CHAPTER SIXTEEN

Chemotherapy-Induced Peripheral Neuropathy Jill C. Fehrenbacher1 Department of Pharmacology and Toxicology, Indiana University School of Medicine, Indianapolis, Indiana, USA Stark Neuroscience Research Institute, Indiana University School of Medicine, Indianapolis, Indiana, USA Department of Anesthesiology, Indiana University School of Medicine, Indianapolis, Indiana, USA 1 Corresponding author: e-mail address: [email protected]

Contents 1. Chemotherapy-Induced Neuropathy: An Introduction 2. Classes of Chemotherapeutics 2.1 Microtubule-targeting agents 2.2 Platinum-containing agents 2.3 Proteasome inhibitors 2.4 Angiogenesis inhibitors 3. Clinical Assessment of CIPN 4. Experimental Studies: Animal Models of CIPN 5. Experimental Studies: In Vitro Models of CIPN 6. Proposed Mechanisms Underlying CIPN 6.1 Mitochondrial dysfunction 6.2 Nitroxidative stress 6.3 DNA damage 6.4 Ion channel modulation 6.5 Inflammation 6.6 Neurotrophic factors 6.7 Microtubule alterations 7. Challenges to CIPN Research References

472 473 473 474 475 475 476 478 481 483 484 486 487 487 489 490 492 492 493

Abstract Chemotherapy-induced peripheral neuropathy (CIPN) is common in patients receiving anticancer treatment and can affect survivability and long-term quality of life of the patient following treatment. The symptoms of CIPN primarily include abnormal sensory discrimination of touch, vibration, thermal information, and pain. There is currently a paucity of pharmacological agents to prevent or treat CIPN. The lack of efficacious therapeutics is due, at least in part, to an incomplete understanding of the mechanisms by which chemotherapies alter the sensitivity of sensory neurons. Although the clinical presentation of CIPN can be similar with the various classes of chemotherapeutic agents,

Progress in Molecular Biology and Translational Science, Volume 131 ISSN 1877-1173 http://dx.doi.org/10.1016/bs.pmbts.2014.12.002

#

2015 Elsevier Inc. All rights reserved.

471

472

Jill C. Fehrenbacher

there are subtle differences, suggesting that each class of drugs might induce neuropathy via different mechanisms. Multiple mechanisms have been proposed to underlie the development and maintenance of neuropathy; however, most pharmacological agents generated from preclinical experiments have failed to alleviate the symptoms of CIPN in the clinic. Further research is necessary to identify the specific mechanisms by which each class of chemotherapeutics induces neuropathy.

1. CHEMOTHERAPY-INDUCED NEUROPATHY: AN INTRODUCTION With considerable advances in the development of anticancer agents over the past 20–30 years, patients diagnosed with cancer are surviving and living longer following treatment.1 With this increase in survival, researchers and patients are now shifting their focus toward a primary side effect of anticancer treatment, chemotherapy-induced peripheral neuropathy (CIPN). Peripheral neuropathy is an adverse effect following treatment with multiple classes of chemotherapeutics, including platinum drugs, microtubuletargeting agents (MTAs), proteasome inhibitors, and angiogenesis inhibitors. The reported incidence of CIPN varies widely among patients and chemotherapeutic drugs, but often ranges between 30% and 40%.2 CIPN can be painful and persist as a significant disability following anticancer treatment, resulting in a decreased quality of life for survivors. The symptoms of peripheral neuropathy are primarily sensory and can be divided into gain in function manifestations including burning pain, tingling, and hypersensitivity to cold or touch and loss-of-function characteristics including loss of proprioception, decreased perception of vibration and pinprick, and numbness.3–8 Numbness, the loss of vibratory sense, and deep tendon ankle reflexes are some of the first signs of neuropathy in patients receiving chemotherapy treatment, followed by the development of paresthesia and loss of positional sense.7,9,10 The neuropathy induced by chemotherapeutics is thought to be induced by varying degrees of axonopathy.11–13 As such, symptoms develop in a stocking and glove distribution,14 suggesting that sensory neurons with long fibers such as the axons innervating the hands and feet, are especially susceptible to neurotoxic insult. At high doses, the effects of chemotherapeutics on the autonomic and motor nervous systems can be observed,7,8,15,16 but these are not observed frequently and are not thought to mediate the symptoms of CIPN most commonly reported by patients.

Chemotherapy-Induced Peripheral Neuropathy

473

There are a number of factors which can contribute to the development of neuropathy in patients receiving chemotherapy, including dose intensity,17 cumulative dose,9,18 concurrent treatment with other neurotoxic agents,19 and preexisting conditions, such as diabetic- and/or alcohol-induced peripheral neuropathy.14,20–22 The onset and reversibility of CIPN symptoms are variable between different chemotherapeutic classes and are also dependent on the cumulative dose of drug administered. In some cases, neuropathy symptoms may worsen following completion of treatment, and this phenomenon is called coasting.23,24 Resolution of peripheral neuropathy following the discontinuation of chemotherapeutic treatment is quite variable among patients and between the drugs used for treatment, but can also be dependent on the cumulative dose of the chemotherapeutics administered25,26 and on the intensity of the neuropathic symptoms at their peak.26,27 Some patients experience persistent disabling neuropathy for 9–13 years following the cessation of therapy.28,29 There are also some patients who report mild neuropathy during treatment and an absence of neuropathy symptoms following the cessation of drug treatment. Even in these asymptomatic patients, however, objective analyses of neuropathy (standardized examinations or nerve conduction studies) indicate the persistence of neuropathy,30 suggesting long-term effects of chemotherapy. Unfortunately, no established biomarkers have been identified to predict the development, intensity, or reversibility of CIPN.

2. CLASSES OF CHEMOTHERAPEUTICS 2.1 Microtubule-targeting agents Alterations in microtubule dynamics are a common mechanism of blocking mitosis within dividing cells, which lead to subsequent apoptosis in cancer cells,31 and MTAs have been developed and used extensively as anticancer drugs.32 There are two main classes of MTAs, which are compounds that bind directly to the tubulin subunits to alter the dynamic rearrangement of microtubules. Compounds that bind free β-tubulin subunits and prevent polymerization into microtubules are called depolymerizing or destabilizing agents and include vincristine, vinblastine, and vinorelbine.33–35 The other class of MTAs binds polymerized β-tubulin comprising the microtubule polymer and prevents the depolymerization of the microtubule. These agents are called polymerizing or stabilizing agents and include paclitaxel, docetaxel, and the epothilones.36,37 Although they have opposite

474

Jill C. Fehrenbacher

effects on microtubules, both MTA classes interfere with the dynamic rearrangement of microtubules. The taxanes are used as treatments for several types of cancer including breast, ovarian, gastric, and lung cancers.22 The clinical onset of symptoms of neuropathy occurs 3–6 weeks following the first dose of paclitaxel.7 Prior to the development of CIPN, many patients also experience paclitaxelassociated acute pain syndrome which presents as pain in the large axial muscles in the shoulders, pelvis, and thighs.38,39 Importantly, recent patients data indicate that patients with intense pain in the days following paclitaxel infusion have a higher risk for the development of CIPN 3–18 weeks following the first paclitaxel infusion.39,40 This association suggests that acute increases in the excitability and/or sensitivity of nociceptive sensory neurons could underlie the chronic changes in neuronal sensitivity induced by paclitaxel. Many questions remain regarding the mechanisms by which sensitization of one population of somatic sensory neurons (those innervating the muscles) can predict enhanced sensitivity of cutaneous sensory neurons. The vinca alkaloids are used to treat acute lymphocytic leukemia, lymphomas, and neuroblastoma. The onset and incidence of vinca-induced neuropathy are similar to the taxanes; however, the vinca alkaloids are also associated with increases in autonomic neuropathy, which is rarely observed with the taxanes.41–44 The only epothilone currently used to treat cancer is ixabepilone, which is administered to patients with metastatic or locally advanced breast cancer. The incidence, dose dependence, and reversibility of neuropathy with ixabepilone treatment are similar to that of the taxanes and vinca alkaloids.45

2.2 Platinum-containing agents The platinum-derived chemotherapeutics, cisplatin, carboplatin, and oxaliplatin induce intrastrand and interstrand DNA cross-linkage, which causes the denaturation of nuclear and mitochondrial DNA.46 This interaction with DNA results in damage to the mitochondria and subsequent inhibition of ATPase activity, cellular arrest, and ultimate death of cancer cells via apoptosis and necrosis.47 In addition to cross-linkage of DNA, both cisplatin and oxaliplatin increase the intracellular production of reactive oxygen species (ROS),48,49 which may or may not contribute to the anticancer activity of the drugs. Cisplatin is administered to treat bladder, ovarian, and testicular cancer. Patients receiving cisplatin generally develop peripheral neuropathy when the cumulative dose of cisplatin reaches 250–300 mg/m2.6,18 Once

Chemotherapy-Induced Peripheral Neuropathy

475

the dosage of cisplatin exceeds 500–600 mg/m2, almost all patients demonstrate symptoms of neuropathy.50 Carboplatin, utilized for ovarian cancer, does not typically produce a significant neuropathy in patients. Doses of carboplatin up to 300 mg/m2, which are effective for anticancer treatment, elicit only a mild transient neuropathy.51,52 Oxaliplatin is used for the treatment of colon and advanced colorectal cancer. Patients treated with oxaliplatin may experience both acute and chronic peripheral neuropathies. For example, 86% of patients receiving oxaliplatin describe immediate neurological symptoms, including, but not limited to, cold-induced perioral or pharyngolaryngeal dysesthesias, jaw pain, and muscle stiffness/cramping.53,54 Approximately, 50% of patients receiving a cumulative dose of oxaliplatin of 1000 mg/m2 or higher, which is the standard treatment protocol, develop chronic peripheral neuropathy.5,55 As observed with paclitaxel, the incidence and severity of acute neurological symptoms induced by oxaliplatin can predict the incidence and severity of chronic neuropathy induced by oxaliplatin in patients.56

2.3 Proteasome inhibitors Inhibition of the ubiquitin-proteasome pathway is another mechanism that is being utilized to kill cancer cells. The proteasome pathway is the primary mechanism for the elimination of damaged or unneeded endogenous proteins in eukaryotic cells.57 This regulatory pathway is critical for numerous cellular functions, including regulation of the cell cycle and gene transcription. Inhibition of the proteasome, especially in cancer cells, increases the rate of apoptosis.58 Bortezomib is the first proteasome inhibitor to be examined in human clinical studies as an anticancer agent and is currently used to treat multiple myeloma.59 Patients receiving bortezomib generally develop chronic sharp, burning pain of moderate-to-severe intensity, self-reporting pain values of 7.8  0.7 out of 10 on a 10-point VAS scale.60 It is common that dose reductions of bortezomib must be made to accommodate for the development of chronic neuropathy symptoms.61 Bortezomib also induces an autonomic neuropathy in a subset of patients.62

2.4 Angiogenesis inhibitors In the past years, angiogenesis has been highlighted as a potential anticancer target because of the reliance of tumor growth and metastatic spread on the development of a vascular supply (see Ref. 63). Vascular endothelial growth factor (VEGF) and fibroblast growth factor 2 (FGF2) are two proangiogenic

476

Jill C. Fehrenbacher

factors which have been shown to promote tumor angiogenesis.64–66 Bevacizumab and thalidomide are two chemotherapeutics which target the signaling of VEGF and FGF2. Bevacizumab is approved for the treatment of colorectal cancer, glioblastoma, renal cell cancer, and nonsmall cell lung carcinoma, while thalidomide is used for multiple myeloma67 and investigations into efficacy for combination treatment with thalidomide for the treatment of prostate cancer are ongoing.68,69 Bevacizumab is a recombinant, humanized monoclonal antibody, which binds to and neutralizes VEGF, preventing its association with endothelial receptors. The exact anticancer mechanism of thalidomide is still debated, but it has been shown to reduce tumor levels of both VEGF and FGF2 to hinder angiogenesis.70,71 Bevacizumab has not been documented to cause extensive neuropathy as a monotherapy, but reports indicate that coadministration of bevacizumab with oxaliplatin can intensify peripheral neuropathy induced by the platin.72 Thalidomide elicits the development of peripheral neuropathy symptoms in 50% of patients within 4 months of the initiation of thalidomide treatment.73,74

3. CLINICAL ASSESSMENT OF CIPN A limitation to the interpretation of some clinical studies is the subjective nature by which the presence and intensity of neuropathy is determined by the clinicians and patients. There is much debate regarding the proper endpoints to use to determine the severity of neuropathy in patients and the intensity of neuropathy is often difficult to gage. Several grading scales have been developed by the World Health Organization, Eastern Cooperative Oncology Group, and the National Cancer Institute (NCI) to categorize the intensity of symptoms experienced by patients.75–78 Using the newly revised NCI-Common Terminology Criteria for Adverse Events (version 4.0), a grade 1 neuropathy indicates mild or asymptomatic neuropathy with loss of deep tendon reflexes or the presence of paresthesia that do not interfere with daily life, grade 2 indicates a moderate neuropathy with some interference with complex activities of daily living, grade 3 describes severe symptoms of neuropathy that limit everyday activities of daily living, and grade 4 indicates life threatening neuropathy.77–79 A limitation of these neuropathy scales, however, is that they do not provide descriptive data on the quality of neuropathic symptoms, which is especially problematic in a clinical trial setting. To overcome the subjective nature of symptom reporting, a composite measure of peripheral nerve function, the total neuropathy score

Chemotherapy-Induced Peripheral Neuropathy

477

(TNS), was established.19,80,81 The TNS is a tool that clinicians can use to objectively assess neurological endpoints of sensory and motor function in the upper and lower limbs, autonomic symptoms, vibration sensory threshold, sensitivity to pinprick, and changes in deep tendon reflexes or grip strength. In addition to clinical examination findings, the TNS also has a neurophysiological component, in which the conduction velocities in the sural and peroneal nerves are determined and factored into the total score.82 Chemotherapy sometimes induces a decrease in the action potential amplitude (SNAP) and/or conduction velocity (sCV) of sensory nerve fibers, with occasional effects on the compound muscle action potential and conduction velocity of motor nerves (CMAP and mCV, respectively).83–90 A major limitation to sensory neurophysiological studies is that the action potential amplitude and conduction velocity of sensory nerves mostly reflect the function of large-diameter sensory neurons. Dysfunction of small- and medium-diameter nerve fibers, which may contribute heavily to burning pain sensations in CIPN patients, is not detected by the recorded evoked potentials.91 Although neurophysiological outcomes do not always correlate with painful neuropathy symptoms in patients,91,92 thus questioning their relevance for the diagnosis and/or grading of CIPN, recent studies have demonstrated a strong correlation between early changes in oxaliplatininduced changes in neurophysiological outcomes and the severity of neuropathy following discontinuation of drug treatment.83 A consistent pathological finding in patients receiving chemotherapy and exhibiting numbness and loss of vibratory senses is intraepidermal nerve fiber (IENF) loss in the hands and feet. The IENF is comprised of unmyelinated axons from small-diameter sensory neurons.93 Thus, observing the degree of change in intraepidermal innervation has been measured using immunohistochemistry and novel functional techniques to specifically assess small-fiber neuropathy in CIPN patients. These novel techniques include nerve excitability-threshold tracking94 and Doppler flowmetry to determine axon reflex flare areas.95 Evaluating these endpoints is preferable for testing the ability of drugs to prevent the development of functional neuropathy, because they focus the evaluation toward the site of primary damage, the peripheral nervous system. A limitation of these techniques is that structural loss of innervation does not always correlate with the presence of painful neuropathy symptoms in patients.96 In addition, chemotherapeutics are not the only cause of IENF loss; it is also observed with diabetes mellitus,97 HIV infection98, and other nerve damage. Paradoxically, a complete loss of IENF is observed both in patients with persistent pain93 and genetic insensitivity to pain,99 suggesting

478

Jill C. Fehrenbacher

that epidermal innervation does not necessarily correlate with pain values. The question remains whether the degeneration of IENF results from changes in the sensitivity of sensory neurons or whether the IENF damage causes an increase in the sensitivity of nerve fibers, contributing to the primary symptom of neuropathy that burdens cancer patients: pain.

4. EXPERIMENTAL STUDIES: ANIMAL MODELS OF CIPN Experimental animal models have been developed to examine the mechanisms of CIPN following exposure to paclitaxel, vincristine, epothilones, cisplatin, oxaliplatin, bortezomib, or thalidomide. Although these animal models have evolved over time to closely mimic the behavioral and neurophysical symptoms observed in patients receiving chemotherapy, the initial animal studies were performed using high doses of the chemotherapeutic drugs. These high doses of drugs caused sensory neuron death, overt axonal damage and involvement of the motor and autonomic nervous systems,100–103 and produced impairments in pain-like behaviors, decreases in movement and coordination and decreases in nerve conduction velocity in the tail nerve, suggesting a functional decrease in the sensitivity of sensory neurons.104–110 Since a common complaint of patients with CIPN is augmented pain, rather than analgesia, models with high doses of drug do not optimally reflect the patient condition. The second generation of animal models utilizes lower dosage treatment protocols. This low-dose treatment paradigm elicits mechanical and cold hypersensitivity without inducing sensory neuron loss and gross axonal degeneration.106,111,112 However, the newer treatment paradigms still produce similar distal morphological changes as observed in patients receiving drug treatment, such as degeneration of the IENFs.4,113–116 Dose–symptom relationships in patients and animal models will always be complicated by the fact that the sensory neural tissues, including dorsal root ganglia (DRG), nerve roots and nerves, accumulate the chemotherapeutics,6,18,117 thus discontinuation of the drug treatments does not necessarily mean that chemotherapeutic exposure ceases. Accumulation of the drugs in the DRG has been proposed to occur via uptake in the terminals and soma and is enabled by the dense vascularization and high permeability of capillaries which surround the ganglia.118–121 The amount of drug accumulation has functional consequences, as the DRG concentration of drug is related to the degree of clinical neurotoxicity in patients.6 The onset of behavioral changes in preclinical models varies with the type of chemotherapeutic administered (Table 1). With oxaliplatin and

Table 1 Chemotherapeutics associated with the development of peripheral neuropathy Low-dose animal model Anticancer agent class

Agent

Clinical onset of sensory symptoms

Mechanical sensitivity

Cold sensitivity

Heat sensitivity

IENF density

Microtubuletargeting agents

Taxanes

Acute and chronic

"111

"111

"111

#115

Vinca alkaloids

Chronic

"112,140

"140

#112

#115

Epothilones

Chronic

ND

ND

ND

#130

DNA-cross-linking agents

Cisplatin139,144,145

Chronic

"106,128

No change145

",128 No change106

#93

Oxaliplatin

Acute and chronic

"12

"12

No change12

#12

Proteasome inhibitor

Bortezomib

Chronic

"116

"116

No change116

#116

Angiogenesis inhibitors

Thalidomide

Chronic

Animal model not yet developed

Bevacizumab

Chronic

Animal model not yet developed

ND: not determined.

480

Jill C. Fehrenbacher

paclitaxel, some studies show an acute behavioral effect within 60 min, whereas others observe effects at 24–48 h after initial dosing.122–126 Longterm effects generally present around 6–10 days following the first drug injection and persist for at least 3–7 weeks following the last injection.105,127,128 Consistent modalities affected by various chemotherapeutics include mechanical allodynia and cold allodynia.105,106,111,112 Sensitivity to warm/heat is quite variable, depending on the type and dose of drug administration and on the investigating laboratory.101,105,111,129 The neurophysiological effects of chemotherapeutic treatment in animal models mimics that observed in patients. Lower doses of chemotherapeutics generate functional effects which are limited to sensory fibers, whereas higher doses also involve motor and autonomic nerves. Endpoints utilized to measure the effects of chemotherapeutics on nerve activity include: nerve conduction velocity,130–133 amplitude of action potentials,134 and changes in blood flow induced by neurostimulation.135 Investigators examining the neurophysiological effects of chemotherapy have reported a slowing of the sensory nerve conduction velocity (SNCV)130,131,136 and an increase in the action potential amplitude when examining Aδ fibers137; however, the effects on C fibers are more variable, with high-dose vincristine causing a decrease in the SNCV136 and high-dose paclitaxel causing no change in the SNCV.122 The data from skin-nerve experiments support an effect of high-dose paclitaxel to decrease the SNCV in all types of sensory neurons (Aβ, Aδ, and C fibers).134 As observed in patients receiving lower doses of chemotherapeutics, the morphological changes in the axonal diameter or myelination of the nerve bundle following lower dose treatment of animals with the various chemotherapeutics are unimpressive.12,111,116,138–140 However, lower doses of chemotherapeutics, which cause an enhanced sensitivity to nociceptive stimuli, cause a degeneration or loss of IENF, similar to that observed in patients.115,116,141–143 One advantage of using animal models is that studies can be designed to determine whether this IENF loss is a cause or an effect of changes in neuronal sensitivity following chemotherapy treatment, although, to date, these studies have not been performed. The data from the animal models suggest that both increases and decreases in neuronal sensitivity occur following treatment with chemotherapeutics, depending on the endpoint measured and the dosing paradigm administered. There is still a lack of understanding, however, as to what population of neurons is affected and the mechanisms by which these changes in sensitivity develop (Figure 1).

Chemotherapy-Induced Peripheral Neuropathy

481

Figure 1 Putative sites of chemotherapeutic toxicity in the peripheral nervous system.

5. EXPERIMENTAL STUDIES: IN VITRO MODELS OF CIPN To accurately delineate the intracellular signaling mechanisms by which chemotherapeutics alter the sensitivity of sensory neurons, investigators have broadened their research to include in vitro models of CIPN. Cultures of sensory neurons provide a tightly controlled system which allows relatively easy manipulation of protein expression and function via genetic and pharmacological tools. The limitation of neuronal cultures is that the neurons have been removed from their native environment and thus, communication with other cell types (direct or indirect), is largely altered. Furthermore, the addition of growth factors to the growth media could alter the effect of chemotherapeutics on neuronal sensitivity and outgrowth (this will be discussed below). In vitro experiments with chemotherapeutics have largely been limited to morphological studies to determine the effects of the drugs on neurite lengths in DRG cultures.130,146–151 Chemotherapyinduced toxicity, as measured by decreasing neurite length, is dependent on distal axonal exposure to the drug.148,150 For example, local application of paclitaxel or vincristine to the soma of cultured neurons does not affect the neurite length, whereas application to the distal terminals causes a shortening of the neurites. It is known that alterations of microtubule dynamics can lead to both neurite retraction and inhibition of neurite outgrowth by altering

482

Jill C. Fehrenbacher

growth cones.152,153 The specific mechanisms underlying chemotherapyinduced decreases in IENFs are under investigation. Regardless of the specific mechanism, it is unclear whether changes in neurite length correlate with changes in the sensitivity of sensory neurons in culture and mechanistic in vitro studies to examine the effects of chemotherapeutics on the sensitivity of subpopulations of sensory neurons are limited. Several investigators have isolated the lumbar DRG from chemotherapy-treated animals to examine the effects of the drugs on electrical excitability and intracellular calcium signaling within the soma of the lumbar neurons, which predominantly innervate the hindpaws.154–156 Increases in the excitability of mediumand large-sized neurons isolated from paclitaxel-treated rats were observed: the percentage of neurons which exhibited spontaneous firing increased, while the current threshold for activation of the remaining quiescent neurons was decreased.156 The contribution of voltage-gated calcium currents to these increases in excitability were investigated by Kawakami and colleagues, who demonstrated that paclitaxel treatment, enhanced voltagedependent calcium current in both small- and medium-sized neurons.155 When neurons were isolated from paclitaxel-treated rats and examined for ATP-stimulated increases in intracellular calcium accumulation; however, a decrease in stimulated calcium levels was observed.154 It is unclear whether stimulation of intracellular calcium accumulation with a general depolarizing stimulus, such as high extracellular potassium, would elicit the same decrease in intracellular calcium. Overall, these data suggest a chemotherapy-induced increase in the excitability of sensory neurons. An effect of chemotherapeutics to modulate integrated neuronal responses, such as the stimulated release of neurotransmitters from the neurites and soma of sensory neurons, has been investigated for paclitaxel.157 Paclitaxel alters the sensitivity of isolated sensory neurons in concentration- and time-dependent manner without altering the viability of the sensory neurons. Paclitaxel can enhance or reduce TRPV1- and TRPA1-mediated release of calcitonin gene-related peptide (CGRP), depending on the magnitude and duration of exposure to the chemotherapeutic. However, paclitaxel augments CGRP release evoked by high extracellular potassium, regardless of the concentration or duration of paclitaxel exposure.149 Exposing sensory neurons in culture to cisplatin results in a concentration-dependent decrease in the TRPV1-mediated release of CGRP, even after accounting for the loss of content via cell death.158 These experiments investigate long-term effects of chemotherapeutic exposure on neuropeptide release, but some studies demonstrate a direct excitatory effect

Chemotherapy-Induced Peripheral Neuropathy

483

of chemotherapeutics on neuropeptide release at higher drug concentrations, suggesting the possibility for an acute effect of the drugs on neuronal sensitivity.159,160 To date, the different effects of the various classes of chemotherapeutics on neuropeptide release suggest that each class of drugs likely has a different mechanism of action to elicit changes in neuronal sensitivity. Using neurophysiological endpoints, such as the release of neurotransmitters from sensory neurons, which integrate the contributions of possible changes in function of ion channels, mitochondria, DNA damage, oxidative stress, and axonal transport, will be critical to determining whether all classes of chemotherapeutics alter neuronal sensitivity via the same mechanisms and identifying the specific signaling mechanisms by which neuronal sensitivity is altered by chemotherapeutics.

6. PROPOSED MECHANISMS UNDERLYING CIPN There have been multiple theories on the mechanisms by which each of the chemotherapeutics alters neuronal function. In this section, the use of pharmacological interventions that have been used for other types of neuropathic pain will be discussed. In addition, various putative mechanisms will be discussed in the context of the clinical and experimental evidence supporting or refuting that mechanism. It is appreciated that many of these proposed mechanisms may overlap, and thus, it is probable that neuronal function is altered via a combination of the proposed cellular mechanisms. Despite differences in disease pathophysiology and symptoms experienced by patients with CIPN,161 the initial approach to investigate CIPN focused on administering drugs utilized for other pain syndromes to treat the painful symptoms of peripheral neuropathy. As such, there have been many clinical studies to examine the effects of opioids,60,162 antidepressants,163–165 and antiepileptics166,167 on the symptoms of neuropathy induced by chemotherapeutics. Some of these drugs have been investigated and shown to be efficacious in reversing mechanical and/ or cold allodynia in animal models of CIPN, including opioids,168 antidepressants,169–171 and antiepileptics.172,173 Of all the aforementioned drugs, only duloxetine, a serotonin/norepinephrine reuptake inhibitor, has been demonstrated to have clinical efficacy in the treatment of chronic neuropathy symptoms induced by oxaliplatin administration.165 In fact, a recent set of guidelines published by the American Society for Clinical

484

Jill C. Fehrenbacher

Oncologists moderately recommends duloxetine for the treatment of established CIPN, but does not endorse any other treatments for the prevention or treatment of CIPN.174

6.1 Mitochondrial dysfunction Arguably, the most investigated putative mechanism by which most chemotherapeutics are thought to alter neuronal function is via the perturbation of neuronal mitochondrial function. Changes in mitochondrial localization, altered fission, and fusion rates and changes in mitochondrial membrane permeability have all been investigated following neuronal exposure to chemotherapeutics. These alterations in mitochondria are associated with changes in calcium handling, release of cytochrome c, mitochondrial DNA damage, and enhanced production of ROS, which can all contribute to neurotoxicity. The localization of mitochondria is dependent on axonal transport via cytoskeletal components, including microtubules and actin filaments,175 and paclitaxel and cisplatin have been shown to impair the axonal movement of organelles, including mitochondria, in neurons or neuronal cells lines.176–178 Multiple different chemotherapeutics have been shown to induce mitochondrial swelling within sensory neuronal axons,116,138,179 sensory neuronal soma,180 and Schwann cells.127 Mitochondrial swelling occurs in normal neurons, but an increased incidence of swelling is associated with changes in mitochondrial function, and can be a result of mitochondrial permeability transition.181,182 Multiple chemotherapeutics have been shown to promote mitochondrial permeability transition and depolarize the mitochondria,146,183–185 theoretically resulting in a decrease in the ability of the mitochondria to respire and generate ATP for the neuron.116,117,182 Depolarization of the mitochondria also enhances the production of ROS,186 which is another putative mechanism by which chemotherapeutics alter neuronal sensitivity (see below). An alternate mechanism by which chemotherapeutics may alter mitochondrial localization and energy production is through the disruption of the balance between mitochondrial fission and fusion. Fission, the defragmentation of mitochondria, leads to a decrease in energy production and enhances the production of ROS.187 Although a change in mitochondrial fission was not observed in DRG from cisplatin-treated rats,188 paclitaxel altered mitochondrial dynamics in differentiating neuroblastoma cells.189 Because of the demonstrated effects of anticancer drugs, of multiple mechanisms, on the localization and function of mitochondria, several

Chemotherapy-Induced Peripheral Neuropathy

485

different mitoprotective drug strategies have been examined to determine whether preventing mitochondrial dysfunction might reverse the signs and symptoms of neurotoxicity. The premiere mitoprotective candidate was acetyl-L-carnitine (ALCAR). ALCAR plays a role in intermediary metabolism and is thought to be neuroprotective via an increase in fatty acid β-oxidation to subsequently increase ATP production.190 ALCAR has also been shown to reverse the depletion of plasma nerve growth factor (NGF) levels induced by cisplatin treatment.131 In preclinical experiments, ALCAR administration reversed mechanical and cold allodynia induced by paclitaxel,131,191 cisplatin,131 and bortezomib.116 ALCAR also reversed the chemotherapeutic-induced swelling and dysfunction of mitochondria.116,131,179 These preclinical data and evidence that ALCAR have therapeutic effects on neuropathy in patients with diabetes192 and HIV,193,194 prompted clinical trials to examine the effects of ALCAR on paclitaxelinduced neuropathy in patients. In a phase II study with 25 patients, the effects of ALCAR were promising, providing improvement of sensory neuropathy in 60% of patients.195 In order to confirm these hopeful findings, a prospective randomized, double-blind, placebo-controlled trial of 409 breast cancer patients receiving taxane-based adjuvant therapy were administered ALCAR. In contrast to the findings of the small-clinical trial, ALCAR treatment in the large trial increased CIPN at 24 weeks following treatment, as determined by patient self-reporting and clinician grading of sensory and motor neuropathy adverse effects.196 Another putative mitoprotective drug, olesoxime, has also been investigated to determine whether it reverses hypersensitivity in animal models of CIPN. Olesoxime binds directly to two different components of the mitochondrial membrane permeability transition pore197 to regulate depolarization of the mitochondria. In addition, treatment with olesoxime restores mitochondrial motility, compromised by treatment with paclitaxel, suggesting that olesoxime can interact with microtubule function.198 Systemic administration of olesoxime reversed mechanical allodynia in animal models of vincristine, paclitaxel, and oxaliplatin-induced CIPN,12,143,197 but did not attenuate the spontaneous activity of the peripheral C- and Aδ-fibers induced by the drugs, suggesting that the site of action for olesoxime is in the CNS.143 The translation of these protective preclinical effects of olesoxime into a clinically relevant therapeutic for CIPN has yet to be determined. A phase II clinical study (NCT00876538) to examine the effects of olesoxime on paclitaxel-induced pain and/or dysesthesia has been completed, but the study outcomes have not been released.

486

Jill C. Fehrenbacher

Despite observations in experimental animals after anticancer drug treatment, the question of mitochondrial involvement in CIPN remains controversial. It is unclear whether mitochondrial dysfunction precedes changes in neuronal sensitivity and/or changes in neuronal morphology. Furthermore, mitochondrial changes such as those listed above have also been associated with general nerve injury199 or insult, such as that resulting from streptozotocin treatment as a preclinical model for diabetes mellitus,200 suggesting that damage to mitochondria might be a common downstream effect of nonspecific damage to sensory neurons rather than a causative mechanism by which chemotherapeutics alter nerve function.

6.2 Nitroxidative stress Chemotherapeutics have also been shown to enhance the production of both ROS and reactive nitrogen species (RNS) in sensory neurons and in the spinal cord.158,201,202 In fact, the generation of nitroxidative stress, induced by ROS and RNS, has been shown to enhance the sensitivity of nociceptive sensory neurons.203,204 Chemotherapy-induced ROS and RNS are generated by neurons, supporting satellite cells and inflammatory cells.205,206 Intracellular generators of ROS and RNS include the mitochondria, via the electron transport chain and an increased incidence of fission, and an increased activity of plasma membrane NADPH oxidase.207 Unfortunately, nitroxidative stress can compromise the survival and function of sensory neurons and supporting cells by causing DNA damage, nerve fiber demyelination, mitochondrial damage and dysfunction, activation of signal transduction pathways, and neuronal death via apoptosis.208–210 Multiple different strategies have been employed preclinically to try reverse chemotherapy-induced nociceptive behaviors with scavengers of ROS (phenyl-N-tert-butylnitrone and TEMPOL) or peroxynitrite decomposition catalysts (FeTMPyP5+ and MnTE-2-PyP5+),201,210–212 and these experiments demonstrate moderate attenuation of chemotherapy-induced hindpaw hypersensitivity to mechanical or cold stimulation. The effects of these compounds on nerve conduction velocity or IENF loss, as other indicators of peripheral neuropathy, have not yet been explored. Glutathione is another antioxidant which facilitates the reduction of the enzyme, glutathione peroxidase, to reduce lipid and hydrogen peroxides.213 N-acetylcysteine (NAC) functions to regenerate the reduced form of glutathione for continued reduction of ROS.214 In preclinical models of CIPN, exogenous administration of glutathione and NAC reduced the severity of neuropathy using nociceptive behaviors,202 sensory nerve conduction

Chemotherapy-Induced Peripheral Neuropathy

487

velocities,215,216 and neuronal morphology217 as endpoints. Following these preclinical studies, several small trials with glutathione and NAC were conducted in cancer patients receiving either cisplatin or oxaliplatin. The trials all reported a reduction in neurotoxicity, albeit the degree of protection varied greatly.218–224 In a large randomized, double-blind, placebo-controlled trial to investigate the ability of glutathione to reverse the neuropathy associated with paclitaxel/carboplatin combination treatment; however, glutathione did not prevent CIPN.225 Several small-clinical trials also demonstrated that vitamin E might have efficacy to decrease the incidence or severity of CIPN following paclitaxel or cisplatin treatment226–229; however, a phase III clinical trial failed to reproduce the beneficial effects of the putative antioxidant.230 There are some concerns about further development of antioxidant therapies to treat CIPN, since the production of ROS and RNS has been shown to be essential for the anticancer effects of chemotherapeutics.49,231

6.3 DNA damage Multiple strategies have been utilized to demonstrate the importance of DNA damage in mediating neurotoxicity induced by the platinum-derived compounds. Enhancing the activity of DNA repair enzymes, involved in either base excision or nucleotide excision repair pathways, mitigates the neurotoxicity associated with exposure to cisplatin.158,180,232 Whether enhanced repair of the nuclear or mitochondrial DNA adducts is critical for the protection against neurotoxicity is still under investigation.180 Mitochondrial DNA damage by cisplatin has been shown to induce the generation of ROS, which can be attenuated by overexpression of APE1, a DNA repair enzyme of the base excision repair pathway.158 Interestingly, carboplatin, which causes the formation of nuclear DNA adducts, but does not generate ROS or mitochondrial DNA damage,49 induces a very mild neuropathy,48,233 leading to speculation that increases in ROS generation secondary to mitochondrial DNA damage is a critical component of neuropathy induced by cisplatin and oxaliplatin.

6.4 Ion channel modulation Changes in ion channel conductance are also a putative target for chemotherapeutics to alter neuronal sensitivity. Both voltage- and ligand-gated ion channels have been shown to play a role in chemotherapy-induced changes in nociceptive behaviors. Included in these are sodium

488

Jill C. Fehrenbacher

channels,137,234,235 calcium channels,155,173,236–238 ATP-gated channels,239 acid-sensing channels,239 and TRP channels,240–242 although much controversy persists regarding which channels mediate different modalities of chemotherapy-induced behaviors. Gabapentin, which binds to the α2δ subunit of the calcium channel,243 has been investigated in preclinical studies and found to have variable effects on nociceptive behaviors, depending on the class of chemotherapy administered. Sensitivity to gabapentin was correlated with chemotherapy-induced upregulation of the α2δ subunit of the calcium channel in the DRG and spinal cord.244,245 Both taxanes and platinum-containing agents elicit an upregulation of the α2δ subunit, and nociceptive behaviors induced by treatment with these chemotherapeutics are reversed by gabapentin administration.12,105,237,244 In contrast, vinca alkaloids do not upregulate the expression of the α2δ subunit, and nociceptive behaviors associated with vinca treatments are not reversed by gabapentin.244,245 There has been only one clinical trial to examine whether gabapentin administration reverses established CIPN in patients receiving vinca alkaloids, taxanes, or platinum-containing agents.166 The results of the trial failed to show an effect of gabapentin to reverse pain as a CIPN symptom.166 Given the dependence of gabapentin efficacy on the class of chemotherapy administered in preclinical studies, it would be interesting to determine whether gabapentin could reverse CIPN in a study with sufficient power to be able to analyze the outcomes based upon the class of chemotherapy which elicited the CIPN. Alternative strategies are being developed to target the pathophysiological function of specific ion channels without complete blockade of ion channel conductance, via the modulation of ion channelbinding proteins.246 This strategy, which is promising in the treatment of pain derived from various etiologies,247 might be a useful approach for the treatment of CIPN once the chemotherapy-induced changes in ion channels are understood. Another cause of ion channel dysfunction, specifically following treatment with oxaliplatin,235,248 is the chelation of extracellular calcium and magnesium by oxalate, a metabolic product of oxaliplatin.249 It was hypothesized that replacement of these ions, via Ca2+/Mg2+ infusions, could help to reverse acute and chronic oxaliplatin-induced neuropathy symptoms.250 Preclinical experiments demonstrated prevention of the hypersensitivity to cold,105,126 but ion replacement did not alter mechanical hypersensitivity induced by either oxaliplatin or oxalate.126 An initial retrospective review of patients receiving oxaliplatin in the absence or presence of Ca2+/Mg2+ infusions supported the theory that Ca2+/Mg2+ therapy prevented

Chemotherapy-Induced Peripheral Neuropathy

489

neuropathy, as patients receiving Ca2+/Mg2+ were administered larger cumulative doses of oxaliplatin (650 vs. 910 mg/m2 for placebo vs. Ca2+/Mg2+, respectively), and reported less acute neuropathic symptoms (laryngopharangeal dysesthesia in 9% vs. 0% of patients), and less development of grade 3 chronic neuropathy (20% vs. 8% of patients).250 Based on this retrospective data, two clinical trials were initiated to examine the efficacy of Ca2+/Mg2+ infusions to prevent oxaliplatin-induced neuropathy (CONcePT trial and N04C7). Trial progress was halted due to concerns that Ca2+/Mg2+ infusions reduced the oncolytic activity of oxaliplatin,251 however, those concerns have since been refuted.252 The data generated from the two discontinued trials were conflicting; in the CONcePT trial, Ca2+/Mg2+ infusions did prevent acute symptoms or chronic neuropathy253; whereas in the early results of the N04C7 trial,254 Ca2+/Mg2+ infusions decreased oxaliplatin-induced acute muscle spasms and the development of grade 2 or higher chronic neuropathy.254 A third trial was initiated (N08CB) to resolve whether Ca2+/Mg2+ infusions were efficacious to prevent oxaliplatin-induced neuropathy.255 An effect of infusions to prevent the development of acute or chronic oxaliplatin-induced neuropathy was not observed; however, there have been several critiques of the data interpretation from the N08CB study. Under questions are the selected choice of sensory scales used as endpoints for neuropathy and the power of the study to determine the reported negative efficacy outcomes.256,257 In light of the preclinical data, which demonstrate an effect of Ca2+/Mg2+ to reverse cold, but not mechanical, hypersensitivity, it would be useful to analyze the clinical data based on individual sensory symptoms. In this manner, information regarding whether Ca2+/Mg2+ has efficacy to prevent specific modalities of the neuropathy could be obtained.

6.5 Inflammation Another plausible mechanism by which chemotherapeutic treatments alter the sensitivity of sensory neurons is through the activation of the immune system and subsequent induction of inflammation. Several investigators have observed the activation of resident immune cells, such as the satellite cells in the DRG258 and the microglia in the dorsal horn of the spinal cord188,259 following chemotherapeutic administrations. Infiltration of macrophages in the DRG and antigen-presenting Langerhans cells in the skin has also been observed following the administration of paclitaxel.115,260 Furthermore, modulation of known inflammatory signaling pathways, such as

490

Jill C. Fehrenbacher

NFκB141 and Toll-like receptor (TLR) signaling261,262 attenuate the behavioral effects of chemotherapy administration in preclinical models. Indeed, paclitaxel administration to rats enhances the expression of the TLR4 in the DRG and spinal cord.262 Intrathecal administration of a TLR4 antagonist, lipopolysaccharide derived from Rhodobacter sphaeroides, partially reversed the mechanical hypersensitivity associated with paclitaxel treatment.262 Several investigators have also demonstrated an upregulation of chemokine (C– C motif ) ligand 2 (CCL2 or MCP-1) following paclitaxel treatment.259,263 This increase in chemokine levels was shown to have functional importance, since intrathecal inhibition of the receptor for CCL2, the C–C chemokine receptor type 2 (CCR2), reverses mechanical allodynia, and the degeneration of IENF induced by paclitaxel administration.263 The signaling pathways by which paclitaxel upregulates CCL2 levels have not been elucidated, but CCL2 is enhanced primarily in small-diameter sensory neurons within the DRG and in astrocytes of the dorsal horn.263 The physiological function of CCL2 is to recruit and activate monocytic cell types, including macrophages and microglia. This function is apparent in the dorsal horn of the spinal cord, where a paclitaxel-induced upregulation of CCL2 activates microglia, reversible by intrathecal application of an anti-CCL2 antibody.259,263 What remains unclear, however, is how paclitaxel-induced activation of the CCL2/CCR2 pathway, which occurs in the entire dorsal column regardless of axonal length,263 can initiate neuronal sensitivity in a “stocking and glove” distribution. Further investigations into possible interactions between inflammation and axonal damage, both induced by chemotherapeutic treatment, to elicit symptoms of CIPN might uncover additional therapeutic targets to prevent or treat the neuropathy.

6.6 Neurotrophic factors An approach to reverse the symptoms of CIPN has been to ameliorate the neuropathy through the addition of neurotrophic factors in an attempt to overcome the nerve damaging effects of chemotherapeutics with growth promoting factors. The first neurotrophin to be investigated was NGF. NGF was discovered in the late 1950s as a nerve growth promoting factor and is essential for the survival of spinal sensory ganglion cells and sympathetic neurons (see review by Zhang et al. 264). NGF levels have been shown to correlate with the intensity of neuropathy symptoms induced by paclitaxel in the clinic and in the laboratory. In patients receiving paclitaxel and cisplatin therapy, treatment-induced decreases in circulating NGF levels were associated with the development of CIPN.265 Although this

Chemotherapy-Induced Peripheral Neuropathy

491

chemotherapy-induced decrease in circulating NGF levels could not be reproduced in preclinical models,110 exogenous administration of NGF with high-dose paclitaxel reversed paclitaxel-induced thermal hypoalgesia.266 In studies investigating the effects of paclitaxel and NGF on cell survival and neurite outgrowth, NGF treatment reversed neuronal degeneration, and death in embryonic DRG explants exposed to paclitaxel (1 μM) for 1–4 days.147,267 Together, these data suggest a protective effect of NGF against neurotoxicity induced by exposure to chemotherapeutics; however, NGF treatment to prevent or treat CIPN has not been tried in the clinic. One limitation to the use of NGF as a treatment for CIPN is the effect that the neurotrophin has on the sensitivity of small-diameter sensory neurons. Inflammation-induced increases in the levels of NGF have been shown to mediate the thermal and mechanical hypersensitivity associated with inflammation,268,269 thus NGF treatment may cause inflammatory-like pain in cancer patients who are already suffering from neuropathy. Indeed, antiNGF pharmacological agents have been investigated for the treatment of inflammatory pain.270 Another limiting factor in the translation of preclinical NGF effects is centered on the interpretation of the experimental data. Embryonic neurons, which were used in many of the preclinical studies, are dependent on NGF for survival.271 NGF is trafficked from the nerve terminals to the nucleus and this retrograde transport is blocked by the MTA, colchicine.272,273 Thus, the decrease in survival observed with paclitaxel exposure could be a result of possible NGF deprivation due to decreased axonal transport, rather than a positive effect of NGF to protect the neurons.274,275 Furthermore, since embryonic neurons are dependent on NGF for survival, it is difficult to interpret the survival findings, since low concentrations of paclitaxel which alter neuronal sensitivity do not commonly cause neuronal death in cultures of DRG neurons derived from adult animals.157 Other neurotrophic agents which have been investigated and have had positive results in preclinical animals include: prosaptide,129 xaliproden,276 retinoic acid,277 and recombinant human glial growth factor 2278, but so far none have demonstrated appreciable efficacy in clinical trials.279 Neurotrophic factors play an integral role not only in the survival of sensory neurons but also in the maintenance of the neurons, thus helping to establish the setpoints for neuronal sensitivity and growth.280,281 Targeting this class of compounds, without a comprehensive understanding of how chemotherapy alters intracellular signaling pathways and the axonal transport mechanisms used to traffic the products of neurotrophin signaling, is proving to be very challenging.

492

Jill C. Fehrenbacher

6.7 Microtubule alterations An obvious mechanism by which MTAs may alter neuronal function is through the perturbation of microtubule function. Microtubules are dynamic cellular structures that have critical roles in many functions including cell signaling, mitosis and cell division, intracellular organelle and vesicular transport, cell shape and motility, and intracellular organization.32,282 The dynamic nature of microtubules, with constant growth and shortening, is necessary for microtubule physiology.283 In nondividing neurons, the MTAs have the potential to induce neurotoxicity in a very specific way by altering axonal transport along the microtubules,11,284,285 by changing the posttranslational status of microtubules,286–288 and by changing protein–protein interactions between the microtubules and cellular proteins.289–291 Bortezomib has also been shown to change microtubule dynamics and thus could mediate changes in neuronal sensitivity via a mechanism similar to the MTAs.292 Unfortunately, the anticancer efficacy of the MTAs is directly proportional to their ability to alter microtubule dynamics,293 thus modulating the microtubule interactions to prevent CIPN would most likely compromise the anticancer properties of the drugs.

7. CHALLENGES TO CIPN RESEARCH There has been very little success in translating preclinical findings to efficacious therapies for patients suffering with CIPN. One caveat that has been largely unaddressed in CIPN research is the possibility that the presence of cancer may predispose patients to neuropathy or induce novel interactions with anticancer drugs to “prime” patients to develop a more robust neuropathy than that which is observed in preclinical models. The presence of cancer alone has been shown to induce central nervous system toxicity, using deficits in cognitive dysfunction as an endpoint for patients.294 To address the possibility that the presence of cancer may also contribute to peripheral neurotoxicity, researchers are shifting their preclinical models from rats to mice.294 The mouse is well developed as a disease model for cancer research and could facilitate the necessary experiments to examine the development of neuropathy in the absence and presence of anticancer treatment in tumorburdened animals. While a possible interaction between cancer and anticancer drugs in preclinical models could underlie the paucity of translational successes from the lab to the clinic, there are other putative pitfalls that may also contribute. CIPN research was originally focused on the

Chemotherapy-Induced Peripheral Neuropathy

493

development of animal models; efforts then shifted to try to reverse evoked nociceptive behaviors induced by anticancer treatment with a variety of “pain” drugs to see if they would alleviate nociceptive behaviors in the animals. Currently, the nociceptive tests that are used to discriminate the symptoms of neuropathy and the success and/or failure of putative chemotherapy drugs in preclinical models, do not assess this modality of spontaneous pain. Indeed, this small-fiber neuropathy is proving to be very difficult to study. Research by Kalliomaki and colleagues has demonstrated that the endpoints commonly used as surrogates for neuronal sensitivity in preclinical models: thermal behaviors, axon flare reflexes, and IENF loss, do not necessarily correlate with pain levels in patients who experience pain as a symptom of the neuropathy,96 suggesting that using appropriate endpoints to monitor neuropathy, even in in vitro studies, is critical for the translational ability of the research. The neuropathy induced by chemotherapeutics is multifaceted and deriving the mechanisms for the various symptoms will take much effort. One of the distinct differences between CIPN and neuropathies induced by physical nerve damage is that the timing of the insult is defined. This predictability allows for the introduction of pharmacological agents to inhibit the neuronal effects of the chemotherapeutics. Only after, we understand the neuronal signaling pathways altered by chemotherapeutics; however, we can subsequently develop therapeutics to specifically prevent the development or maintenance of neurotoxicity without compromising the oncolytic activity of the drug.

REFERENCES 1. Burton AW, Fanciullo GJ, Beasley RD, Fisch MJ. Chronic pain in the cancer survivor: a new frontier. Pain Med. 2007;8(2):189–198. 2. Wolf S, Barton D, Kottschade L, Grothey A, Loprinzi C. Chemotherapy-induced peripheral neuropathy: prevention and treatment strategies. Eur J Cancer. 2008;44(11):1507–1515. 3. Iniguez C, Larrode P, Mayordomo JI, et al. Reversible peripheral neuropathy induced by a single administration of high-dose paclitaxel. Neurology. 1998;51(3):868–870. 4. Boyette-Davis JA, Cata JP, Driver LC, et al. Persistent chemoneuropathy in patients receiving the plant alkaloids paclitaxel and vincristine. Cancer Chemother Pharmacol. 2013;71(3):619–626. 5. Krishnan AV, Goldstein D, Friedlander M, Kiernan MC. Oxaliplatin-induced neurotoxicity and the development of neuropathy. Muscle Nerve. 2005;32(1):51–60. 6. Gregg RW, Molepo JM, Monpetit VJ, et al. Cisplatin neurotoxicity: the relationship between dosage, time, and platinum concentration in neurologic tissues, and morphologic evidence of toxicity. J Clin Oncol. 1992;10(5):795–803. 7. Forsyth PA, Balmaceda C, Peterson K, Seidman AD, Brasher P, DeAngelis LM. Prospective study of paclitaxel-induced peripheral neuropathy with quantitative sensory testing. J Neurooncol. 1997;35(1):47–53.

494

Jill C. Fehrenbacher

8. Lipton RB, Apfel SC, Dutcher JP, et al. Taxol produces a predominantly sensory neuropathy. Neurology. 1989;39(3):368–373. 9. Postma TJ, Vermorken JB, Liefting AJ, Pinedo HM, Heimans JJ. Paclitaxel-induced neuropathy. Ann Oncol. 1995;6(5):489–494. 10. Park SB, Lin CS, Krishnan AV, Friedlander ML, Lewis CR, Kiernan MC. Early, progressive, and sustained dysfunction of sensory axons underlies paclitaxel-induced neuropathy. Muscle Nerve. 2011;43(3):367–374. 11. Theiss C, Meller K. Taxol impairs anterograde axonal transport of microinjected horseradish peroxidase in dorsal root ganglia neurons in vitro. Cell Tissue Res. 2000;299(2):213–224. 12. Xiao WH, Zheng H, Bennett GJ. Characterization of oxaliplatin-induced chronic painful peripheral neuropathy in the rat and comparison with the neuropathy induced by paclitaxel. Neuroscience. 2012;203:194–206. 13. Cavaletti G, Tredici G, Marmiroli P, Petruccioli MG, Barajon I, Fabbrica D. Morphometric study of the sensory neuron and peripheral nerve changes induced by chronic cisplatin (DDP) administration in rats. Acta Neuropathol. 1992;84(4):364–371. 14. Rowinsky EK, Chaudhry V, Cornblath DR, Donehower RC. Neurotoxicity of Taxol. J Natl Cancer Inst Monogr. 1993;15:107–115. 15. Wiernik PH, Schwartz EL, Strauman JJ, Dutcher JP, Lipton RB, Paietta E. Phase I clinical and pharmacokinetic study of taxol. Cancer Res. 1987;47(9):2486–2493. 16. Dougherty PM, Cata JP, Cordella JV, Burton A, Weng HR. Taxol-induced sensory disturbance is characterized by preferential impairment of myelinated fiber function in cancer patients. Pain. 2004;109(1–2):132–142. 17. Nabholtz JM, Gelmon K, Bontenbal M, et al. Multicenter, randomized comparative study of two doses of paclitaxel in patients with metastatic breast cancer. J Clin Oncol. 1996;14(6):1858–1867. 18. Thompson SW, Davis LE, Kornfeld M, Hilgers RD, Standefer JC. Cisplatin neuropathy. Clinical, electrophysiologic, morphologic, and toxicologic studies. Cancer. 1984;54(7):1269–1275. 19. Chaudhry V, Rowinsky EK, Sartorius SE, Donehower RC, Cornblath DR. Peripheral neuropathy from taxol and cisplatin combination chemotherapy: clinical and electrophysiological studies. Ann Neurol. 1994;35(3):304–311. 20. Lee JJ, Swain SM. Peripheral neuropathy induced by microtubule-stabilizing agents. J Clin Oncol. 2006;24(10):1633–1642. 21. Windebank AJ, Grisold W. Chemotherapy-induced neuropathy. J Peripher Nerv Syst. 2008;13(1):27–46. 22. Rowinsky EK, Gilbert MR, McGuire WP, et al. Sequences of taxol and cisplatin: a phase I and pharmacologic study. J Clin Oncol. 1991;9(9):1692–1703. 23. Siegal T, Haim N. Cisplatin-induced peripheral neuropathy. Frequent off-therapy deterioration, demyelinating syndromes, and muscle cramps. Cancer. 1990;66(6):1117–1123. 24. von Schlippe M, Fowler CJ, Harland SJ. Cisplatin neurotoxicity in the treatment of metastatic germ cell tumour: time course and prognosis. Br J Cancer. 2001; 85(6):823–826. 25. Brydoy M, Oldenburg J, Klepp O, et al. Observational study of prevalence of longterm Raynaud-like phenomena and neurological side effects in testicular cancer survivors. J Natl Cancer Inst. 2009;101(24):1682–1695. 26. Park SB, Lin CS, Krishnan AV, Goldstein D, Friedlander ML, Kiernan MC. Long-term neuropathy after oxaliplatin treatment: challenging the dictum of reversibility. Oncologist. 2011;16(5):708–716. 27. Haim N, Epelbaum R, Ben-Shahar M, Yarnitsky D, Simri W, Robinson E. Full dose vincristine (without 2-mg dose limit) in the treatment of lymphomas. Cancer. 1994;73(10):2515–2519.

Chemotherapy-Induced Peripheral Neuropathy

495

28. Moser EC, Noordijk EM, Carde P, et al. Late non-neoplastic events in patients with aggressive non-Hodgkin’s lymphoma in four randomized European Organisation for Research and Treatment of Cancer trials. Clin Lymphoma Myeloma. 2005;6(2):122–130. 29. Strumberg D, Brugge S, Korn MW, et al. Evaluation of long-term toxicity in patients after cisplatin-based chemotherapy for non-seminomatous testicular cancer. Ann Oncol. 2002;13(2):229–236. 30. Ramchandren S, Leonard M, Mody RJ, et al. Peripheral neuropathy in survivors of childhood acute lymphoblastic leukemia. J Peripher Nerv Syst. 2009;14(3):184–189. 31. Jordan MA, Wendell K, Gardiner S, Derry WB, Copp H, Wilson L. Mitotic block induced in HeLa cells by low concentrations of paclitaxel (Taxol) results in abnormal mitotic exit and apoptotic cell death. Cancer Res. 1996;56(4):816–825. 32. Wilson L, Jordan MA. New microtubule/tubulin-targeted anticancer drugs and novel chemotherapeutic strategies. J Chemother. 2004;16(suppl 4):83–85. 33. Johnson IS, Armstrong JG, Gorman M, Burnett Jr JP. The vinca alkaloids: a new class of oncolytic agents. Cancer Res. 1963;23:1390–1427. 34. Sartorelli AC, Creasey WA. Cancer chemotherapy. Annu Rev Pharmacol. 1969;9:51–72. 35. Zhou XJ, Rahmani R. Preclinical and clinical pharmacology of vinca alkaloids. Drugs. 1992;44(suppl 4):1–16 [discussion 66–69]. 36. Altaha R, Fojo T, Reed E, Abraham J. Epothilones: a novel class of non-taxane microtubule-stabilizing agents. Curr Pharm Des. 2002;8(19):1707–1712. 37. Sackett DL, Fojo T. Taxanes and other microtubule stabilizing agents. Cancer Chemother Biol Response Modif. 1999;18:59–80. 38. Loprinzi CL, Maddocks-Christianson K, Wolf SL, Rao RD, Dyck PJ, Mantyh P. The paclitaxel acute pain syndrome: sensitization of nociceptors as the putative mechanism. Cancer J. 2007;13(6):399–403. 39. Loprinzi CL, Reeves BN, Dakhil SR, et al. Natural history of paclitaxel-associated acute pain syndrome: prospective cohort study NCCTG N08C1. J Clin Oncol. 2011;29(11):1472–1478. 40. Reeves BN, Dakhil SR, Sloan JA, et al. Further data supporting that paclitaxelassociated acute pain syndrome is associated with development of peripheral neuropathy: North Central Cancer Treatment Group trial N08C1. Cancer. 2012;118(20):5171–5178. 41. Legha SS. Vincristine neurotoxicity. Pathophysiology and management. Med Toxicol. 1986;1(6):421–427. 42. Rowinsky EK, Eisenhauer EA, Chaudhry V, Arbuck SG, Donehower RC. Clinical toxicities encountered with paclitaxel (Taxol). Semin Oncol. 1993;20(4 suppl 3): 1–15. 43. Jerian SM, Sarosy GA, Link Jr CJ, Fingert HJ, Reed E, Kohn EC. Incapacitating autonomic neuropathy precipitated by taxol. Gynecol Oncol. 1993;51(2):277–280. 44. Miltenburg NC, Boogerd W. Chemotherapy-induced neuropathy: a comprehensive survey. Cancer Treat Rev. 2014;40(7):872–882. 45. Vahdat LT, Thomas ES, Roche HH, et al. Ixabepilone-associated peripheral neuropathy: data from across the phase II and III clinical trials. Support Care Cancer. 2012;20(11):2661–2668. 46. Roberts JJ, Knox RJ, Friedlos F, Lydall DA. DNA as the target for the cytotoxic and anti-tumour action of platinum coordination complexes: comparative in vitro and in vivo studies of cisplatin and carboplatin. In: McBrien DCH, Slater TF, eds. Biochemical Mechanisms of Platinum Antitumour Drugs. Oxford: IRL Press; 1986:29–64. 47. Eastman A. Activation of programmed cell death by anticancer agents: cisplatin as a model system. Cancer Cells. 1990;2(8–9):275–280.

496

Jill C. Fehrenbacher

48. Kelley MR, Jiang Y, Guo C, Reed A, Meng H, Vasko MR. Role of the DNA base excision repair protein, APE1 in cisplatin, oxaliplatin, or carboplatin induced sensory neuropathy. PLoS One. 2014;9(9):e106485. 49. Marullo R, Werner E, Degtyareva N, et al. Cisplatin induces a mitochondrial-ROS response that contributes to cytotoxicity depending on mitochondrial redox status and bioenergetic functions. PLoS One. 2013;8(11):e81162. 50. van der Hoop RG, van der Burg ME, ten Bokkel Huinink WW, van Houwelingen C, Neijt JP. Incidence of neuropathy in 395 patients with ovarian cancer treated with or without cisplatin. Cancer. 1990;66(8):1697–1702. 51. Anderson H, Wagstaff J, Crowther D, et al. Comparative toxicity of cisplatin, carboplatin (CBDCA) and iproplatin (CHIP) in combination with cyclophosphamide in patients with advanced epithelial ovarian cancer. Eur J Cancer Clin Oncol. 1988;24(9):1471–1479. 52. Gurney H, Crowther D, Anderson H, et al. Five year follow-up and dose delivery analysis of cisplatin, iproplatin or carboplatin in combination with cyclophosphamide in advanced ovarian carcinoma. Ann Oncol. 1990;1(6):427–433. 53. Argyriou AA, Cavaletti G, Briani C, et al. Clinical pattern and associations of oxaliplatin acute neurotoxicity: a prospective study in 170 patients with colorectal cancer. Cancer. 2013;119(2):438–444. 54. Wilson RH, Lehky T, Thomas RR, Quinn MG, Floeter MK, Grem JL. Acute oxaliplatin-induced peripheral nerve hyperexcitability. J Clin Oncol. 2002;20(7): 1767–1774. 55. Andre T, Boni C, Mounedji-Boudiaf L, et al. Oxaliplatin, fluorouracil, and leucovorin as adjuvant treatment for colon cancer. N Engl J Med. 2004;350(23):2343–2351. 56. Park SB, Goldstein D, Lin CS, Krishnan AV, Friedlander ML, Kiernan MC. Acute abnormalities of sensory nerve function associated with oxaliplatin-induced neurotoxicity. J Clin Oncol. 2009;27(8):1243–1249. 57. Sorokin AV, Kim ER, Ovchinnikov LP. Proteasome system of protein degradation and processing. Biochemistry (Mosc). 2009;74(13):1411–1442. 58. Testa U. Proteasome inhibitors in cancer therapy. Curr Drug Targets. 2009; 10(10):968–981. 59. Doss S, Hay N, Sutcliffe F. NICE guidance on bortezomib and thalidomide for firstline treatment of multiple myeloma. Lancet Oncol. 2011;12(9):837–838. 60. Cata JP, Weng HR, Burton AW, Villareal H, Giralt S, Dougherty PM. Quantitative sensory findings in patients with bortezomib-induced pain. J Pain. 2007;8(4):296–306. 61. Cavaletti G, Jakubowiak AJ. Peripheral neuropathy during bortezomib treatment of multiple myeloma: a review of recent studies. Leuk Lymphoma. 2010;51(7):1178–1187. 62. Stratogianni A, Tosch M, Schlemmer H, et al. Bortezomib-induced severe autonomic neuropathy. Clin Auton Res. 2012;22(4):199–202. 63. Folkman J. Tumor angiogenesis: therapeutic implications. N Engl J Med. 1971; 285(21):1182–1186. 64. Folkman J, Merler E, Abernathy C, Williams G. Isolation of a tumor factor responsible for angiogenesis. J Exp Med. 1971;133(2):275–288. 65. Leung DW, Cachianes G, Kuang WJ, Goeddel DV, Ferrara N. Vascular endothelial growth factor is a secreted angiogenic mitogen. Science. 1989;246(4935):1306–1309. 66. Grose R, Dickson C. Fibroblast growth factor signaling in tumorigenesis. Cytokine Growth Factor Rev. 2005;16(2):179–186. 67. Singhal S, Mehta J, Desikan R, et al. Antitumor activity of thalidomide in refractory multiple myeloma. N Engl J Med. 1999;341(21):1565–1571. 68. Drake MJ, Robson W, Mehta P, Schofield I, Neal DE, Leung HY. An open-label phase II study of low-dose thalidomide in androgen-independent prostate cancer. Br J Cancer. 2003;88(6):822–827.

Chemotherapy-Induced Peripheral Neuropathy

497

69. Chen L, Qiu X, Wang R, Xie X. The efficacy and safety of docetaxel plus thalidomide vs. docetaxel alone in patients with androgen-independent prostate cancer: a systematic review. Sci Rep. 2014;4:4818. 70. D’Amato RJ, Loughnan MS, Flynn E, Folkman J. Thalidomide is an inhibitor of angiogenesis. Proc Natl Acad Sci USA. 1994;91(9):4082–4085. 71. Verheul HM, Panigrahy D, Yuan J, D’Amato RJ. Combination oral antiangiogenic therapy with thalidomide and sulindac inhibits tumour growth in rabbits. Br J Cancer. 1999;79(1):114–118. 72. Alejandro LM, Behrendt CE, Chen K, Openshaw H, Shibata S. Predicting acute and persistent neuropathy associated with oxaliplatin. Am J Clin Oncol. 2013;36(4):331–337. 73. Plasmati R, Pastorelli F, Cavo M, et al. Neuropathy in multiple myeloma treated with thalidomide: a prospective study. Neurology. 2007;69(6):573–581. 74. Glasmacher A, Hahn C, Hoffmann F, et al. A systematic review of phase-II trials of thalidomide monotherapy in patients with relapsed or refractory multiple myeloma. Br J Haematol. 2006;132(5):584–593. 75. Miller AB, Hoogstraten B, Staquet M, Winkler A. Reporting results of cancer treatment. Cancer. 1981;47(1):207–214. 76. Oken MM, Creech RH, Tormey DC, et al. Toxicity and response criteria of the Eastern Cooperative Oncology Group. Am J Clin Oncol. 1982;5(6):649–655. 77. Postma TJ, Heimans JJ. Grading of chemotherapy-induced peripheral neuropathy. Ann Oncol. 2000;11(5):509–513. 78. National Cancer Institute Common Terminology Criteria for Adverse Events Criteria (NCI CTCAE) for Peripheral Neuropathy Version 4.0. ; Accessed 01.08.14. 79. Carlson K, Ocean AJ. Peripheral neuropathy with microtubule-targeting agents: occurrence and management approach. Clin Breast Cancer. 2011;11(2):73–81. 80. Cavaletti G, Bogliun G, Marzorati L, et al. Grading of chemotherapy-induced peripheral neurotoxicity using the Total Neuropathy Scale. Neurology. 2003;61(9):1297–1300. 81. Cornblath DR, Chaudhry V, Carter K, et al. Total neuropathy score: validation and reliability study. Neurology. 1999;53(8):1660–1664. 82. Cavaletti G, Jann S, Pace A, et al. Multi-center assessment of the total neuropathy score for chemotherapy-induced peripheral neurotoxicity. J Peripher Nerv Syst. 2006;11(2):135–141. 83. Velasco R, Bruna J, Briani C, et al. Early predictors of oxaliplatin-induced cumulative neuropathy in colorectal cancer patients. J Neurol Neurosurg Psychiatry. 2014;85(4):392–398. 84. Briani C, Campagnolo M, Lucchetta M, et al. Ultrasound assessment of oxaliplatininduced neuropathy and correlations with neurophysiologic findings. Eur J Neurol. 2013;20(1):188–192. 85. McHugh JC, Tryfonopoulos D, Fennelly D, Crown J, Connolly S. Electroclinical biomarkers of early peripheral neurotoxicity from oxaliplatin. Eur J Cancer Care. 2012;21(6):782–789. 86. Stubblefield MD, Vahdat LT, Balmaceda CM, Troxel AB, Hesdorffer CS, Gooch CL. Glutamine as a neuroprotective agent in high-dose paclitaxel-induced peripheral neuropathy: a clinical and electrophysiologic study. Clin Oncol (R Coll Radiol). 2005;17(4):271–276. 87. Riggs JE, Ashraf M, Snyder RD, Gutmann L. Prospective nerve conduction studies in cisplatin therapy. Ann Neurol. 1988;23(1):92–94. 88. Boogerd W, ten Bokkel Huinink WW, Dalesio O, Hoppenbrouwers WJ, van der Sande JJ. Cisplatin induced neuropathy: central, peripheral and autonomic nerve involvement. J Neurooncol. 1990;9(3):255–263.

498

Jill C. Fehrenbacher

89. Sarosy G, Kohn E, Stone DA, et al. Phase I study of taxol and granulocyte colonystimulating factor in patients with refractory ovarian cancer. J Clin Oncol. 1992;10(7):1165–1170. 90. New PZ, Jackson CE, Rinaldi D, Burris H, Barohn RJ. Peripheral neuropathy secondary to docetaxel (Taxotere). Neurology. 1996;46(1):108–111. 91. Devigili G, Tugnoli V, Penza P, et al. The diagnostic criteria for small fibre neuropathy: from symptoms to neuropathology. Brain. 2008;131(Pt 7):1912–1925. 92. Cavaletti G, Beronio A, Reni L, et al. Thalidomide sensory neurotoxicity: a clinical and neurophysiologic study. Neurology. 2004;62(12):2291–2293. 93. Lauria G, Morbin M, Lombardi R, et al. Expression of capsaicin receptor immunoreactivity in human peripheral nervous system and in painful neuropathies. J Peripher Nerv Syst. 2006;11(3):262–271. 94. Park SB, Lin CS, Kiernan MC. Nerve excitability assessment in chemotherapy-induced neurotoxicity. J Vis Exp. 2012;62:e3439. 95. Namer B, Pfeffer S, Handwerker HO, Schmelz M, Bickel A. Axon reflex flare and quantitative sudomotor axon reflex contribute in the diagnosis of small fiber neuropathy. Muscle Nerve. 2013;47(3):357–363. 96. Kalliomaki M, Kieseritzky JV, Schmidt R, et al. Structural and functional differences between neuropathy with and without pain? Exp Neurol. 2011;231(2):199–206. 97. Lauria G, McArthur JC, Hauer PE, Griffin JW, Cornblath DR. Neuropathological alterations in diabetic truncal neuropathy: evaluation by skin biopsy. J Neurol Neurosurg Psychiatry. 1998;65(5):762–766. 98. Polydefkis M, Yiannoutsos CT, Cohen BA, et al. Reduced intraepidermal nerve fiber density in HIV-associated sensory neuropathy. Neurology. 2002;58(1):115–119. 99. Nolano M, Crisci C, Santoro L, et al. Absent innervation of skin and sweat glands in congenital insensitivity to pain with anhidrosis. Clin Neurophysiol. 2000;111(9): 1596–1601. 100. Cavaletti G, Tredici G, Braga M, Tazzari S. Experimental peripheral neuropathy induced in adult rats by repeated intraperitoneal administration of taxol. Exp Neurol. 1995;133(1):64–72. 101. Cliffer KD, Siuciak JA, Carson SR, et al. Physiological characterization of Taxolinduced large-fiber sensory neuropathy in the rat. Ann Neurol. 1998;43(1):46–55. 102. Roytta M, Raine CS. Taxol-induced neuropathy: further ultrastructural studies of nerve fibre changes in situ. J Neurocytol. 1985;14(1):157–175. 103. Roytta M, Raine CS. Taxol-induced neuropathy: chronic effects of local injection. J Neurocytol. 1986;15(4):483–496. 104. Cavaletti G, Cavalletti E, Montaguti P, Oggioni N, De Negri O, Tredici G. Effect on the peripheral nervous system of the short-term intravenous administration of paclitaxel in the rat. Neurotoxicology. 1997;18(1):137–145. 105. Ling B, Authier N, Balayssac D, Eschalier A, Coudore F. Behavioral and pharmacological description of oxaliplatin-induced painful neuropathy in rat. Pain. 2007; 128(3):225–234. 106. Cata JP, Weng HR, Dougherty PM. Behavioral and electrophysiological studies in rats with cisplatin-induced chemoneuropathy. Brain Res. 2008;1230:91–98. 107. Authier N, Gillet JP, Fialip J, Eschalier A, Coudore F. Description of a shortterm Taxol-induced nociceptive neuropathy in rats. Brain Res. 2000;887(2): 239–249. 108. Contreras PC, Vaught JL, Gruner JA, et al. Insulin-like growth factor-I prevents development of a vincristine neuropathy in mice. Brain Res. 1997;774(1–2):20–26. 109. Verdu E, Vilches JJ, Rodriguez FJ, Ceballos D, Valero A, Navarro X. Physiological and immunohistochemical characterization of cisplatin-induced neuropathy in mice. Muscle Nerve. 1999;22(3):329–340.

Chemotherapy-Induced Peripheral Neuropathy

499

110. Persohn E, Canta A, Schoepfer S, et al. Morphological and morphometric analysis of paclitaxel and docetaxel-induced peripheral neuropathy in rats. Eur J Cancer. 2005;41(10):1460–1466. 111. Polomano RC, Mannes AJ, Clark US, Bennett GJ. A painful peripheral neuropathy in the rat produced by the chemotherapeutic drug, paclitaxel. Pain. 2001;94(3): 293–304. 112. Aley KO, Reichling DB, Levine JD. Vincristine hyperalgesia in the rat: a model of painful vincristine neuropathy in humans. Neuroscience. 1996;73(1):259–265. 113. Boyette-Davis J, Xin W, Zhang H, Dougherty PM. Intraepidermal nerve fiber loss corresponds to the development of Taxol-induced hyperalgesia and can be prevented by treatment with minocycline. Pain. 2011;152(2):308–313. 114. Boyette-Davis JA, Cata JP, Zhang H, et al. Follow-up psychophysical studies in bortezomib-related chemoneuropathy patients. J Pain. 2011;12(9):1017–1024. 115. Siau C, Xiao W, Bennett GJ. Paclitaxel- and vincristine-evoked painful peripheral neuropathies: loss of epidermal innervation and activation of Langerhans cells. Exp Neurol. 2006;201(2):507–514. 116. Zheng H, Xiao WH, Bennett GJ. Mitotoxicity and bortezomib-induced chronic painful peripheral neuropathy. Exp Neurol. 2012;238(2):225–234. 117. Xiao WH, Zheng H, Zheng FY, Nuydens R, Meert TF, Bennett GJ. Mitochondrial abnormality in sensory, but not motor, axons in paclitaxel-evoked painful peripheral neuropathy in the rat. Neuroscience. 2011;199:461–469. 118. Olsson Y. Topographical differences in the vascular permeability of the peripheral nervous system. Acta Neuropathol. 1968;10(1):26–33. 119. Somjen GG, Herman SP, Klein R, et al. The uptake of methyl mercury (203Hg) in different tissues related to its neurotoxic effects. J Pharmacol Exp Ther. 1973;187(3):602–611. 120. Arvidson B, Tjalve H. Distribution of 109Cd in the nervous system of rats after intravenous injection. Acta Neuropathol. 1986;69(1–2):111–116. 121. Jimenez-Andrade JM, Herrera MB, Ghilardi JR, Vardanyan M, Melemedjian OK, Mantyh PW. Vascularization of the dorsal root ganglia and peripheral nerve of the mouse: implications for chemical-induced peripheral sensory neuropathies. Mol Pain. 2008;4:10. 122. Dina OA, Chen X, Reichling D, Levine JD. Role of protein kinase Cepsilon and protein kinase A in a model of paclitaxel-induced painful peripheral neuropathy in the rat. Neuroscience. 2001;108(3):507–515. 123. Joseph EK, Chen X, Bogen O, Levine JD. Oxaliplatin acts on IB4-positive nociceptors to induce an oxidative stress-dependent acute painful peripheral neuropathy. J Pain. 2008;9(5):463–472. 124. Rigo FK, Dalmolin GD, Trevisan G, et al. Effect of omega-conotoxin MVIIA and Phalpha1beta on paclitaxel-induced acute and chronic pain. Pharmacol Biochem Behav. 2013;114–115:16–22. 125. Ferrier J, Bayet-Robert M, Pereira B, et al. A polyamine-deficient diet prevents oxaliplatin-induced acute cold and mechanical hypersensitivity in rats. PLoS One. 2013;8(10):e77828. 126. Sakurai M, Egashira N, Kawashiri T, Yano T, Ikesue H, Oishi R. Oxaliplatin-induced neuropathy in the rat: involvement of oxalate in cold hyperalgesia but not mechanical allodynia. Pain. 2009;147(1–3):165–174. 127. Cavaletti G, Gilardini A, Canta A, et al. Bortezomib-induced peripheral neurotoxicity: a neurophysiological and pathological study in the rat. Exp Neurol. 2007;204(1):317–325. 128. Joseph EK, Levine JD. Comparison of oxaliplatin- and cisplatin-induced painful peripheral neuropathy in the rat. J Pain. 2009;10(5):534–541.

500

Jill C. Fehrenbacher

129. Campana WM, Eskeland N, Calcutt NA, Misasi R, Myers RR, O’Brien JS. Prosaptide prevents paclitaxel neurotoxicity. Neurotoxicology. 1998;19(2):237–244. 130. Chiorazzi A, Nicolini G, Canta A, et al. Experimental epothilone B neurotoxicity: results of in vitro and in vivo studies. Neurobiol Dis. 2009;35(2):270–277. 131. Pisano C, Pratesi G, Laccabue D, et al. Paclitaxel and cisplatin-induced neurotoxicity: a protective role of acetyl-L-carnitine. Clin Cancer Res. 2003;9(15):5756–5767. 132. Wang MS, Davis AA, Culver DG, Wang Q, Powers JC, Glass JD. Calpain inhibition protects against Taxol-induced sensory neuropathy. Brain. 2004;127(Pt 3):671–679. 133. Xiao WH, Bennett GJ. Effects of mitochondrial poisons on the neuropathic pain produced by the chemotherapeutic agents, paclitaxel and oxaliplatin. Pain. 2012;153(3):704–709. 134. Zhang J, Tuckett RP. Comparison of paclitaxel and cisplatin effects on the slowly adapting type I mechanoreceptor. Brain Res. 2008;1214:50–57. 135. Gracias NG, Cummins TR, Kelley MR, Basile DP, Iqbal T, Vasko MR. Vasodilatation in the rat dorsal hindpaw induced by activation of sensory neurons is reduced by paclitaxel. Neurotoxicology. 2011;32(1):140–149. 136. Tanner KD, Reichling DB, Levine JD. Nociceptor hyper-responsiveness during vincristine-induced painful peripheral neuropathy in the rat. J Neurosci. 1998;18(16):6480–6491. 137. Adelsberger H, Quasthoff S, Grosskreutz J, Lepier A, Eckel F, Lersch C. The chemotherapeutic oxaliplatin alters voltage-gated Na(+) channel kinetics on rat sensory neurons. Eur J Pharmacol. 2000;406(1):25–32. 138. Flatters SJ, Bennett GJ. Studies of peripheral sensory nerves in paclitaxel-induced painful peripheral neuropathy: evidence for mitochondrial dysfunction. Pain. 2006; 122(3):245–257. 139. Authier N, Gillet JP, Fialip J, Eschalier A, Coudore F. An animal model of nociceptive peripheral neuropathy following repeated cisplatin injections. Exp Neurol. 2003;182(1):12–20. 140. Authier N, Gillet JP, Fialip J, Eschalier A, Coudore F. A new animal model of vincristine-induced nociceptive peripheral neuropathy. Neurotoxicology. 2003;24(6): 797–805. 141. Boyette-Davis J, Dougherty PM. Protection against oxaliplatin-induced mechanical hyperalgesia and intraepidermal nerve fiber loss by minocycline. Exp Neurol. 2011;229(2):353–357. 142. Melli G, Jack C, Lambrinos GL, Ringkamp M, Hoke A. Erythropoietin protects sensory axons against paclitaxel-induced distal degeneration. Neurobiol Dis. 2006;24(3):525–530. 143. Xiao WH, Zheng FY, Bennett GJ, Bordet T, Pruss RM. Olesoxime (cholest-4en-3-one, oxime): analgesic and neuroprotective effects in a rat model of painful peripheral neuropathy produced by the chemotherapeutic agent, paclitaxel. Pain. 2009;147(1–3):202–209. 144. Boehmerle W, Huehnchen P, Peruzzaro S, Balkaya M, Endres M. Electrophysiological, behavioral and histological characterization of paclitaxel, cisplatin, vincristine and bortezomib-induced neuropathy in C57Bl/6 mice. Sci Rep. 2014;4:6370. 145. Deuis JR, Lim YL, Rodrigues de Sousa S, et al. Analgesic effects of clinically used compounds in novel mouse models of polyneuropathy induced by oxaliplatin and cisplatin. Neuro Oncol. 2014;16(10):1324–1332. 146. Melli G, Taiana M, Camozzi F, et al. Alpha-lipoic acid prevents mitochondrial damage and neurotoxicity in experimental chemotherapy neuropathy. Exp Neurol. 2008;214(2):276–284. 147. Scuteri A, Nicolini G, Miloso M, et al. Paclitaxel toxicity in post-mitotic dorsal root ganglion (DRG) cells. Anticancer Res. 2006;26(2A):1065–1070.

Chemotherapy-Induced Peripheral Neuropathy

501

148. Yang IH, Siddique R, Hosmane S, Thakor N, Hoke A. Compartmentalized microfluidic culture platform to study mechanism of paclitaxel-induced axonal degeneration. Exp Neurol. 2009;218(1):124–128. 149. Anand U, Otto WR, Anand P. Sensitization of capsaicin and icilin responses in oxaliplatin treated adult rat DRG neurons. Mol Pain. 2010;6:82. 150. Silva A, Wang Q, Wang M, Ravula SK, Glass JD. Evidence for direct axonal toxicity in vincristine neuropathy. J Peripher Nerv Syst. 2006;11(3):211–216. 151. Ustinova EE, Shurin GV, Gutkin DW, Shurin MR. The role of TLR4 in the paclitaxel effects on neuronal growth in vitro. PLoS One. 2013;8(2):e56886. 152. Erturk A, Hellal F, Enes J, Bradke F. Disorganized microtubules underlie the formation of retraction bulbs and the failure of axonal regeneration. J Neurosci. 2007;27(34):9169–9180. 153. Gracias NG. Paclitaxel Alters the Function of the Small Diameter Sensory Neurons [dissertation]. Indianapolis, IN: University Graduate School, Indiana University; 2011. 154. Mo M, Erdelyi I, Szigeti-Buck K, Benbow JH, Ehrlich BE. Prevention of paclitaxel-induced peripheral neuropathy by lithium pretreatment. FASEB J. 2012; 26(11):4696–4709. 155. Kawakami K, Chiba T, Katagiri N, et al. Paclitaxel increases high voltage-dependent calcium channel current in dorsal root ganglion neurons of the rat. J Pharmacol Sci. 2012;120(3):187–195. 156. Zhang H, Dougherty PM. Enhanced excitability of primary sensory neurons and altered gene expression of neuronal ion channels in dorsal root ganglion in paclitaxel-induced peripheral neuropathy. Anesthesiology. 2014;120(6):1463–1475. 157. Pittman SK, Gracias NG, Vasko MR, Fehrenbacher JC. Paclitaxel alters the evoked release of calcitonin gene-related peptide from rat sensory neurons in culture. Exp Neurol. 2014;253:146–153. 158. Jiang Y, Guo C, Vasko MR, Kelley MR. Implications of apurinic/apyrimidinic endonuclease in reactive oxygen signaling response after cisplatin treatment of dorsal root ganglion neurons. Cancer Res. 2008;68(15):6425–6434. 159. Materazzi S, Fusi C, Benemei S, et al. TRPA1 and TRPV4 mediate paclitaxel-induced peripheral neuropathy in mice via a glutathione-sensitive mechanism. Pflugers Arch. 2012;463(4):561–569. 160. Miyano K, Tang HB, Nakamura Y, Morioka N, Inoue A, Nakata Y. Paclitaxel and vinorelbine, evoked the release of substance P from cultured rat dorsal root ganglion cells through different PKC isoform-sensitive ion channels. Neuropharmacology. 2009;57(1):25–32. 161. Wolf SL, Barton DL, Qin R, et al. The relationship between numbness, tingling, and shooting/burning pain in patients with chemotherapy-induced peripheral neuropathy (CIPN) as measured by the EORTC QLQ-CIPN20 instrument, N06CA. Support Care Cancer. 2012;20(3):625–632. 162. Nagashima M, Ooshiro M, Moriyama A, et al. Efficacy and tolerability of controlledrelease oxycodone for oxaliplatin-induced peripheral neuropathy and the extension of FOLFOX therapy in advanced colorectal cancer patients. Support Care Cancer. 2014;22(6):1579–1584. 163. Kautio AL, Haanpaa M, Saarto T, Kalso E. Amitriptyline in the treatment of chemotherapy-induced neuropathic symptoms. J Pain Symptom Manag. 2008;35(1):31–39. 164. Hammack JE, Michalak JC, Loprinzi CL, et al. Phase III evaluation of nortriptyline for alleviation of symptoms of cis-platinum-induced peripheral neuropathy. Pain. 2002;98(1–2):195–203. 165. Smith EM, Pang H, Cirrincione C, et al. Effect of duloxetine on pain, function, and quality of life among patients with chemotherapy-induced painful peripheral neuropathy: a randomized clinical trial. J Am Med Assoc. 2013;309(13):1359–1367.

502

Jill C. Fehrenbacher

166. Rao RD, Michalak JC, Sloan JA, et al. Efficacy of gabapentin in the management of chemotherapy-induced peripheral neuropathy: a phase 3 randomized, double-blind, placebo-controlled, crossover trial (N00C3). Cancer. 2007;110(9):2110–2118. 167. Rao RD, Flynn PJ, Sloan JA, et al. Efficacy of lamotrigine in the management of chemotherapy-induced peripheral neuropathy: a phase 3 randomized, double-blind, placebo-controlled trial, N01C3. Cancer. 2008;112(12):2802–2808. 168. Pascual D, Goicoechea C, Burgos E, Martin MI. Antinociceptive effect of three common analgesic drugs on peripheral neuropathy induced by paclitaxel in rats. Pharmacol Biochem Behav. 2010;95(3):331–337. 169. Sada H, Egashira N, Ushio S, Kawashiri T, Shirahama M, Oishi R. Repeated administration of amitriptyline reduces oxaliplatin-induced mechanical allodynia in rats. J Pharmacol Sci. 2012;118(4):547–551. 170. Xiao W, Naso L, Bennett GJ. Experimental studies of potential analgesics for the treatment of chemotherapy-evoked painful peripheral neuropathies. Pain Med. 2008;9(5):505–517. 171. Balayssac D, Ling B, Ferrier J, Pereira B, Eschalier A, Authier N. Assessment of thermal sensitivity in rats using the thermal place preference test: description and application in the study of oxaliplatin-induced acute thermal hypersensitivity and inflammatory pain models. Behav Pharmacol. 2014;25(2):99–111. 172. Chogtu B, Bairy KL, Smitha D, Dhar S, Himabindu P. Comparison of the efficacy of carbamazepine, gabapentin and lamotrigine for neuropathic pain in rats. Indian J Pharm. 2011;43(5):596–598. 173. Xiao W, Boroujerdi A, Bennett GJ, Luo ZD. Chemotherapy-evoked painful peripheral neuropathy: analgesic effects of gabapentin and effects on expression of the alpha-2delta type-1 calcium channel subunit. Neuroscience. 2007;144(2):714–720. 174. Hershman DL, Lacchetti C, Dworkin RH, et al. Prevention and management of chemotherapy-induced peripheral neuropathy in survivors of adult cancers: American Society of Clinical Oncology clinical practice guideline. J Clin Oncol. 2014;32(18):1941–1967. 175. Morris RL, Hollenbeck PJ. Axonal transport of mitochondria along microtubules and F-actin in living vertebrate neurons. J Cell Biol. 1995;131(5):1315–1326. 176. Shemesh OA, Spira ME. Paclitaxel induces axonal microtubules polar reconfiguration and impaired organelle transport: implications for the pathogenesis of paclitaxelinduced polyneuropathy. Acta Neuropathol. 2010;119(2):235–248. 177. Shprung T, Gozes I. A novel method for analyzing mitochondrial movement: inhibition by paclitaxel in a pheochromocytoma cell model. J Mol Neurosci. 2009;37(3):254–262. 178. Russell JW, Windebank AJ, McNiven MA, Brat DJ, Brimijoin WS. Effect of cisplatin and ACTH4-9 on neural transport in cisplatin induced neurotoxicity. Brain Res. 1995;676(2):258–267. 179. Jin HW, Flatters SJ, Xiao WH, Mulhern HL, Bennett GJ. Prevention of paclitaxelevoked painful peripheral neuropathy by acetyl-L-carnitine: effects on axonal mitochondria, sensory nerve fiber terminal arbors, and cutaneous Langerhans cells. Exp Neurol. 2008;210(1):229–237. 180. Podratz JL, Knight AM, Ta LE, et al. Cisplatin induced mitochondrial DNA damage in dorsal root ganglion neurons. Neurobiol Dis. 2011;41(3):661–668. 181. Nowikovsky K, Schweyen RJ, Bernardi P. Pathophysiology of mitochondrial volume homeostasis: potassium transport and permeability transition. Biochim Biophys Acta. 2009;1787(5):345–350. 182. Zheng H, Xiao WH, Bennett GJ. Functional deficits in peripheral nerve mitochondria in rats with paclitaxel- and oxaliplatin-evoked painful peripheral neuropathy. Exp Neurol. 2011;232(2):154–161.

Chemotherapy-Induced Peripheral Neuropathy

503

183. Kidd JF, Pilkington MF, Schell MJ, et al. Paclitaxel affects cytosolic calcium signals by opening the mitochondrial permeability transition pore. J Biol Chem. 2002;277(8):6504–6510. 184. Andre N, Braguer D, Brasseur G, et al. Paclitaxel induces release of cytochrome c from mitochondria isolated from human neuroblastoma cells’. Cancer Res. 2000; 60(19):5349–5353. 185. Evtodienko YV, Teplova VV, Sidash SS, Ichas F, Mazat JP. Microtubule-active drugs suppress the closure of the permeability transition pore in tumour mitochondria. FEBS Lett. 1996;393(1):86–88. 186. Batandier C, Leverve X, Fontaine E. Opening of the mitochondrial permeability transition pore induces reactive oxygen species production at the level of the respiratory chain complex I. J Biol Chem. 2004;279(17):17197–17204. 187. Knott AB, Bossy-Wetzel E. Impairing the mitochondrial fission and fusion balance: a new mechanism of neurodegeneration. Ann N Y Acad Sci. 2008;1147:283–292. 188. Corsetti G, Rodella L, Rezzani R, Stacchiotti A, Bianchi R. Cytoplasmic changes in satellite cells of spinal ganglia induced by cisplatin treatment in rats. Ultrastruct Pathol. 2000;24(4):259–265. 189. Voccoli V, Colombaioni L. Mitochondrial remodeling in differentiating neuroblasts. Brain Res. 2009;1252:15–29. 190. Fritz IB, Yue KT. Long-chain carnitine acyltransferase and the role of acylcarnitine derivatives in the catalytic increase of fatty acid oxidation induced by carnitine. J Lipid Res. 1963;4:279–288. 191. Flatters SJ, Xiao WH, Bennett GJ. Acetyl-L-carnitine prevents and reduces paclitaxelinduced painful peripheral neuropathy. Neurosci Lett. 2006;397(3):219–223. 192. De Grandis D, Minardi C. Acetyl-L-carnitine (levacecarnine) in the treatment of diabetic neuropathy. A long-term, randomised, double-blind, placebo-controlled study. Drugs R&D. 2002;3(4):223–231. 193. Youle M, Osio M. A double-blind, parallel-group, placebo-controlled, multicentre study of acetyl L-carnitine in the symptomatic treatment of antiretroviral toxic neuropathy in patients with HIV-1 infection. HIV Med. 2007;8(4):241–250. 194. Osio M, Muscia F, Zampini L, et al. Acetyl-L-carnitine in the treatment of painful antiretroviral toxic neuropathy in human immunodeficiency virus patients: an open label study. J Peripher Nerv Syst. 2006;11(1):72–76. 195. Bianchi G, Vitali G, Caraceni A, et al. Symptomatic and neurophysiological responses of paclitaxel- or cisplatin-induced neuropathy to oral acetyl-L-carnitine. Eur J Cancer. 2005;41(12):1746–1750. 196. Hershman DL, Unger JM, Crew KD, et al. Randomized double-blind placebocontrolled trial of acetyl-L-carnitine for the prevention of taxane-induced neuropathy in women undergoing adjuvant breast cancer therapy. J Clin Oncol. 2013;31(20):2627–2633. 197. Bordet T, Buisson B, Michaud M, et al. Identification and characterization of cholest4-en-3-one, oxime (TRO19622), a novel drug candidate for amyotrophic lateral sclerosis. J Pharmacol Exp Ther. 2007;322(2):709–720. 198. Rovini A, Carre M, Bordet T, Pruss RM, Braguer D. Olesoxime prevents microtubule-targeting drug neurotoxicity: selective preservation of EB comets in differentiated neuronal cells. Biochem Pharmacol. 2010;80(6):884–894. 199. Renehan WE, Munger BL. Degeneration and regeneration of peripheral nerve in the rat trigeminal system. II. Response to nerve lesions. J Comp Neurol. 1986;249(4): 429–459. 200. Huang TJ, Price SA, Chilton L, et al. Insulin prevents depolarization of the mitochondrial inner membrane in sensory neurons of type 1 diabetic rats in the presence of sustained hyperglycemia. Diabetes. 2003;52(8):2129–2136.

504

Jill C. Fehrenbacher

201. Doyle T, Chen Z, Muscoli C, et al. Targeting the overproduction of peroxynitrite for the prevention and reversal of paclitaxel-induced neuropathic pain. J Neurosci. 2012;32(18):6149–6160. 202. Barriere DA, Rieusset J, Chanteranne D, et al. Paclitaxel therapy potentiates cold hyperalgesia in streptozotocin-induced diabetic rats through enhanced mitochondrial reactive oxygen species production and TRPA1 sensitization. Pain. 2012;153(3): 553–561. 203. Ferrari LF, Chum A, Bogen O, Reichling DB, Levine JD. Role of Drp1, a key mitochondrial fission protein, in neuropathic pain. J Neurosci. 2011;31(31):11404–11410. 204. Ridger VC, Greenacre SA, Handy RL, et al. Effect of peroxynitrite on plasma extravasation, microvascular blood flow and nociception in the rat. Br J Pharmacol. 1997;122(6):1083–1088. 205. Hamanaka RB, Chandel NS. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem Sci. 2010;35(9):505–513. 206. Kim D, You B, Jo EK, Han SK, Simon MI, Lee SJ. NADPH oxidase 2-derived reactive oxygen species in spinal cord microglia contribute to peripheral nerve injury-induced neuropathic pain. Proc Natl Acad Sci USA. 2010;107(33):14851–14856. 207. Jang HJ, Hwang S, Cho KY, Kim do K, Chay KO, Kim JK. Taxol induces oxidative neuronal cell death by enhancing the activity of NADPH oxidase in mouse cortical cultures. Neurosci Lett. 2008;443:17–22. 208. Russell JW, Golovoy D, Vincent AM, et al. High glucose-induced oxidative stress and mitochondrial dysfunction in neurons. FASEB J. 2002;16(13):1738–1748. 209. Hamilton RT, Bhattacharya A, Walsh ME, et al. Elevated protein carbonylation, and misfolding in sciatic nerve from db/db and Sod1(/) mice: plausible link between oxidative stress and demyelination. PLoS One. 2013;8(6):e65725. 210. Janes K, Doyle T, Bryant L, et al. Bioenergetic deficits in peripheral nerve sensory axons during chemotherapy-induced neuropathic pain resulting from peroxynitritemediated post-translational nitration of mitochondrial superoxide dismutase. Pain. 2013;154(11):2432–2440. 211. Fidanboylu M, Griffiths LA, Flatters SJ. Global inhibition of reactive oxygen species (ROS) inhibits paclitaxel-induced painful peripheral neuropathy. PLoS One. 2011;6(9): e25212. 212. Kim HK, Zhang YP, Gwak YS, Abdi S. Phenyl N-tert-butylnitrone, a free radical scavenger, reduces mechanical allodynia in chemotherapy-induced neuropathic pain in rats. Anesthesiology. 2010;112(2):432–439. 213. Arthur JR. The glutathione peroxidases. Cell Mol Life Sci. 2000;57(13–14):1825–1835. 214. Moldeus P, Cotgreave IA, Berggren M. Lung protection by a thiol-containing antioxidant: N-acetylcysteine. Respiration. 1986;50(suppl 1):31–42. 215. Tredici G, Cavaletti G, Petruccioli MG, Fabbrica D, Tedeschi M, Venturino P. Lowdose glutathione administration in the prevention of cisplatin-induced peripheral neuropathy in rats. Neurotoxicology. 1994;15(3):701–704. 216. Hamers FP, Brakkee JH, Cavalletti E, et al. Reduced glutathione protects against cisplatin-induced neurotoxicity in rats. Cancer Res. 1993;53(3):544–549. 217. Cavaletti G, Minoia C, Schieppati M, Tredici G. Protective effects of glutathione on cisplatin neurotoxicity in rats. Int J Radiat Oncol Biol Phys. 1994;29(4):771–776. 218. Bogliun G, Marzorati L, Cavaletti G, Frattola L. Evaluation by somatosensory evoked potentials of the neurotoxicity of cisplatin alone or in combination with glutathione. Ital J Neurol Sci. 1992;13(8):643–647. 219. Cascinu S, Cordella L, Del Ferro E, Fronzoni M, Catalano G. Neuroprotective effect of reduced glutathione on cisplatin-based chemotherapy in advanced gastric cancer: a randomized double-blind placebo-controlled trial. J Clin Oncol. 1995;13(1):26–32.

Chemotherapy-Induced Peripheral Neuropathy

505

220. Colombo N, Bini S, Miceli D, et al. Weekly cisplatin +/ glutathione in relapsed ovarian carcinoma. Int J Gynecol Cancer. 1995;5(2):81–86. 221. Smyth JF, Bowman A, Perren T, et al. Glutathione reduces the toxicity and improves quality of life of women diagnosed with ovarian cancer treated with cisplatin: results of a double-blind, randomised trial. Ann Oncol. 1997;8(6):569–573. 222. Cascinu S, Catalano V, Cordella L, et al. Neuroprotective effect of reduced glutathione on oxaliplatin-based chemotherapy in advanced colorectal cancer: a randomized, double-blind, placebo-controlled trial. J Clin Oncol. 2002;20(16):3478–3483. 223. Milla P, Airoldi M, Weber G, Drescher A, Jaehde U, Cattel L. Administration of reduced glutathione in FOLFOX4 adjuvant treatment for colorectal cancer: effect on oxaliplatin pharmacokinetics, Pt-DNA adduct formation, and neurotoxicity. Anticancer Drugs. 2009;20(5):396–402. 224. Lin PC, Lee MY, Wang WS, et al. N-acetylcysteine has neuroprotective effects against oxaliplatin-based adjuvant chemotherapy in colon cancer patients: preliminary data. Support Care Cancer. 2006;14(5):484–487. 225. Leal AD, Qin R, Atherton PJ, 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(12):1890–1897. 226. Argyriou AA, Chroni E, Koutras A, et al. Preventing paclitaxel-induced peripheral neuropathy: a phase II trial of vitamin E supplementation. J Pain Symptom Manag. 2006;32(3):237–244. 227. Argyriou AA, Chroni E, Koutras A, et al. A randomized controlled trial evaluating the efficacy and safety of vitamin E supplementation for protection against cisplatin-induced peripheral neuropathy: final results. Support Care Cancer. 2006;14(11):1134–1140. 228. Pace A, Savarese A, Picardo M, et al. Neuroprotective effect of vitamin E supplementation in patients treated with cisplatin chemotherapy. J Clin Oncol. 2003;21(5):927–931. 229. Pace A, Giannarelli D, Galie E, et al. Vitamin E neuroprotection for cisplatin neuropathy: a randomized, placebo-controlled trial. Neurology. 2010;74(9):762–766. 230. Kottschade LA, Sloan JA, Mazurczak MA, et al. The use of vitamin E for the prevention of chemotherapy-induced peripheral neuropathy: results of a randomized phase III clinical trial. Support Care Cancer. 2011;19(11):1769–1777. 231. Alexandre J, Batteux F, Nicco C, et al. Accumulation of hydrogen peroxide is an early and crucial step for paclitaxel-induced cancer cell death both in vitro and in vivo. Int J Cancer. 2006;119(1):41–48. 232. Dzagnidze A, Katsarava Z, Makhalova J, et al. Repair capacity for platinum-DNA adducts determines the severity of cisplatin-induced peripheral neuropathy. J Neurosci. 2007;27(35):9451–9457. 233. Screnci D, McKeage MJ. Platinum neurotoxicity: clinical profiles, experimental models and neuroprotective approaches. J Inorg Biochem. 1999;77(1–2):105–110. 234. Nieto FR, Entrena JM, Cendan CM, Pozo ED, Vela JM, Baeyens JM. Tetrodotoxin inhibits the development and expression of neuropathic pain induced by paclitaxel in mice. Pain. 2008;137(3):520–531. 235. Grolleau F, Gamelin L, Boisdron-Celle M, Lapied B, Pelhate M, Gamelin E. A possible explanation for a neurotoxic effect of the anticancer agent oxaliplatin on neuronal voltage-gated sodium channels. J Neurophysiol. 2001;85(5):2293–2297. 236. Flatters SJ, Bennett GJ. Ethosuximide reverses paclitaxel- and vincristine-induced painful peripheral neuropathy. Pain. 2004;109(1–2):150–161. 237. Matsumoto M, Inoue M, Hald A, Xie W, Ueda H. Inhibition of paclitaxelinduced A-fiber hypersensitization by gabapentin. J Pharmacol Exp Ther. 2006;318(2): 735–740.

506

Jill C. Fehrenbacher

238. Okubo K, Takahashi T, Sekiguchi F, et al. Inhibition of T-type calcium channels and hydrogen sulfide-forming enzyme reverses paclitaxel-evoked neuropathic hyperalgesia in rats. Neuroscience. 2011;188:148–156. 239. Hori K, Ozaki N, Suzuki S, Sugiura Y. Upregulations of P2X(3) and ASIC3 involve in hyperalgesia induced by cisplatin administration in rats. Pain. 2010;149(2):393–405. 240. Hara T, Chiba T, Abe K, et al. Effect of paclitaxel on transient receptor potential vanilloid 1 in rat dorsal root ganglion. Pain. 2013;154(6):882–889. 241. Ta LE, Bieber AJ, Carlton SM, Loprinzi CL, Low PA, Windebank AJ. Transient Receptor Potential Vanilloid 1 is essential for cisplatin-induced heat hyperalgesia in mice. Mol Pain. 2010;6:15. 242. Chen Y, Yang C, Wang ZJ. Proteinase-activated receptor 2 sensitizes transient receptor potential vanilloid 1, transient receptor potential vanilloid 4, and transient receptor potential ankyrin 1 in paclitaxel-induced neuropathic pain. Neuroscience. 2011;193:440–451. 243. Gee NS, Brown JP, Dissanayake VU, Offord J, Thurlow R, Woodruff GN. The novel anticonvulsant drug, gabapentin (Neurontin), binds to the alpha2delta subunit of a calcium channel. J Biol Chem. 1996;271(10):5768–5776. 244. Luo ZD, Calcutt NA, Higuera ES, et al. Injury type-specific calcium channel alpha 2 delta-1 subunit up-regulation in rat neuropathic pain models correlates with antiallodynic effects of gabapentin. J Pharmacol Exp Ther. 2002;303(3):1199–1205. 245. Gauchan P, Andoh T, Ikeda K, et al. Mechanical allodynia induced by paclitaxel, oxaliplatin and vincristine: different effectiveness of gabapentin and different expression of voltage-dependent calcium channel alpha(2)delta-1 subunit. Biol Pharm Bull. 2009;32(4):732–734. 246. Brittain JM, Duarte DB, Wilson SM, et al. Suppression of inflammatory and neuropathic pain by uncoupling CRMP-2 from the presynaptic Ca(2)(+) channel complex. Nat Med. 2011;17(7):822–829. 247. Wilson SM, Brittain JM, Piekarz AD, et al. Further insights into the antinociceptive potential of a peptide disrupting the N-type calcium channel-CRMP-2 signaling complex. Channels (Austin). 2011;5(5):449–456. 248. Armstrong CM, Cota G. Calcium block of Na+ channels and its effect on closing rate. Proc Natl Acad Sci USA. 1999;96(7):4154–4157. 249. Jacobsen D, McMartin KE. Methanol and ethylene glycol poisonings. Mechanism of toxicity, clinical course, diagnosis and treatment. Med Toxicol. 1986;1(5):309–334. 250. Gamelin L, Boisdron-Celle M, Delva R, et al. Prevention of oxaliplatin-related neurotoxicity by calcium and magnesium infusions: a retrospective study of 161 patients receiving oxaliplatin combined with 5-fluorouracil and leucovorin for advanced colorectal cancer. Clin Cancer Res. 2004;10(12 Pt 1):4055–4061. 251. Hochster HS, Grothey A, Childs BH. Use of calcium and magnesium salts to reduce oxaliplatin-related neurotoxicity. J Clin Oncol. 2007;25(25):4028–4029. 252. Gamelin L, Boisdron-Celle M, Morel A, et al. Oxaliplatin-related neurotoxicity: interest of calcium–magnesium infusion and no impact on its efficacy. J Clin Oncol. 2008;26(7):1188–1189 [author reply 1189–1190]. 253. Grothey A, Hart L, Rowland KM. Intermittent oxaliplatin administration and time-to-treatment-failure in metastatic colorectal cancer: final results of the phase III CONcePT trial. J Clin Oncol. 2008;26(Suppl):4010. 254. Grothey A, Nikcevich DA, Sloan JA, et al. Intravenous calcium and magnesium for oxaliplatin-induced sensory neurotoxicity in adjuvant colon cancer: NCCTG N04C7. J Clin Oncol. 2011;29(4):421–427. 255. Loprinzi CL, Qin R, Dakhil SR, et al. Phase III randomized, placebo-controlled, doubleblind study of intravenous calcium and magnesium to prevent oxaliplatin-induced sensory neurotoxicity (N08CB/Alliance). J Clin Oncol. 2014;32(10):997–1005.

Chemotherapy-Induced Peripheral Neuropathy

507

256. Atkins CD. Estimating the value of intravenous calcium and magnesium in ameliorating oxaliplatin-induced neuropathy. J Clin Oncol. 2014;32:3341. 257. Avan A, Giovannetti E, Peters GJ. Calcium/magnesium infusion for oxaliplatininduced neuropathy: protective or not? J Clin Oncol. 2014;32(29):3341. 258. Warwick RA, Hanani M. The contribution of satellite glial cells to chemotherapyinduced neuropathic pain. Eur J Pain. 2013;17(4):571–580. 259. Pevida M, Lastra A, Hidalgo A, Baamonde A, Menendez L. Spinal CCL2 and microglial activation are involved in paclitaxel-evoked cold hyperalgesia. Brain Res Bull 2013;95:21–27. 260. Peters CM, Jimenez-Andrade JM, Jonas BM, et al. Intravenous paclitaxel administration in the rat induces a peripheral sensory neuropathy characterized by macrophage infiltration and injury to sensory neurons and their supporting cells. Exp Neurol. 2007;203(1):42–54. 261. Park HJ, Stokes JA, Corr M, Yaksh TL. Toll-like receptor signaling regulates cisplatininduced mechanical allodynia in mice. Cancer Chemother Pharmacol. 2014;73(1):25–34. 262. Li Y, Zhang H, Kosturakis AK, Jawad AB, Dougherty PM. Toll-like receptor 4 signaling contributes to paclitaxel-induced peripheral neuropathy. J Pain. 2014;15(7):712–725. 263. Zhang H, Boyette-Davis JA, Kosturakis AK, et al. Induction of monocyte chemoattractant protein-1 (MCP-1) and its receptor CCR2 in primary sensory neurons contributes to paclitaxel-induced peripheral neuropathy. J Pain. 2013;14(10): 1031–1044. 264. Levi-Montalcini R, Angeletti PU. Nerve growth factor. Physiol Rev. 1968;48(3):534–569. 265. Cavaletti G, Bogliun G, Marzorati L, et al. Early predictors of peripheral neurotoxicity in cisplatin and paclitaxel combination chemotherapy. Ann Oncol. 2004;15(9): 1439–1442. 266. Apfel SC, Lipton RB, Arezzo JC, Kessler JA. Nerve growth factor prevents toxic neuropathy in mice. Ann Neurol. 1991;29(1):87–90. 267. Peterson ER, Crain SM. Nerve growth factor attenuates neurotoxic effects of taxol on spinal cord-ganglion explants from fetal mice. Science. 1982;217(4557):377–379. 268. Woolf CJ, Ma QP, Allchorne A, Poole S. Peripheral cell types contributing to the hyperalgesic action of nerve growth factor in inflammation. J Neurosci. 1996;16(8):2716–2723. 269. Fukuoka T, Kondo E, Dai Y, Hashimoto N, Noguchi K. Brain-derived neurotrophic factor increases in the uninjured dorsal root ganglion neurons in selective spinal nerve ligation model. J Neurosci. 2001;21(13):4891–4900. 270. Holmes D. Anti-NGF painkillers back on track? Nat Rev Drug Discov. 2012;11(5): 337–338. 271. Zhu W, Galoyan SM, Petruska JC, Oxford GS, Mendell LM. A developmental switch in acute sensitization of small dorsal root ganglion (DRG) neurons to capsaicin or noxious heating by NGF. J Neurophysiol. 2004;92(5):3148–3152. 272. Claude P, Hawrot E, Dunis DA, Campenot RB. Binding, internalization, and retrograde transport of 125I-nerve growth factor in cultured rat sympathetic neurons. J Neurosci. 1982;2(4):431–442. 273. Hendry IA, Stockel K, Thoenen H, Iversen LL. The retrograde axonal transport of nerve growth factor. Brain Res. 1974;68(1):103–121. 274. Acosta CG, Fabrega AR, Masco DH, Lopez HS. A sensory neuron subpopulation with unique sequential survival dependence on nerve growth factor and basic fibroblast growth factor during development. J Neurosci. 2001;21(22):8873–8885. 275. Buchman VL, Davies AM. Different neurotrophins are expressed and act in a developmental sequence to promote the survival of embryonic sensory neurons. Development. 1993;118(3):989–1001.

508

Jill C. Fehrenbacher

276. Andoh T, Sakamoto A, Kuraishi Y. Effects of xaliproden, a 5-HT(1)A agonist, on mechanical allodynia caused by chemotherapeutic agents in mice. Eur J Pharmacol. 2013;721(1–3):231–236. 277. Arrieta O, Hernandez-Pedro N, Fernandez-Gonzalez-Aragon MC, et al. Retinoic acid reduces chemotherapy-induced neuropathy in an animal model and patients with lung cancer. Neurology. 2011;77(10):987–995. 278. ter Laak MP, Hamers FP, Kirk CJ, Gispen WH. rhGGF2 protects against cisplatininduced neuropathy in the rat. J Neurosci Res. 2000;60(2):237–244. 279. Porzner M, Muller T, Seufferlein T. SR 57746A/xaliproden, a non-peptide neurotrophic compound: prospects and constraints for the treatment of nervous system diseases. Expert Opin Invest Drugs. 2009;18(11):1765–1772. 280. Park KA, Fehrenbacher JC, Thompson EL, Duarte DB, Hingtgen CM, Vasko MR. Signaling pathways that mediate nerve growth factor-induced increase in expression and release of calcitonin gene-related peptide from sensory neurons. Neuroscience. 2010;171(3):910–923. 281. Schmutzler BS, Roy S, Hingtgen CM. Glial cell line-derived neurotrophic factor family ligands enhance capsaicin-stimulated release of calcitonin gene-related peptide from sensory neurons. Neuroscience. 2009;161(1):148–156. 282. Burnside B. The form and arrangement of microtubules: an historical, primarily morphological, review. Ann N Y Acad Sci. 1975;253:14–26. 283. Gelfand VI, Bershadsky AD. Microtubule dynamics: mechanism, regulation, and function. Annu Rev Cell Biol. 1991;7:93–116. 284. Nennesmo I, Reinholt FP. Effects of intraneural injection of taxol on retrograde axonal transport and morphology of corresponding nerve cell bodies. Virchows Archiv B Cell Pathol Incl Mol Pathol. 1988;55(4):241–246. 285. Nakata T, Yorifuji H. Morphological evidence of the inhibitory effect of taxol on the fast axonal transport. Neurosci Res. 1999;35(2):113–122. 286. Zhou J, Liu M, Aneja R, Chandra R, Joshi HC. Enhancement of paclitaxel-induced microtubule stabilization, mitotic arrest, and apoptosis by the microtubule-targeting agent EM012. Biochem Pharmacol. 2004;68(12):2435–2441. 287. Zhuang SH, Hung YE, Hung L, et al. Evidence for microtubule target engagement in tumors of patients receiving ixabepilone. Clin Cancer Res. 2007;13(24):7480–7486. 288. Hammond JW, Huang CF, Kaech S, Jacobson C, Banker G, Verhey KJ. Posttranslational modifications of tubulin and the polarized transport of kinesin-1 in neurons. Mol Biol Cell. 2010;21(4):572–583. 289. Nishio K, Arioka H, Ishida T, et al. Enhanced interaction between tubulin and microtubule-associated protein 2 via inhibition of MAP kinase and CDC2 kinase by paclitaxel. Int J Cancer. 1995;63(5):688–693. 290. Goswami C, Dreger M, Jahnel R, Bogen O, Gillen C, Hucho F. Identification and characterization of a Ca2+-sensitive interaction of the vanilloid receptor TRPV1 with tubulin. J Neurochem. 2004;91(5):1092–1103. 291. Maldonado EN, Patnaik J, Mullins MR, Lemasters JJ. Free tubulin modulates mitochondrial membrane potential in cancer cells. Cancer Res. 2010;70(24):10192–10201. 292. Staff NP, Podratz JL, Grassner L, et al. Bortezomib alters microtubule polymerization and axonal transport in rat dorsal root ganglion neurons. Neurotoxicology. 2013;39:124–131. 293. Orr GA, Verdier-Pinard P, McDaid H, Horwitz SB. Mechanisms of Taxol resistance related to microtubules. Oncogene. 2003;22(47):7280–7295. 294. Carozzi VA, Canta A, Oggioni N, et al. Neurophysiological and neuropathological characterization of new murine models of chemotherapy-induced chronic peripheral neuropathies. Exp Neurol. 2010;226(2):301–309.

Chemotherapy-induced peripheral neuropathy.

Chemotherapy-induced peripheral neuropathy (CIPN) is common in patients receiving anticancer treatment and can affect survivability and long-term qual...
630KB Sizes 0 Downloads 19 Views