Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1959 word(s) 1959 2021-07-22 08:11:53 |
2 update references and layout + 89 word(s) 2048 2021-08-04 07:49:19 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Cauli, O. Oxidative stress, brain and chemotherapy. Encyclopedia. Available online: https://encyclopedia.pub/entry/12741 (accessed on 29 March 2024).
Cauli O. Oxidative stress, brain and chemotherapy. Encyclopedia. Available at: https://encyclopedia.pub/entry/12741. Accessed March 29, 2024.
Cauli, Omar. "Oxidative stress, brain and chemotherapy" Encyclopedia, https://encyclopedia.pub/entry/12741 (accessed March 29, 2024).
Cauli, O. (2021, August 04). Oxidative stress, brain and chemotherapy. In Encyclopedia. https://encyclopedia.pub/entry/12741
Cauli, Omar. "Oxidative stress, brain and chemotherapy." Encyclopedia. Web. 04 August, 2021.
Oxidative stress, brain and chemotherapy
Edit

Preclinical evidence shows that several chemotherapeutic drugs widely used in cancer patients such as anthracyclines, taxanes, and platinum derivatives induced oxidative stress noted in the blood and brain, which may affect both neurons and glia cells. In animal models, the oxidative stress induced by chemotherapeutic drugs is accompanied by cognitive deficits. Administration of several antioxidants decreased or prevented these effects and helped pinpoint the potential role of antioxidants as drugs that may be able to reduce both oxidative stress and cognitive dysfunction caused by chemotherapy.

taxanes anthracyclines platinum cognition biomarker acetylcysteine clinical trial

1. Introduction

The study of cancer treatments currently occupies a prominent place in research and public health policies. These focus not only on curative or palliative purposes, but also on reducing the toxicity of cancer treatments which can, in turn, increase adherence to oncology protocols and improve quality of life and survival [1][2][3]. Antineoplastic therapies entail multiple side effects and must be closely monitored to promote better assimilation of the treatment as well as to encourage patient adherence to chemotherapy regimens and mitigate reductions in their quality of life. Chemotherapy is currently one of the most important tools in the fight against cancer; many of its side effects are well known, however, others such as cognitive impairment are still being studied.

Cognitive impairment appears in up to 50–75% of people who undergo chemotherapy [4][5]. Some cancer patients report difficulties in concentration, memory, and attention both during and after the process of treating the disease, referred colloquially as ‘chemofog’ or ‘chemobrain’. Although in most cases the damage may be subtle and temporary, in a subgroup of patients, these alterations are more severe and can persist for years [6][7]. However, it should be borne in mind that even a subtle cognitive deterioration can have substantial repercussions on daily life [8][9]. Several factors such as the combination of cancer treatments, dose used, administration route, previous genetic vulnerability, and some psycho-social characteristics, among others, could give rise to these individual differences. However, chemotherapy drugs such as doxorubicin, cisplatin, 5-fluorouracil, methotrexate, and other anti-neoplastic agents trigger cognitive dysfunction in many patients [10][11][12][13][14]. These drugs are widely used in cancer chemotherapy to treat many cancers, including lymphoma, sarcoma, breast cancer, and many pediatric cancers [15][16]. About one third of women with breast cancer and half of children with cancer are treated with anthracyclines, including doxorubicin and epirubicin, and the class of taxanes, which includes paclitaxel among other drugs that are commonly used [17][18].

Oxidative stress is a dynamic and complex condition characterised by an imbalance between the generation of reactive oxygen species (ROS) and the availability and action of antioxidants [19][20][21]. The central nervous system consumes large amounts of oxygen to carry out physiological processes, leading to elevated free radical generation [22][23]. Some factors make the CNS susceptible to ROS attack, such as the deficit of antioxidant mechanisms, presence of high levels of polyunsaturated fatty acids, and selectivity of the blood-brain barrier, which reduces the diffusion of some antioxidants [23][24]. The escape of ROS from antioxidant mechanisms and their progressive accumulation trigger lipid peroxidation mechanisms as well as structural damage to proteins and DNA [25].

Oxidative stress is a factor in several neurodegenerative diseases, including Alzheimer’s disease, Parkinson’s disease, and amyotrophic lateral sclerosis [26][27][28] and is also thought to be a cornerstone of the pathophysiological mechanisms of drug-induced damage in several organs and tissues [29][30][31][32]. Indeed, several reports have shown that several pathophysiological factors lead to cognitive impairment after chemotherapy administration [33][34][35][36]. For instance, an increase in inflammatory markers in blood leading to their entry into the brain, white matter alterations, impaired neurogenesis, and cerebrovascular alterations have all have been proposed as possible factors that contribute to cognitive impairment. Moreover, both in vivo data obtained in animal models and in vitro experiments in cultured cells suggest that chemotherapy produces both an increase in oxidative stress and a decrease in antioxidative enzymes.

2. Neuronal and Glial Oxidative Stress Induced by Chemotherapeutic Drugs in Preclinical Studies

Compared to in vivo studies, relatively few in vitro studies have investigated the direct effects of anticancer drugs on neurons [37][38][39]. In primary cultures of rat neural stem cells or progenitor cells and hippocampal neurons, cisplatin and temozolomide induced mitochondrial DNA damage, impaired respiratory activity, and increased oxidative stress [38][39]. However, the presence of antioxidants in the culture medium used in cell culture experiments may be sufficient to block the effect of anticancer drugs and therefore, these systems are not suitable for examining ROS-mediated neurotoxicity because they may not reflect the actual conditions of chemotherapy-treated animal models or patients.

Thus, to mimic the in vivo conditions of chemotherapy, an elegant study showed that methotrexate, 5-fluorouracil, or cisplatin neurotoxicity only occurred when primary cell cultures obtained from the cortical and striatal neurons from rat embryos were incubated with low concentrations of antioxidant substances [37][40]. Furthermore, the co-administration of methotrexate and 5-fluorouracil through incubation, as assessed with a cell-permeable fluorogenic probe (DHR123), showed a significant increase in intra-mitochondrial ROS. Together, these data indicate that oxidative stress plays a fundamental role in the mediation of in vitro neuronal toxicity. In other work, a far-red photostable fluorogenic probe (CellROX Deep Red Reagent) was used in primary cultures of rat hippocampal neurons to show that ROS was not increased by cyclophosphamide but that exposure to doxorubicin led to a 3-fold increase in CellROX signal intensities [41]. These data suggest that neurotoxicity is drug-dependent, and the main mechanism of chemotherapy-related cognitive impairment is unlikely to be increased oxidative stress.

In addition to mitochondria, peroxisomes also generate ROS, which in turn, promote cell senescence [42][43]. These ubiquitous cytoplasmic organelles are single-membrane vesicles that are found in most eukaryotic cells [44]. Peroxisomes are oxidative organelles in which molecular oxygen acts as a co-substrate for the formation of hydrogen peroxide. Of note, the anthracycline derivative, doxorubicin, affects peroxisomal homeostasis in neurons [45]. Moreover, in an H 2O 2 environment, the level of oxidative stress was enhanced in neurons [46] and primary cultured doxorubicin-treated neurons from rat embryos displayed an increase in oxidative stress in peroxisomes [47]. Besides neurons, some chemotherapeutic drugs have been shown to promote oxidative stress in glia cells [48]. In cultures of primary rat astrocytes, oxaliplatin induced an increase in superoxide anion production up to 10-fold compared to the controls and also induced the oxidation of lipids, proteins, and DNA [48][49]. Indeed, the level of protein carbonylation was approximately doubled in oxaliplatin-treated cells compared to control samples. Furthermore, in astrocyte cultures, the basal concentration of the oxidative stress marker 8-OH-dG also increased up to 9-fold after incubation with oxaliplatin [48]. Oxidative stress is often reported to be accompanied by DNA damage. For instance in primary neurons, the anthracycline doxorubicin, promotes the formation of DNA double-strand breaks and reduced synaptic and neurite density [50].

3. Treatment to Prevent Oxidative Stress and Cognitive Dysfunction Induced In Vivo by Chemotherapeutic Drugs

Different in vitro studies in neuronal and glia cell cultures have demonstrated that the oxidative stress induced by exposure to the chemotherapeutic drugs used in cancer treatment is reduced by molecules with antioxidant properties, as summarised in Table 1 [48][49][51][52][53]. However, treatment with acetyl-carnitine, which has been shown to decrease the ROS formation induced by doxorubicin exposure to neurons in vitro [54], has not been tested in vivo in animal models. Co-administration of different antioxidant compounds in animal models reduced oxidative stress levels and improved the cognitive deficits elicited by the administration of chemotherapeutic drugs [55][51][56][57][58][59][60].

Table 1. Drugs proven to decrease both the oxidative stress and cognitive dysfunction induced by the administration of chemotherapeutic drugs in animal models.
Antioxidant Compound Drug-Induced Cognitive
Impairment and
Oxidative Stress
Behavioural Test Used to
Assess Cognitive Function
N-acetylcysteine [88]   Recognition memory task
Cisplatin Fear conditioning learning
  Object discrimination
N-acetylcysteine [54] Doxorubicin Recognition memory task
Cyclophosphamide
Gamma-glutamyl
cysteine ethyl ester [89]
Adriamycin Recognition memory task
   
Polydatin [56]   Morris water-maze task
Doxorubicin Step-down avoidance task
   
   
Caffeic acid phenethyl ester [57] Doxorubicin Passive avoidance test
    Morris water-maze task
MESNA [53] Doxorubicin Recognition memory task

MESNA, 2-mercaptoethane sulfonate sodium.

N- acetylcysteine treatment (250 mg/kg/day) prevented free radical production, ameliorated apoptotic cellular death and dendritic spine loss, and partially reversed cisplatin-induced cognitive impairments [38]. A regimen of repeated cisplatin treatment in rats led to impaired cognitive performance (contextual fear conditioning, context object discrimination, and novel object recognition tasks), but this effect was partially mitigated by concomitant N- acetylcysteine treatment [59]. Moreover, administration of gamma-glutamyl cysteine ethyl ester, a glutathione precursor (150 mg/kg) prior to adriamycin administration (20 mg/kg body weight) led to a decreased production of protein oxidation and lipid peroxidation [60]. Furthermore, N- acetylcysteine reversed the anxiety-like behaviour and recognition memory task inhibition induced by doxorubicin and cyclophosphamide in rats [56]; this effect was also accompanied by a parallel improvement in the rats’ hippocampal GSH/GSSG ratios [56].

Polydatin, a resveratrol glycoside and potent natural antioxidant [61] extracted from the root of Polygonum cuspidatum, also counteracted the effect of the anthracycline drug doxorubicin. Prior treatment with polydatin inhibited doxorubicin-induced cognitive deficits in rats, both at the neurobehavioral and hippocampal histopathological levels [57]. Administration of caffeic acid phenethyl ester, a natural polyphenolic compound that exhibits unique context-dependent antioxidant activity was also able to counteract behavioural impairment and oxidative stress in hippocampal and prefrontal cortical tissues in rats treated with doxorubicin, as measured by the reduced glutathione content and malondialdehyde concentration in these brain areas [58].

Finally, administration of the antioxidant drug 2-mercaptoethane sulfonate sodium (MESNA) improved the production of oxidative stress markers in the blood and brains of rats treated with doxorubicin (determined using protein carbonyl and protein-bound 4-hydroxy-2-nonenal as indicators of protein oxidation and lipid peroxidation, respectively) [51][52]. In parallel, MESNA administration prevented the memory deficits induced by doxorubicin in the object recognition task [51]. These latter results are very promising for future clinical trials because MESNA is already being used in oncology patients to prevent urothelial toxicity including haemorrhagic cystitis, microhaematuria, and macrohaematuria in patients treated with chemotherapeutic drugs belonging to the oxazaphosphorine family (ifosfamide and cyclophosphamide) at doses considered to be urotoxic [62][63].

4. Oxidative Stress Markers After Chemotherapy Administration in Cancer Patients

In lung cancer patients, the concentration of the oxidative DNA damage markers 8-oxoguanine (8-oxoGua) and levels of 8-oxo-2’-deoxyguanosine (8-oxodG) in urine and whole blood were higher than in controls [64]. In addition, patients with stage IV cancer had higher urinary 8-oxoGua and 8-oxodG levels than patients with stage I–III disease. These results suggest that cancer promotes oxidative stress per se [65][66][67] and so, oxidative stress markers should be measured before chemotherapy administration in order to assess the effects of chemotherapeutic drugs versus those of the cancer alone. Urinary 8-oxodG levels have been shown to increase after radiotherapy and after six cycles of chemotherapy in lung cancer. Moreover, DNA oxidation parameters were increased both after radiotherapy and chemotherapy, suggesting that a pathophysiological mechanism such as the anti-cancer effects of these drugs may underlie these effects [64].

In previously untreated cancer patients (mainly with breast or endometrial tumours), Cadeddu et al. [68] evaluated various effects of the anthracycline drug epirubicin, including oxidative stress markers [68]. The levels of ROS were determined in blood samples using the free oxygen radicals test (FORT, with 1 FORT-U corresponding to the oxidative stress elicited by 0.26 mg/L of H 2O 2) [69]. In addition, they also measured the antioxidant enzyme glutathione peroxidase in red blood cells [68]. The administration of epirubicin promoted ROS formation and reduced the expression of glutathione peroxidase, suggesting that this drug induced oxidative stress.

Cisplatin-induced toxicities mainly seem to be caused by the formation of free radicals, leading to oxidative organ damage [70][71]. In fact, the plasma concentrations of the antioxidants vitamin C, E, and ceruloplasmin decreased after the administration of cisplatin or cisplatin-containing chemotherapy regimens. This appears to be a drug-induced effect because the concentrations of these substances returned to their initial levels just before the start of the next chemotherapy cycle. In addition, the levels of the antioxidants bilirubin and albumin also gradually decreased when measured just before the start of the next chemotherapy cycle. Furthermore, the copper/ceruloplasmin ratio, a marker of pro-oxidative status, significantly increased in the first cycle of cisplatin-based regimens [70]. In analogy of in vitro and in vivo studies as well as in human studies, direct DNA damage has been reported as a result of the mechanism of the action of chemotherapeutic drugs and oxidative stress. This has only been studied in lymphocytes derived from patients treated with chemotherapy. A substantial increase in both oxidative and direct DNA damage measured in the peripheral lymphocytes assessed by the Comet assay have been reported from before to shortly after chemotherapy administration in cancer patients [72][73][74][75][76].

References

  1. Janelsins, M.C.; Kesler, S.R.; Ahles, T.A.; Morrow, G.R. Prevalence, mechanisms, and management of cancer-related cognitive impairment. Int. Rev. Psychiatry 2014, 26, 102–113.
  2. Matikas, A.; Foukakis, T.; Bergh, J. Dose intense, dose dense and tailored dose adjuvant chemotherapy for early breast cancer: An evolution of concepts. Acta Oncol. 2017, 56, 1143–1151.
  3. Balducci, L.; Phillips, D.M.; Wallace, C.; Hardy, C. Cancer chemotherapy in the elderly. Am. Fam. Physician 1987, 35, 133–143.
  4. Ahles, T.A.; Saykin, A.J.; McDonald, B.C.; Li, Y.; Furstenberg, C.T.; Hanscom, B.S.; Mulrooney, T.J.; Schwartz, G.N.; Kaufman, P.A. Longitudinal assessment of cognitive changes associated with adjuvant treatment for breast cancer: Impact of age and cognitive reserve. J. Clin. Oncol. 2010, 28, 4434–4440.
  5. Wefel, J.S.; Kesler, S.R.; Noll, K.R.; Schagen, S.B. Clinical characteristics, pathophysiology, and management of noncentral nervous system cancer-related cognitive impairment in adults. CA. Cancer J. Clin. 2015, 65, 123–138.
  6. Miao, H.; Li, J.; Hu, S.; He, X.; Partridge, S.C.; Ren, J.; Bian, Y.; Yu, Y.; Qiu, B. Long-term cognitive impairment of breast cancer patients after chemotherapy: A functional MRI study. Eur. J. Radiol. 2016, 85, 1053–1057.
  7. Christie, L.A.; Acharya, M.M.; Parihar, V.K.; Nguyen, A.; Martirosian, V.; Limoli, C.L. Impaired cognitive function and hippocampal neurogenesis following cancer chemotherapy. Clin Cancer Res. 2012, 18, 1954–1965.
  8. Fehlauer, F.; Tribius, S.; Mehnert, A.; Rades, D. Health-related quality of life in long term breast cancer survivors treated with breast conserving therapy: Impact of age at therapy. Breast Cancer Res. Treat. 2005, 92, 217–222.
  9. Walczak, P.; Janowski, M. Chemobrain as a Product of Growing Success in Chemotherapy-Focus On Glia as Both a Victim and a Cure. Neuropsychiatry 2019, 9, 2207–2216.
  10. Soussain, C.; Ricard, D.; Fike, J.R.; Mazeron, J.J.; Psimaras, D.; Delattre, J.Y. CNS complications of radiotherapy and chemotherapy. Lancet 2009, 374, 1639–1651.
  11. Winocur, G.; Berman, H.; Nguyen, M.; Binns, M.A.; Henkelman, M.; van Eede, M.; Piquette-Miller, M.; Sekeres, M.J.; Wojtowicz, J.M.; Yu, J.; et al. Neurobiological Mechanisms of Chemotherapy-induced Cognitive Impairment in a Transgenic Model of Breast Cancer. Neuroscience 2018, 369, 51–65.
  12. Joly, F.; Alibhai, S.M.H.; Galica, J.; Park, A.; Yi, Q.L.; Wagner, L.; Tannock, I.F. Impact of Androgen Deprivation Therapy on Physical and Cognitive Function, as Well as Quality of Life of Patients With Nonmetastatic Prostate Cancer. J. Urol. 2006, 176, 2443–2447.
  13. Vardy, J.; Wefel, J.S.; Ahles, T.; Tannock, I.F.; Schagen, S.B. Cancer and cancer-therapy related cognitive dysfunction: An international perspective from the Venice cognitive workshop. Ann. Oncol. 2008, 19, 623–629.
  14. Tannock, I.F.; Ahles, T.A.; Ganz, P.A.; van Dam, F.S. Cognitive impairment associated with chemotherapy for cancer: Report of a workshop. J. Clin. Oncol. 2004, 22, 2233–2239.
  15. McGowan, J.V.; Chung, R.; Maulik, A.; Piotrowska, I.; Walker, J.M.; Yellon, D.M. Anthracycline Chemotherapy and Cardiotoxicity. Cardiovasc. Drugs Ther. 2017, 31, 63–75.
  16. Tripaydonis, A.; Conyers, R.; Elliott, D.A. Pediatric Anthracycline-Induced Cardiotoxicity: Mechanisms, Pharmacogenomics, and Pluripotent Stem-Cell Modeling. Clin. Pharmacol. Ther. 2019, 105, 614–624.
  17. Anampa, J.; Makower, D.; Sparano, J.A. Progress in adjuvant chemotherapy for breast cancer: An overview. BMC Med. 2015, 13, 195.
  18. Jasra, S.; Anampa, J. Anthracycline Use for Early Stage Breast Cancer in the Modern Era: A Review. Curr. Treat. Options Oncol. 2018, 19, 30.
  19. Costantini, D. Understanding diversity in oxidative status and oxidative stress: The opportunities and challenges ahead. J. Exp. Biol. 2019, 222, jeb194688.
  20. Ji, L.L.; Yeo, D. Oxidative stress: An evolving definition. Fac. Rev. 2021, 10, 13.
  21. Jones, D.P. Redefining oxidative stress. Antioxid. Redox Signal. 2006, 8, 1865–1879.
  22. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative stress: A key modulator in neurodegenerative diseases. Molecules 2019, 24, 1583.
  23. Salim, S. Oxidative stress and the central nervous system. J. Pharmacol. Exp. Ther. 2017, 360, 201–205.
  24. Franco, R.; Navarro, G.; Martínez-Pinilla, E. Antioxidant Defense Mechanisms in Erythrocytes and in the Central Nervous System. Antioxidants 2019, 8, 46.
  25. Poprac, P.; Jomova, K.; Simunkova, M.; Kollar, V.; Rhodes, C.J.; Valko, M. Targeting Free Radicals in Oxidative Stress-Related Human Diseases. Trends Pharmacol. Sci. 2017, 38, 592–607.
  26. Radi, E.; Formichi, P.; Battisti, C.; Federico, A. Apoptosis and oxidative stress in neurodegenerative diseases. J. Alzheimer’s Dis. 2014, 42, S125–S152.
  27. Neves Carvalho, A.; Firuzi, O.; Joao Gama, M.; van Horssen, J.; Saso, L. Oxidative Stress and Antioxidants in Neurological Diseases: Is There Still Hope? Curr. Drug Targets 2016, 18, 705–718.
  28. Ortiz, G.G.; Pacheco Moisés, F.P.; Mireles-Ramírez, M.; Flores-Alvarado, L.J.; González-Usigli, H.; Sánchez-González, V.J.; Sánchez-López, A.L.; Sánchez-Romero, L.; Díaz-Barba, E.I.; Santoscoy-Gutiérrez, J.F.; et al. Oxidative Stress: Love and Hate History in Central Nervous System. In Advances in Protein Chemistry and Structural Biology; Academic Press Inc.: Cambridge, MA, USA, 2017; Volume 108, pp. 1–31.
  29. Pereira, C.V.; Nadanaciva, S.; Oliveira, P.J.; Will, Y. The contribution of oxidative stress to drug-induced organ toxicity and its detection in vitro and in vivo. Expert Opin. Drug Metab. Toxicol. 2012, 8, 219–237.
  30. Tafazoli, S.; Spehar, D.D.; O’Brien, P.J. Oxidative stress mediated idiosyncratic drug toxicity. Drug Metab Rev. 2005, 37, 311–325.
  31. Martins, M.R.; Petronilho, F.C.; Gomes, K.M.; Dal-Pizzol, F.; Streck, E.L.; Quevedo, J. Antipsychotic-induced oxidative stress in rat brain. Neurotox. Res. 2008, 13, 63–69.
  32. Kannarkat, G.; Lasher, E.E.; Schiff, D. Neurologic complications of chemotherapy agents. Curr. Opin. Neurol. 2007, 20, 719–725.
  33. Mounier, N.M.; Abdel-Maged, A.E.S.; Wahdan, S.A.; Gad, A.M.; Azab, S.S. Chemotherapy-induced cognitive impairment (CICI): An overview of etiology and pathogenesis. Life Sci. 2020, 258, 118071.
  34. Ren, X.; Boriero, D.; Chaiswing, L.; Bondada, S.; St. Clair, D.K.; Butterfield, D.A. Plausible biochemical mechanisms of chemotherapy-induced cognitive impairment (“chemobrain”), a condition that significantly impairs the quality of life of many cancer survivors. Biochim. Biophys. Acta-Mol. Basis Dis. 2019, 1865, 1088–1097.
  35. Vitali, M.; Ripamonti, C.I.; Roila, F.; Proto, C.; Signorelli, D.; Imbimbo, M.; Corrao, G.; Brissa, A.; Rosaria, G.; de Braud, F.; et al. Cognitive impairment and chemotherapy: A brief overview. Crit. Rev. Oncol. Hematol. 2017, 118, 7–14.
  36. Williams, A.L.M.; Shah, R.; Shayne, M.; Huston, A.J.; Krebs, M.; Murray, N.; Thompson, B.D.; Doyle, K.; Korotkin, J.; van Wijngaarden, E.; et al. Associations between inflammatory markers and cognitive function in breast cancer patients receiving chemotherapy. J. Neuroimmunol. 2018, 314, 17–23.
  37. Jiang, Z.G.; Fuller, S.A.; Ghanbari, H.A. PAN-811 Blocks Chemotherapy Drug-Induced in Vitro Neurotoxicity, while Not Affecting Suppression of Cancer Cell Growth. Oxid. Med. Cell. Longev. 2016, 2016, 569807.
  38. Lomeli, N.; Di, K.; Pearre, D.C.; Chung, T.F.; Bota, D.A. Mitochondrial-associated impairments of temozolomide on neural stem/progenitor cells and hippocampal neurons. Mitochondrion 2020, 52, 56–66.
  39. Lomeli, N.; Lepe, J.; Gupta, K.; Bota, D.A. Cognitive complications of cancer and cancer-related treatments—Novel paradigms. Neurosci Lett. 2021, 749, 135720.
  40. Qian, X.; Li, J.; Xu, S.; Wan, Y.; Li, Y.; Jiang, Y.; Zhao, H.; Zhou, Y.; Liao, J.; Liu, H.; et al. Prenatal exposure to phthalates and neurocognitive development in children at two years of age. Environ. Int. 2019, 131, 105023.
  41. Bagnall-Moreau, C.; Chaudhry, S.; Salas-Ramirez, K.; Ahles, T.; Hubbard, K. Chemotherapy-Induced Cognitive Impairment Is Associated with Increased Inflammation and Oxidative Damage in the Hippocampus. Mol. Neurobiol. 2019, 56, 7159–7172.
  42. Vallée, A.; Lecarpentier, Y. Crosstalk between peroxisome proliferator-activated receptor gamma and the canonical WNT/β-catenin pathway in chronic inflammation and oxidative stress during carcinogenesis. Front. Immunol. 2018, 9, 745.
  43. Ganguli, G.; Mukherjee, U.; Sonawane, A. Peroxisomes and oxidative stress: Their implications in the modulation of cellular immunity during mycobacterial infection. Front. Microbiol. 2019, 10, 1121.
  44. Wilkinson, C.F.; Lamb IV, J.C. The potential health effects of phthalate esters in children’s toys: A review and risk assessment. Regul. Toxicol. Pharmacol. 1999, 30, 140–155.
  45. Moruno-Manchon, J.F.; Uzor, N.E.; Kesler, S.R.; Wefel, J.S.; Townley, D.M.; Nagaraja, A.S.; Pradeep, S.; Mangala, L.S.; Sood, A.K.; Tsvetkov, A.S. TFEB ameliorates the impairment of the autophagy-lysosome pathway in neurons induced by doxorubicin. Aging 2016, 8, 3507–3519.
  46. Walker, C.L.; Pomatto, L.C.D.; Tripathi, D.N.; Davies, K.J.A. Redox regulation of homeostasis and proteostasis in peroxisomes. Physiol. Rev. 2018, 98, 89–115.
  47. Moruno-Manchon, J.F.; Uzor, N.E.; Kesler, S.R.; Wefel, J.S.; Townley, D.M.; Nagaraja, A.S.; Pradeep, S.; Mangala, L.S.; Sood, A.K.; Tsvetkov, A.S. Peroxisomes contribute to oxidative stress in neurons during doxorubicin-based chemotherapy. Mol. Cell. Neurosci. 2018, 86, 65–71.
  48. Di Cesare Mannelli, L.; Zanardelli, M.; Failli, P.; Ghelardini, C. Oxaliplatin-induced oxidative stress in nervous system-derived cellular models: Could it correlate with in vivo neuropathy? Free Radic. Biol. Med. 2013, 61, 143–150.
  49. Di Cesare Mannelli, L.; Zanardelli, M.; Landini, I.; Pacini, A.; Ghelardini, C.; Mini, E.; Bencini, A.; Valtancoli, B.; Failli, P. Effect of the SOD mimetic MnL4 on in vitro and in vivo oxaliplatin toxicity: Possible aid in chemotherapy induced neuropathy. Free Radic. Biol. Med. 2016, 93, 67–76.
  50. Manchon, J.F.M.; Dabaghian, Y.; Uzor, N.E.; Kesler, S.R.; Wefel, J.S.; Tsvetkov, A.S. Levetiracetam mitigates doxorubicin-induced DNA and synaptic damage in neurons. Sci. Rep. 2016, 6, 25705.
  51. Keeney, J.T.R.; Ren, X.; Warrier, G.; Noel, T.; Powell, D.K.; Brelsfoard, J.M.; Sultana, R.; Saatman, K.E.; St. Clair, D.K.; Butterfield, D.A. Doxorubicin-induced elevated oxidative stress and neurochemical alterations in brain and cognitive decline: Protection by MESNA and insights into mechanisms of chemotherapy-induced cognitive impairment (“chemobrain”). Oncotarget 2018, 9, 30324–30339.
  52. Aluise, C.D.; St Clair, D.; Vore, M.; Butterfield, D.A. In vivo amelioration of adriamycin induced oxidative stress in plasma by gamma-glutamylcysteine ethyl ester (GCEE). Cancer Lett. 2009, 282, 25–29.
  53. Keeney, J.T.R.; Miriyala, S.; Noel, T.; Moscow, J.A.; St. Clair, D.K.; Butterfield, D.A. Superoxide induces protein oxidation in plasma and TNF-α elevation in macrophage culture: Insights into mechanisms of neurotoxicity following doxorubicin chemotherapy. Cancer Lett. 2015, 367, 157–161.
  54. Lee, S.; Rauch, J.; Kolch, W. Targeting MAPK Signaling in Cancer: Mechanisms of Drug Resistance and Sensitivity. Int. J. Mol. Sci. 2020, 21, 1102.
  55. Joshi, G.; Aluise, C.D.; Cole, M.P.; Sultana, R.; Pierce, W.M.; Vore, M.; St Clair, D.K.; Butterfield, D.A. Alterations in brain antioxidant enzymes and redox proteomic identification of oxidized brain proteins induced by the anti-cancer drug adriamycin: Implications for oxidative stress-mediated chemobrain. Neuroscience 2010, 166, 796–807.
  56. Kitamura, Y.; Ushio, S.; Sumiyoshi, Y.; Wada, Y.; Miyazaki, I.; Asanuma, M.; Sendo, T. N-acetylcysteine attenuates the anxiety-like behavior and spatial cognition impairment induced by doxorubicin and cyclophosphamide combination treatment in rats. Pharmacology 2020, 106, 286–293.
  57. Tong, Y.; Wang, K.; Sheng, S.; Cui, J. Polydatin ameliorates chemotherapy-induced cognitive impairment (chemobrain) by inhibiting oxidative stress, inflammatory response, and apoptosis in rats. Biosci. Biotechnol. Biochem. 2020, 84, 1201–1210.
  58. Ali, M.A.; Menze, E.T.; Tadros, M.G.; Tolba, M.F. Caffeic acid phenethyl ester counteracts doxorubicin-induced chemobrain in Sprague-Dawley rats: Emphasis on the modulation of oxidative stress and neuroinflammation. Neuropharmacology 2020, 181, 108334.
  59. Lomeli, N.; Di, K.; Czerniawski, J.; Guzowski, J.F.; Bota, D.A. Cisplatin-induced mitochondrial dysfunction is associated with impaired cognitive function in rats. Free Radic. Biol. Med. 2017, 102, 274–286.
  60. Joshi, G.; Hardas, S.; Sultana, R.; St. Clair, D.K.; Vore, M.; Butterfield, D.A. Glutathione elevation by γ-glutamyl cysteine ethyl ester as a potential therapeutic strategy for preventing oxidative stress in brain mediated by in vivo administration of adriamycin: Implication for chemobrain. J. Neurosci. Res. 2007, 85, 497–503.
  61. Du, Q.H.; Peng, C.; Zhang, H. Polydatin: A review of pharmacology and pharmacokinetics. Pharm. Biol. 2013, 51, 1347–1354.
  62. Shaw, I.C.; Graham, M.I. Mesna-a short review. Cancer Treat. Rev. 1987, 14, 67–86.
  63. Shaw, I.C. Mesna and oxazaphosphorine cancer chemotherapy. Cancer Treat. Rev. 1987, 14, 359–364.
  64. Crohns, M.; Saarelainen, S.; Erhola, M.; Alho, H.; Kellokumpu-Lehtinen, P. Impact of radiotherapy and chemotherapy on biomarkers of oxidative DNA damage in lung cancer patients. Clin. Biochem. 2009, 42, 1082–1090.
  65. Klaunig, J.E. Oxidative Stress and Cancer. Curr. Pharm. Des. 2019, 24, 4771–4778.
  66. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative stress, inflammation, and cancer: How are they linked? Free Radic. Biol. Med. 2010, 49, 1603–1616.
  67. Gill, J.G.; Piskounova, E.; Morrison, S.J. Cancer, oxidative stress, and metastasis. Cold Spring Harb. Symp. Quant. Biol. 2016, 81, 163–175.
  68. Cadeddu, C.; Piras, A.; Mantovani, G.; Deidda, M.; Dessì, M.; Madeddu, C.; Massa, E.; Mercuro, G. Protective effects of the angiotensin II receptor blocker telmisartan on epirubicin-induced inflammation, oxidative stress, and early ventricular impairment. Am. Heart J. 2010, 160, 487.e1–487.e7.
  69. Pavlatou, M.G.; Papastamataki, M.; Apostolakou, F.; Papassotiriou, I.; Tentolouris, N. FORT and FORD: Two simple and rapid assays in the evaluation of oxidative stress in patients with type 2 diabetes mellitus. Metabolism 2009, 58, 1657–1662.
  70. Weijl, N.I.; Elsendoorn, T.J.; Lentjes, E.G.W.M.; Hopman, G.D.; Wipkink-Bakker, A.; Zwinderman, A.H.; Cleton, F.J.; Osanto, S. Supplementation with antioxidant micronutrients and chemotherapy-induced toxicity in cancer patients treated with cisplatin-based chemotherapy: A randomised, double-blind, placebo-controlled study. Eur. J. Cancer 2004, 40, 1713–1723.
  71. Weijl, N.I.; Hopman, G.D.; Wipkink-Bakker, A.; Lentjes, E.G.W.M.; Berger, H.M.; Cleton, F.J.; Osanto, S. Cisplatin combination chemotherapy induces a fall in plasma antioxidants of cancer patients. Ann. Oncol. 1998, 9, 1331–1337.
  72. Conroy, S.K.; McDonald, B.C.; Smith, D.J.; Moser, L.R.; West, J.D.; Kamendulis, L.M.; Klaunig, J.E.; Champion, V.L.; Unverzagt, F.W.; Saykin, A.J. Alterations in brain structure and function in breast cancer survivors: Effect of post-chemotherapy interval and relation to oxidative DNA damage. Breast Cancer Res. Treat. 2013, 137, 493–502.
  73. Tomasello, B.; Malfa, G.A.; Strazzanti, A.; Gangi, S.; Di Giacomo, C.; Basile, F.; Renis, M. Effects of physical activity on systemic oxidative/DNA status in breast cancer survivors. Oncol. Lett. 2017, 13, 441–448.
  74. Scuric, Z.; Carroll, J.E.; Bower, J.E.; Ramos-Perlberg, S.; Petersen, L.; Esquivel, S.; Hogan, M.; Chapman, A.M.; Irwin, M.R.; Breen, E.C.; et al. Biomarkers of aging associated with past treatments in breast cancer survivors. NPJ Breast Cancer 2017, 3, 50.
  75. Cruz-Jentoft, A.J.; Bahat, G.; Bauer, J.; Boirie, Y.; Bruyère, O.; Cederholm, T.; Cooper, C.; Landi, F.; Rolland, Y.; Sayer, A.A.; et al. Sarcopenia: Revised European consensus on definition and diagnosis. Age Ageing 2019, 48, 16–31.
  76. Root, J.C.; Pergolizzi, D.; Pan, H.; Orlow, I.; Passik, S.D.; Silbersweig, D.; Stern, E.; Ahles, T.A. Prospective evaluation of functional brain activity and oxidative damage in breast cancer: Changes in task-induced deactivation during a working memory task. Brain Imaging Behav. 2020, 1–10.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 411
Revisions: 2 times (View History)
Update Date: 29 Mar 2022
1000/1000