-specific role of histone deacetylase 6 in chemotherapy-induced mechanical allodynia and loss of intraepidermal nerve fibers. Pain 2019, 160, 2877–2890. [Google Scholar] [CrossRef]
  172. Zhang, J.; Ma, J.; Trinh, R.T.; Heijnen, C.J.; Kavelaars, A. An HDAC6 inhibitor reverses chemotherapy-induced mechanical hypersensitivity via an IL-10 and macrophage dependent pathway. Brain Behav. Immun. 2022, 100, 287–296. [Google Scholar] [CrossRef] [PubMed]
  173. Zhang, J.; Junigan, J.M.; Trinh, R.; Kavelaars, A.; Heijnen, C.J.; Grace, P.M. HDAC6 Inhibition Reverses Cisplatin-Induced Mechanical Hypersensitivity via Tonic Delta Opioid Receptor Signaling. J. Neurosci. 2022, 42, 7862–7874. [Google Scholar] [CrossRef] [PubMed]
  174. Van Helleputte, L.; Kater, M.; Cook, D.P.; Eykens, C.; Rossaert, E.; Haeck, W.; Jaspers, T.; Geens, N.; Vanden Berghe, P.; Gysemans, C.; et al. Inhibition of histone deacetylase 6 (HDAC6) protects against vincristine-induced peripheral neuropathies and inhibits tumor growth. Neurobiol. Dis. 2018, 111, 59–69. [Google Scholar] [CrossRef] [PubMed]
  175. Steen, E.H.; Wang, X.; Balaji, S.; Butte, M.J.; Bollyky, P.L.; Keswani, S.G. The Role of the Anti-Inflammatory Cytokine Interleukin-10 in Tissue Fibrosis. Adv. Wound Care 2020, 9, 184–198. [Google Scholar] [CrossRef]
  176. Saraiva, M.; Vieira, P.; O’Garra, A. Biology and therapeutic potential of interleukin-10. J. Exp. Med. 2020, 217, e20190418. [Google Scholar] [CrossRef] [PubMed]
  177. Krukowski, K.; Eijkelkamp, N.; Laumet, G.; Hack, C.E.; Li, Y.; Dougherty, P.M.; Heijnen, C.J.; Kavelaars, A. CD8+ T Cells and Endogenous IL-10 Are Required for Resolution of Chemotherapy-Induced Neuropathic Pain. J. Neurosci. 2016, 36, 11074–11083. [Google Scholar] [CrossRef]
  178. Singh, S.K.; Krukowski, K.; Laumet, G.O.; Weis, D.; Alexander, J.F.; Heijnen, C.J.; Kavelaars, A. CD8+ T cell-derived IL-13 increases macrophage IL-10 to resolve neuropathic pain. JCI Insight 2022, 7, e154194. [Google Scholar] [CrossRef]
  179. Hu, L.Y.; Zhou, Y.; Cui, W.Q.; Hu, X.M.; Du, L.X.; Mi, W.L.; Chu, Y.X.; Wu, G.C.; Wang, Y.Q.; Mao-Ying, Q.L. Triggering receptor expressed on myeloid cells 2 (TREM2) dependent microglial activation promotes cisplatin-induced peripheral neuropathy in mice. Brain Behav. Immun. 2018, 68, 132–145. [Google Scholar] [CrossRef] [PubMed]
  180. Goode, D.; Whitaker, E.E.; Mecum, N.E. Ovariectomy Reduces Anti-inflammatory CD4+ T Cells in the Dorsal Root Ganglion of NaïVe and Paclitaxel-treated Female Mice. J. Pain 2022, 23 (Suppl. 5), 58. [Google Scholar] [CrossRef]
  181. Whitaker, E.E.; Mecum, N.E.; Cott, R.C.; Goode, D.J. Expression of MHC II in DRG neurons attenuates paclitaxel-induced cold hypersensitivity in male and female mice. PLoS ONE 2024, 19, e0298396. [Google Scholar] [CrossRef]
  182. Sankaranarayanan, I.; Tavares-Ferreira, D.; Mwirigi, J.M.; Mejia, G.L.; Burton, M.D.; Price, T.J. Inducible co-stimulatory molecule (ICOS) alleviates paclitaxel-induced neuropathic pain via an IL-10-mediated mechanism in female mice. J. Neuroinflamm. 2023, 20, 32. [Google Scholar] [CrossRef] [PubMed]
  183. Hammad, A.S.A.; Sayed-Ahmed, M.M.; Abdel Hafez, S.M.N.; Ibrahim, A.R.N.; Khalifa, M.M.A.; El-Daly, M. Trimetazidine alleviates paclitaxel-induced peripheral neuropathy through modulation of TLR4/p38/NF-κB and klotho protein expression. Chem. Biol. Interact. 2023, 376, 110446. [Google Scholar] [CrossRef]
  184. Boukelmoune, N.; Laumet, G.; Tang, Y.; Ma, J.; Mahant, I.; Singh, S.K.; Nijboer, C.; Benders, M.; Kavelaars, A.; Heijnen, C.J. Nasal administration of mesenchymal stem cells reverses chemotherapy-induced peripheral neuropathy in mice. Brain Behav. Immun. 2021, 93, 43–54. [Google Scholar] [CrossRef]
  185. Fernandes, E.S.; Fernandes, M.A.; Keeble, J.E. The functions of TRPA1 and TRPV1: Moving away from sensory nerves. Br. J. Pharmacol. 2012, 166, 510–521. [Google Scholar] [CrossRef] [PubMed]
  186. de Jong, C.; Herder, G.J.M.; van Haarlem, S.W.A.; van der Meer, F.S.; van Lindert, A.S.R.; Ten Heuvel, A.; Brouwer, J.; Egberts, T.C.G.; Deneer, V.H.M. Association between Genetic Variants and Peripheral Neuropathy in Patients with NSCLC Treated with First-Line Platinum-Based Therapy. Genes 2023, 14, 170. [Google Scholar] [CrossRef] [PubMed]
  187. Akhilesh; Chouhan, D.; Ummadisetty, O.; Verma, N.; Tiwari, V. Bergenin ameliorates chemotherapy-induced neuropathic pain in rats by modulating TRPA1/TRPV1/NR2B signalling. Int. Immunopharmacol. 2023, 125 Pt A, 111100. [Google Scholar] [CrossRef]
  188. Chukyo, A.; Chiba, T.; Kambe, T.; Yamamoto, K.; Kawakami, K.; Taguchi, K.; Abe, K. Oxaliplatin-induced changes in expression of transient receptor potential channels in the dorsal root ganglion as a neuropathic mechanism for cold hypersensitivity. Neuropeptides 2018, 67, 95–101. [Google Scholar] [CrossRef]
  189. Leo, M.; Schmitt, L.I.; Küsterarent, P.; Kutritz, A.; Rassaf, T.; Kleinschnitz, C.; Hendgen-Cotta, U.B.; Hagenacker, T. Platinum-Based Drugs Cause Mitochondrial Dysfunction in Cultured Dorsal Root Ganglion Neurons. Int. J. Mol. Sci. 2020, 21, 8639. [Google Scholar] [CrossRef] [PubMed]
  190. Pittman, S.K.; Gracias, N.G.; Vasko, M.R.; Fehrenbacher, J.C. Paclitaxel alters the evoked release of calcitonin gene-related peptide from rat sensory neurons in culture. Exp. Neurol. 2014, 253, 146–153. [Google Scholar] [CrossRef] [PubMed]
  191. Shim, H.S.; Bae, C.; Wang, J.; Lee, K.H.; Hankerd, K.M.; Kim, H.K.; Chung, J.M.; La, J.H. Peripheral and central oxidative stress in chemotherapy-induced neuropathic pain. Mol. Pain 2019, 15, 1744806919840098. [Google Scholar] [CrossRef]
  192. Luo, X.; Gu, Y.; Tao, X.; Serhan, C.N.; Ji, R.R. Resolvin D5 Inhibits Neuropathic and Inflammatory Pain in Male But Not Female Mice: Distinct Actions of D-Series Resolvins in Chemotherapy-Induced Peripheral Neuropathy. Front. Pharmacol. 2019, 10, 745. [Google Scholar] [CrossRef] [PubMed]
  193. Sałat, K.; Cios, A.; Wyska, E.; Sałat, R.; Mogilski, S.; Filipek, B.; Więckowski, K.; Malawska, B. Antiallodynic and antihyperalgesic activity of 3-[4-(3-trifluoromethyl-phenyl)-piperazin-1-yl]-dihydrofuran-2-one compared to pregabalin in chemotherapy-induced neuropathic pain in mice. Pharmacol. Biochem. Behav. 2014, 122, 173–181. [Google Scholar] [CrossRef] [PubMed]
  194. Navia-Pelaez, J.M.; Borges Paes Lemes, J.; Gonzalez, L.; Delay, L.; Dos Santos Aggum Capettini, L.; Lu, J.W.; Gonçalves Dos Santos, G.; Gregus, A.M.; Dougherty, P.M.; Yaksh, T.L.; et al. AIBP regulates TRPV1 activation in chemotherapy-induced peripheral neuropathy by controlling lipid raft dynamics and proximity to TLR4 in dorsal root ganglion neurons. Pain 2023, 164, e274–e285. [Google Scholar] [CrossRef] [PubMed]
  195. Ademi, M.; Yang, X.; Coleman, M.P.; Gilley, J. Natural variants of human SARM1 cause both intrinsic and dominant loss-of-function influencing axon survival. Sci. Rep. 2022, 12, 13846. [Google Scholar] [CrossRef] [PubMed]
  196. Li, Y.; Pazyra-Murphy, M.F.; Avizonis, D.; de Sá Tavares Russo, M.; Tang, S.; Chen, C.Y.; Hsueh, Y.P.; Bergholz, J.S.; Jiang, T.; Zhao, J.J.; et al. Sarm1 activation produces cADPR to increase intra-axonal Ca++ and promote axon degeneration in PIPN. J. Cell Biol. 2022, 221, e202106080. [Google Scholar] [CrossRef]
  197. Geisler, S.; Doan, R.A.; Cheng, G.C.; Cetinkaya-Fisgin, A.; Huang, S.X.; Höke, A.; Milbrandt, J.; DiAntonio, A. Vincristine and bortezomib use distinct upstream mechanisms to activate a common SARM1-dependent axon degeneration program. JCI Insight 2019, 4, e129920. [Google Scholar] [CrossRef] [PubMed]
  198. Geisler, S.; Doan, R.A.; Strickland, A.; Huang, X.; Milbrandt, J.; DiAntonio, A. Prevention of vincristine-induced peripheral neuropathy by genetic deletion of SARM1 in mice. Brain 2016, 139 Pt 12, 3092–3108. [Google Scholar] [CrossRef]
  199. Turkiew, E.; Falconer, D.; Reed, N.; Höke, A. Deletion of Sarm1 gene is neuroprotective in two models of peripheral neuropathy. J. Peripher. Nerv. Syst. 2017, 22, 162–171. [Google Scholar] [CrossRef]
  200. Bosanac, T.; Hughes, R.O.; Engber, T.; Devraj, R.; Brearley, A.; Danker, K.; Young, K.; Kopatz, J.; Hermann, M.; Berthemy, A.; et al. Pharmacological SARM1 inhibition protects axon structure and function in paclitaxel-induced peripheral neuropathy. Brain 2021, 144, 3226–3238. [Google Scholar] [CrossRef]
  201. Gould, S.A.; White, M.; Wilbrey, A.L.; Pór, E.; Coleman, M.P.; Adalbert, R. Protection against oxaliplatin-induced mechanical and thermal hypersensitivity in Sarm1(-/-) mice. Exp. Neurol. 2021, 338, 113607. [Google Scholar] [CrossRef]
  202. Nasim, N.; Sandeep, I.S.; Mohanty, S. Plant-derived natural products for drug discovery: Current approaches and prospects. Nucleus 2022, 65, 399–411. [Google Scholar] [CrossRef] [PubMed]
  203. Balkrishna, A.; Sakat, S.S.; Karumuri, S.; Singh, H.; Tomer, M.; Kumar, A.; Sharma, N.; Nain, P.; Haldar, S.; Varshney, A. Herbal Decoction Divya-Peedantak-Kwath Alleviates Allodynia and Hyperalgesia in Mice Model of Chemotherapy-Induced Peripheral Neuropathy via Modulation in Cytokine Response. Front. Pharmacol. 2020, 11, 566490. [Google Scholar] [CrossRef] [PubMed]
  204. Bahar, M.A.; Andoh, T.; Ogura, K.; Hayakawa, Y.; Saiki, I.; Kuraishi, Y. Herbal Medicine Goshajinkigan Prevents Paclitaxel-Induced Mechanical Allodynia without Impairing Antitumor Activity of Paclitaxel. Evid.-Based Complement. Altern. Med. 2013, 2013, 849754. [Google Scholar] [CrossRef] [PubMed]
  205. Takanashi, K.; Shibata, K.; Mizuno, K.; Komatsu, R.; Koizumi, S. Goshajinkigan attenuates paclitaxel-induced neuropathic pain via cortical astrocytes. Pharmacol. Res. Perspect. 2021, 9, e00850. [Google Scholar] [CrossRef] [PubMed]
  206. Toume, K.; Hou, Z.; Yu, H.; Kato, M.; Maesaka, M.; Bai, Y.; Hanazawa, S.; Ge, Y.; Andoh, T.; Komatsu, K. Search of anti-allodynic compounds from Plantaginis Semen, a crude drug ingredient of Kampo formula “Goshajinkigan”. J. Nat. Med. 2019, 73, 761–768. [Google Scholar] [CrossRef] [PubMed]
  207. Lee, J.H.; Min, D.; Lee, D.; Kim, W. Zingiber officinale Roscoe Rhizomes Attenuate Oxaliplatin-Induced Neuropathic Pain in Mice. Molecules 2021, 26, 548. [Google Scholar] [CrossRef] [PubMed]
  208. Micheli, L.; Maggini, V.; Ciampi, C.; Gallo, E.; Bogani, P.; Fani, R.; Pistelli, L.; Ghelardini, C.; Di Cesare Mannelli, L.; De Leo, M.; et al. Echinacea purpurea against neuropathic pain: Alkamides versus polyphenols efficacy. Phytother. Res. 2023, 37, 1911–1923. [Google Scholar] [CrossRef] [PubMed]
  209. Al-Romaiyan, A.; Barakat, A.; Jose, L.; Masocha, W. An aqueous Commiphora myrrha extract ameliorates paclitaxel-induced peripheral neuropathic pain in mice. Front. Pharmacol. 2023, 14, 1295096. [Google Scholar] [CrossRef] [PubMed]
  210. Lee, K.; Ku, J.M.; Choi, Y.J.; Hwang, H.H.; Jeong, M.; Kim, Y.G.; Kim, M.J.; Ko, S.G. Herbal Prescription SH003 Alleviates Docetaxel-Induced Neuropathic Pain in C57BL/6 Mice. Evid.-Based Complement. Altern. Med. 2021, 2021, 4120334. [Google Scholar] [CrossRef]
  211. Zhang, P.; Lu, Y.; Yang, C.; Zhang, Q.; Qian, Y.; Suo, J.; Cheng, P.; Zhu, J. Based on Systematic Pharmacology: Molecular Mechanism of Siwei Jianbu Decoction in Preventing Oxaliplatin-Induced Peripheral Neuropathy. Neural Plast. 2020, 2020, 8880543. [Google Scholar] [CrossRef]
  212. Lee, J.H.; Kim, B.; Ko, S.G.; Kim, W. Analgesic Effect of SH003 and Trichosanthes kirilowii Maximowicz in Paclitaxel-Induced Neuropathic Pain in Mice. Curr. Issues Mol. Biol. 2022, 44, 718–730. [Google Scholar] [CrossRef] [PubMed]
  213. Wang, C.Y.; Lin, T.T.; Hu, L.; Xu, C.J.; Hu, F.; Wan, L.; Yang, X.; Wu, X.F.; Zhang, X.T.; Li, Y.; et al. Neutrophil extracellular traps as a unique target in the treatment of chemotherapy-induced peripheral neuropathy. EBioMedicine 2023, 90, 104499. [Google Scholar] [CrossRef] [PubMed]
  214. Cho, H.S.; Choi, Y.I.; Park, S.U.; Han, Y.S.; Kwon, J.; Jung, S.J. Prevention of Chemotherapy-Induced Peripheral Neuropathy by Inhibiting C-X-C Motif Chemokine Receptor 2. Int. J. Mol. Sci. 2023, 24, 1855. [Google Scholar] [CrossRef] [PubMed]
  215. Hamity, M.V.; White, S.R.; Blum, C.; Gibson-Corley, K.N.; Hammond, D.L. Nicotinamide riboside relieves paclitaxel-induced peripheral neuropathy and enhances suppression of tumor growth in tumor-bearing rats. Pain 2020, 161, 2364–2375. [Google Scholar] [CrossRef] [PubMed]
  216. Hamity, M.V.; White, S.R.; Walder, R.Y.; Schmidt, M.S.; Brenner, C.; Hammond, D.L. Nicotinamide riboside, a form of vitamin B3 and NAD+ precursor, relieves the nociceptive and aversive dimensions of paclitaxel-induced peripheral neuropathy in female rats. Pain 2017, 158, 962–972. [Google Scholar] [CrossRef]
  217. Kim, N.; Chung, G.; Son, S.R.; Park, J.H.; Lee, Y.H.; Park, K.T.; Cho, I.H.; Jang, D.S.; Kim, S.K. Magnolin Inhibits Paclitaxel-Induced Cold Allodynia and ERK1/2 Activation in Mice. Plants 2023, 12, 2283. [Google Scholar] [CrossRef] [PubMed]
  218. Agnes, J.P.; Santos, V.W.D.; das Neves, R.N.; Gonçalves, R.M.; Delgobo, M.; Girardi, C.S.; Lückemeyer, D.D.; Ferreira, M.A.; Macedo-Júnior, S.J.; Lopes, S.C.; et al. Antioxidants Improve Oxaliplatin-Induced Peripheral Neuropathy in Tumor-Bearing Mice Model: Role of Spinal Cord Oxidative Stress and Inflammation. J. Pain 2021, 22, 996–1013. [Google Scholar] [CrossRef]
  219. Guindon, J.; Deng, L.; Fan, B.; Wager-Miller, J.; Hohmann, A.G. Optimization of a cisplatin model of chemotherapy-induced peripheral neuropathy in mice: Use of vitamin C and sodium bicarbonate pretreatments to reduce nephrotoxicity and improve animal health status. Mol. Pain 2014, 10, 56. [Google Scholar] [CrossRef] [PubMed]
  220. Zhai, M.; Hu, H.; Zheng, Y.; Wu, B.; Sun, W. PGC1α: An emerging therapeutic target for chemotherapy-induced peripheral neuropathy. Ther. Adv. Neurol. Disord. 2023, 16, 17562864231163361. [Google Scholar] [CrossRef]
  221. Benvenutti, L.; Wolff, F.R.; Corrêa, T.P.; Melato, J.; Goldoni, F.C.; De Faveri, R.; Patel, Y.B.K.; de Souza, J.A.; Grockoski, H.A.; Nilz, P.M.; et al. A partial agonist of PPARγ prevents paclitaxel-induced peripheral neuropathy in mice, by inhibiting neuroinflammation. Br. J. Pharmacol. 2024, 181, 1128–1149. [Google Scholar] [CrossRef]
  222. Yang, Y.; He, Z.; Wu, S. Ursolic acid alleviates paclitaxel-induced peripheral neuropathy through PPARγ activation. Toxicol. Appl. Pharmacol. 2024, 484, 116883. [Google Scholar] [CrossRef] [PubMed]
  223. Yang, X.; Jia, R.; Hu, F.; Fan, W.; Lin, T.; Zhang, X.; Xu, C.; Ruan, S.; Jiang, C.; Li, Y.; et al. Promoting AMPK/SR-A1-mediated clearance of HMGB1 attenuates chemotherapy-induced peripheral neuropathy. Cell Commun. Signal. 2023, 21, 99. [Google Scholar] [CrossRef] [PubMed]
  224. Zhang, X.T.; Zong, L.J.; Jia, R.M.; Qin, X.M.; Ruan, S.R.; Lu, L.L.; Wang, P.; Hu, L.; Liu, W.T.; Yang, Y.; et al. Ozone attenuates chemotherapy-induced peripheral neuropathy via upregulating the AMPK-SOCS3 axis. J. Cancer Res. Ther. 2023, 19, 1031–1039. [Google Scholar] [CrossRef]
  225. Vargas-Aliaga, A.; De la Haba, M.; Contreras, M.J.; Morales Estevez, C.; Porras, I.; Cano, M.T.; Pulido, G.; Gómez, M.A.; Flores-Paco, P.; Juan, H.R.; et al. NeuroPredict: Study of the predictive value of ABCB1 genetic polymorphisms and associated clinical factors in chronic chemotherapy-induced peripheral neuropathy (CIPN). Front. Pharmacol. 2024, 15, 1352939. [Google Scholar] [CrossRef] [PubMed]
  226. Dunn, P.J.; Griffiths, L.R.; Yates, P.; Haupt, L.M.; Alexander, K.E. GSTM1 and GSTT1 polymorphisms associated with pain in a chemotherapy-induced peripheral neuropathy cohort. J. Cancer Res. Clin. Oncol. 2023, 149, 7405–7412. [Google Scholar] [CrossRef] [PubMed]
  227. Li, X.C.; Chen, H.; Chen, Y.; Chu, Y.X.; Mi, W.L.; Wang, Y.Q.; Mao-Ying, Q.L. Spinal Neuronal miR-124 Inhibits Microglial Activation and Contributes to Preventive Effect of Electroacupuncture on Chemotherapy-Induced Peripheral Neuropathy in Mice. J. Immunol. 2024, 212, 410–420. [Google Scholar] [CrossRef]
  228. Becker, G.; Fialho, M.F.P.; Brusco, I.; Oliveira, S.M. Kinin B(1) and B(2) Receptors Contribute to Cisplatin-Induced Painful Peripheral Neuropathy in Male Mice. Pharmaceutics 2023, 15, 852. [Google Scholar] [CrossRef]
  229. Becker, G.; Fialho, M.F.P.; Brum, E.S.; Oliveira, S.M. Kinin B(2) Receptor Mediates Cisplatin-Induced Painful Peripheral Neuropathy by Intracellular Kinase Pathways and TRPA1 Channel Sensitisation. Pharmaceuticals 2023, 16, 959. [Google Scholar] [CrossRef]
  230. Whitaker, E.E.; Mecum, N.E.; Cott, R.C.; Goode, D.J. Novel expression of major histocompatibility complex II in dorsal root ganglion neurons attenuates paclitaxel-induced cold hypersensitivity in male and female mice. bioRxiv 2023. [Google Scholar] [CrossRef]
  231. Ju, Y.; Baek, D.H.; Choi, G.E.; Jang, A. Exploring microRNA patterns as biomarkers of FOLFOX chemotherapy-induced peripheral neuropathy in patients with colorectal cancer. Biochim. Biophys. Acta Mol. Basis Dis. 2024, 1870, 167209. [Google Scholar] [CrossRef]
  232. Pereira, A.F.; Lino, J.A.; Alves, B.W.F.; Lisboa, M.R.P.; Pontes, R.B.; Leite, C.; Nogueira, R.B.; Lima-Júnior, R.C.P.; Vale, M.L. Amifostine protects from the peripheral sensory neuropathy induced by oxaliplatin in mice. Braz. J. Med. Biol. Res. 2020, 53, e10263. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, H.; Li, X.; Qiao, Y.; Wang, M.; Wang, W.; McIntosh, J.M.; Zhangsun, D.; Luo, S. αO-Conotoxin GeXIVA[1,2] Reduced Neuropathic Pain and Changed Gene Expression in Chronic Oxaliplatin-Induced Neuropathy Mice Model. Mar. Drugs 2024, 22, 49. [Google Scholar] [CrossRef]
  234. Kume, M.; Ahmad, A.; DeFea, K.A.; Vagner, J.; Dussor, G.; Boitano, S.; Price, T.J. Protease-Activated Receptor 2 (PAR2) Expressed in Sensory Neurons Contributes to Signs of Pain and Neuropathy in Paclitaxel Treated Mice. J. Pain 2023, 24, 1980–1993. [Google Scholar] [CrossRef] [PubMed]
  235. Hayduk, S.A.; Hughes, A.C.; Winter, R.L.; Milton, M.D.; Ward, S.J. Single and Combined Effects of Cannabigerol (CBG) and Cannabidiol (CBD) in Mouse Models of Oxaliplatin-Associated Mechanical Sensitivity, Opioid Antinociception, and Naloxone-Precipitated Opioid Withdrawal. Biomedicines 2024, 12, 1145. [Google Scholar] [CrossRef]
  236. Xie, S.; Gao, Z.; Zhang, J.; Xing, C.; Dong, Y.; Wang, L.; Wang, Z.; Li, Y.; Li, G.; Han, G.; et al. Monoclonal Antibody Targeting CGRP Relieves Cisplatin-Induced Neuropathic Pain by Attenuating Neuroinflammation. Neurotox. Res. 2024, 42, 8. [Google Scholar] [CrossRef]
  237. Lee, M.; Cho, S.; Roh, K.; Chae, J.; Park, J.H.; Park, J.; Lee, M.A.; Kim, J.; Auh, C.K.; Yeom, C.H.; et al. Glutathione alleviated peripheral neuropathy in oxaliplatin-treated mice by removing aluminum from dorsal root ganglia. Am. J. Transl. Res. 2017, 9, 926–939. [Google Scholar] [PubMed]
  238. Klein, I.; Isensee, J.; Wiesen, M.H.J.; Imhof, T.; Wassermann, M.K.; Müller, C.; Hucho, T.; Koch, M.; Lehmann, H.C. Glycyrrhizic Acid Prevents Paclitaxel-Induced Neuropathy via Inhibition of OATP-Mediated Neuronal Uptake. Cells 2023, 12, 1249. [Google Scholar] [CrossRef]
  239. Starobova, H.; Mueller, A.; Allavena, R.; Lohman, R.J.; Sweet, M.J.; Vetter, I. Minocycline Prevents the Development of Mechanical Allodynia in Mouse Models of Vincristine-Induced Peripheral Neuropathy. Front. Neurosci. 2019, 13, 653. [Google Scholar] [CrossRef]
  240. Pan, C.; Xu, Y.; Jiang, Z.; Fan, C.; Chi, Z.; Zhang, Y.; Miao, M.; Ren, Y.; Wu, Z.; Xu, L.; et al. Naringenin relieves paclitaxel-induced pain by suppressing calcitonin gene-related peptide signalling and enhances the anti-tumour action of paclitaxel. Br. J. Pharmacol. 2024, 181, 3136–3159. [Google Scholar] [CrossRef]
  241. Park, K.T.; Ko, S.G.; Kim, W. Phlomidis Radix Extract Alleviates Paclitaxel-Induced Neuropathic Pain by Modulating Spinal TRPV1 in Mice. Plants 2023, 12, 3819. [Google Scholar] [CrossRef]
  242. Basu, A.; Yang, J.Y.; Tsirukis, V.E.; Loiacono, A.; Koch, G.; Khwaja, I.A.; Krishnamurthy, M.; Fazio, N.; White, E.; Jha, A.; et al. Phosphosulindac (OXT-328) prevents and reverses chemotherapy induced peripheral neuropathy in mice. Front. Neurosci. 2023, 17, 1240372. [Google Scholar] [CrossRef] [PubMed]
  243. Mroué, M.; Bessaguet, F.; Nizou, A.; Richard, L.; Sturtz, F.; Magy, L.; Bourthoumieu, S.; Danigo, A.; Demiot, C. Neuroprotective Effect of Polyvalent Immunoglobulins on Mouse Models of Chemotherapy-Induced Peripheral Neuropathy. Pharmaceutics 2024, 16, 139. [Google Scholar] [CrossRef] [PubMed]
  244. Wagner, M.A.; Smith, E.M.L.; Ayyash, N.; Holden, J.E. Prazosin as an Adjuvant to Increase Effectiveness of Duloxetine in a Rat Model of Oxaliplatin-Induced Peripheral Neuropathy. Semin. Oncol. Nurs. 2024; in press. [Google Scholar] [CrossRef]
  245. Sulaiman, M.I.; Alabsi, W.; Szabo, L.; Hay, M.; Polt, R.; Largent-Milnes, T.M.; Vanderah, T.W. PNA6, a Lactosyl Analogue of Angiotensin-(1-7), Reverses Pain Induced in Murine Models of Inflammation, Chemotherapy-Induced Peripheral Neuropathy, and Metastatic Bone Disease. Int. J. Mol. Sci. 2023, 24, 5007. [Google Scholar] [CrossRef] [PubMed]
  246. Itoh, K.; Shimoyama, M.; Schiller, P.W.; Toyama, S. Protective effect of a mitochondria-targeting peptide against paclitaxel-induced peripheral neuropathy. Chem. Biol. Drug Des. 2023, 101, 1012–1018. [Google Scholar] [CrossRef] [PubMed]
  247. Mei, C.; Pan, C.; Xu, L.; Miao, M.; Lu, Q.; Yu, Y.; Lin, P.; Wu, W.; Ni, F.; Gao, Y.; et al. Trimethoxyflavanone relieves Paclitaxel-induced neuropathic pain via inhibiting expression and activation of P2X7 and production of CGRP in mice. Neuropharmacology 2023, 236, 109584. [Google Scholar] [CrossRef]
  248. Anand, U.; Dey, A.; Chandel, A.K.S.; Sanyal, R.; Mishra, A.; Pandey, D.K.; De Falco, V.; Upadhyay, A.; Kandimalla, R.; Chaudhary, A.; et al. Cancer chemotherapy and beyond: Current status, drug candidates, associated risks and progress in targeted therapeutics. Genes Dis. 2023, 10, 1367–1401. [Google Scholar] [CrossRef]
  249. Smith, L.A.; Azariah, F.; Lavender, V.T.; Stoner, N.S.; Bettiol, S. Cannabinoids for nausea and vomiting in adults with cancer receiving chemotherapy. Cochrane Database Syst. Rev. 2015, 2015, Cd009464. [Google Scholar] [CrossRef]
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

Mattar, M.; Umutoni, F.; Hassan, M.A.; Wamburu, M.W.; Turner, R.; Patton, J.S.; Chen, X.; Lei, W. Chemotherapy-Induced Peripheral Neuropathy: A Recent Update on Pathophysiology and Treatment. Life 2024, 14, 991. https://doi.org/10.3390/life14080991

AMA Style

Mattar M, Umutoni F, Hassan MA, Wamburu MW, Turner R, Patton JS, Chen X, Lei W. Chemotherapy-Induced Peripheral Neuropathy: A Recent Update on Pathophysiology and Treatment. Life. 2024; 14(8):991. https://doi.org/10.3390/life14080991

Chicago/Turabian Style

Mattar, Marina, Florence Umutoni, Marwa A. Hassan, M. Wambui Wamburu, Reagan Turner, James S. Patton, Xin Chen, and Wei Lei. 2024. "Chemotherapy-Induced Peripheral Neuropathy: A Recent Update on Pathophysiology and Treatment" Life 14, no. 8: 991. https://doi.org/10.3390/life14080991

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