Cancer Research and Molecular Mechanisms-Sci Forschen

Full Text

Research Article
Preclinical Assessment of Low Doses of Cisplatin in the Management of Acute Promyelocytic Leukemia

  Shaloam R Dasari       Venkatramreddy Velma      Clement G Yedjou      Paul B Tchounwou*   

Cellomics and Toxicogenomics Research Laboratory, NIH/NIMHD RCMI-Center for Environmental Health, College of Science, Engineering and Technology, Jackson State University, 1400 Lynch Street, Box 18750, Jackson, MS 39217, USA

*Corresponding author: Paul B. Tchounwou, Sc.D. Presidential Distinguished Professor, Associate Dean, CSET, Jackson State University, 1400 Lynch Street, P.O. Box 18540, Jackson, Mississippi, USA, Tel: (601)979-0777; Fax: (601)979-0570; E-mail: paul.b.tchounwou@jsums.edu


Article Information

Aritcle Type: Research Article

Citation: Dasari SR, Velma V, Yedjou CG, Tchounwou PB (2015) Preclinical Assessment of Low Doses of Cisplatin in the Management of Acute Promyelocytic Leukemia. Int J Cancer Res Mol Mech 1(3): doi http:// dx.doi.org/10.16966/2381-3318.113

Copyright: © Dasari SR. et al. This is an openaccess article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Publication history: 

  • Received date: 7 Sep 2015

  • Accepted date: 13 Oct 2015

  • Published date: 15 Oct 2015
  •  

    Abstract

    Cis-diamminedichloroplatinum (II) (cisplatin) is the most widely used chemotherapeutic drug for various cancers, but its effectiveness is limited by tumor cell resistance and the severe side effects it causes. Since high level of cisplatin is cytotoxic to both cancer and normal cells, the goal of the present study was to explore the effectiveness of prolonged low doses of cisplatin in the management of leukemia. To achieve our goal, human leukemia (HL-60) cells were treated with different doses (1, 2, or 3 µM) of cisplatin for 24, 48, 72 and 96 hours. Cell viability was assessed by MTS assay. Both oxidative stress damage and genotoxicity were estimated by antioxidants, lipid peroxidation, and comet assays, respectively. Data obtained from the MTS assay demonstrated that cisplatin treatment decreased the number of viable tumor cells by direct cell killing or by simply decreasing the rate of cellular proliferation in a dose- and time-dependent fashion. The results of the lipid peroxidation showed a significant increase (p<0.05) of malondialdehyde levels with increasing cisplatin doses. Results obtained from super oxide dismutase and catalase assays showed a gradual increase in antioxidant enzyme activity in cisplatin-treated cells compared to control cells. Data generated from the Comet assay demonstrated a significant dose-dependent increase in genotoxicity with respect to DNA damage as a result of cisplatin treatment. Taken together, our research demonstrated that cisplatin-induced cytotoxicity in HL-60 cells is mediated at least in part via induction of oxidative stress and oxidative damage.

    Keywords

    Cisplatin; HL-60 cells; Cell viability; Oxidative stress; Oxidative damage

    Introduction

    Cisplatin, cis-diamminedichloroplatinum (II), is a well-known chemotherapeutic drug and is one of the most widely used drugs for the treatment of various cancers [1,2]. Cisplatin is clinically proven to combat different types of cancers including sarcomas, cancers of soft tissue, bones, muscles, and blood vessels. Recent developments in the chemotherapy of such cancers have yielded better prognosis and therefore have led to these diseases becoming less life threatening [3]. From the molecular perspective of cisplatin, it represents a perfect example of how a small alteration in chemical structure can significantly affect biological activity in target cells. Nine platinum analogs are currently in clinical trials around the world [4,5].

    Cisplatin is known to bind cellular components such as DNA and proteins, and to form DNA and protein adducts as a result of cross links with DNA and protein molecules that hamper transcription and translation mechanisms [6,7]. Those DNA adducts impact cell cycle progression check points and determine whether the cell has to die or survive with the help of repair mechanisms [8]. In addition to the formation of DNA adducts, cisplatin induces oxidative stress, and modulates calcium signaling, and cell apoptosis [9,10]. It also regulates the expression of many proteins that play a major role in cell cycle regulation and signal transduction [8,11,12]. Cisplatin also mediates the mitogen activated protein kinases (MAPK) and jun amino-terminal kinase (JNK) pathways that coordinate various extra cellular signals to regulate cell growth, survival, and apoptosis [13,14].

    Additionally, cisplatin induced oxidative stress may trigger cell death besides DNA damage. Oxidative stress is one of the important processes that induce toxicity by targeting the mitochondrion membrane potential, and eventually leads to the inhibition of calcium uptake due to loss of mitochondrial protein sulfhydryl group. The degree of oxidative stress induction is dependent on the exposure time and drug dose [15]. The thiol group (– SH) containing molecules play a major role by maintaining the intracellular redox homeostasis. Reactive oxygen species are generated when thiol radicals interact with molecular oxygen under induced conditions [3]. If the cells cannot control the generation of reactive oxygen species, they may eventually damage cellular membranes and trigger intrinsic and/or extrinsic pathways of apoptosis [16]. However, in contrary to entering the apoptotic pathway in some - patients, cells develop resistance to cisplatin due to a decrease in drug intake or a passively diffusion of drug [17,18].

    Although some of the biochemical effects of cisplatin are well studied, the detailed mode of its potential therapeutic action at low doses has not yet been elucidated. The goal of present research was to assess the low dose effects and elucidate the cytotoxic mechanisms of cisplatin on HL-60 cells. Our results provide a scientific basis to identify the lowest dose of cisplatin that has a maximum impact in reducing cancer cell proliferation, thereby minimizing its side effects on normal/non-cancer cells. Hence, in this study, we have investigated the cytotoxicity, oxidative stress, lipid peroxidation and genotoxicity potentials of cisplatin at low levels of exposure to HL-60 cells.

    Methods and Materials
    Cell culture, chemicals and reagents

    Cisplatin was obtained from University of Mississippi Medical Center (Jackson, MS). Iscove’s Modified Dulbecco’s Medium (IMDM) and HL-60 cells were purchased from American Type Culture Collection (ATCC) (Manassas, VA, USA). Fetal bovine serum (FBS), and penicillin – streptomycin were purchased from Sigma (St. Louis, MO, USA). IMDM contains 4 mM L-glutamine, 4500 mg/L glucose, and 1500 mg/L sodium bicarbonate, and is supplemented with 10% (v/v) FBS, and 1% (W/V) antibiotics. The live cells were incubated at 37° C in a 5% CO2 incubator (Thermoscientific, Waltham, MA, USA).

    Cell treatment

    The experiments were conducted with three replicates for each control or treatment group. Uniform cell density of 1 × 106 cells/mL was maintained for all the treatments and controls. Cells were treated with 1, 2, or 3 µM of cisplatin or left untreated, and incubated for various time periods (24, 48, 72, or 96 h) at 37°C in a humidified 5% CO2 incubator (Thermoscientific, Waltham, MA, USA).

    Cell proliferation assay

    Cell viability was assessed as previously described [19] with CellTiter 96® AQueous One Solution Cell Proliferation Assay kit from Promega (Madison, WI, USA) with few modifications. Briefly, 100 µL aliquots of the treated or untreated cell suspension were seeded into 96-well polystyrene tissue culture plates and 20 µL of assay reagent was added to each well. After 60 min of incubation at 37°C, the absorbance was read at 490 nm with a 96 well plate reader from BMG LABTECH GmbH (Ortenberg, Germany).

    Lipid peroxidation assay

    Malondialdehyde (MDA) concentrations were quantified as previously described [20] using lipid peroxidation assay kit (Abcam, Cambridge, MA, USA) with few modifications. After the each treatment, cell suspension was collected, centrifuged at 1,230 rpm for 5 min to get the cell pellet and the cell pellet was suspended in cell lysis buffer. After centrifugation at 13,000g for 10 min, a 200 μl aliquot was assayed for MDA levels according to the lipid peroxidation assay kit protocol. The absorbance of the sample was read at 532 nm with a 96-well plate reader from BMG Labtech GmbH (Ortenberg, Germany). The concentrations of MDA were determined from the standard curve.

    Superoxide dismutase (sod) and catalase assays

    SOD and Catalase assays were carried out as previously described [21,22] with few modifications, using commercially available SOD assay kit and Catalase assay kit (Abcam, Cambridge, MA, USA), respectively. After each treatment, both the control and treated HL-60 cells were collected. Cell lysates were quantified for SOD and catalase activities according to the manufacturer’s protocol. The final absorbance for SOD was measured at 450 nm and catalase was measured at 570 nm using 96- well plate reader from BMG Labtech GmbH (Ortenberg, Germany).

    Single cell gel electrophoresis

    Cisplatin genotoxicity in treated and untreated HL-60 cells was analyzed by alkaline single cell gel electrophoresis (Comet) assay as described earlier [23] with few modifications using Comet assay kit for single cell gel electrophoresis from Trevigen (Gaithersburg, MD, USA). All the precautions were taken to avoid the UV light effect on DNA. Low melting agarose was melted in boiling water and cooled down to 37°C. For each treatment, 75 μL of the agarose and cells mixtucre (ratio of 1:10) was placed on comet slides and solidified at 4°C. Cells were lysed with lysis solution for 30 min at 4°C, followed by denaturing the DNA with alkaline solution for 40 min at 37°C. The prepared slides were subjected to electrophoresis (1 volt/cm) for 10 min in TBE (Tris borate EDTA) buffer. After the electrophoresis, cells were fixed with 70% ethanol followed by staining with SYBR green. Epifluorescent microscope (Olympus BX51 TRF, USA) was used to observe the comet slides. The data was evaluated using the DNA damage analysis software (Loats Associates Inc., USA).

    Statistical analysis

    Experiments were carried out in triplicates, and the data were presented as means ± SDs. To test for differences among and between experimental groups, one–way analysis of variance (ANOVA) and Student’s t-test were performed respectively, using SAS software available in the Biostatistics Core Laboratory available at the RCMI Center for Environmental Health at Jackson State University for testing differences. Data were considered statistically significant for p-values less than 0.05.

    Results
    Cisplatin inhibits HL60 cell proliferation

    Cell survival was measured in HL60 cells treated with 1, 2, or 3 µM cisplatin for 24, 48, 72, and 96 h incubation periods compared to the respective controls (Figure 1). The results indicate that the cell viability was decreased with increased cisplatin dose and incubation period. For 1 µM cisplatin, the percentages of cell viability were 92.0 ± 1.7%, 81.0 ± 3.8%, 59.0 ± 2.2% and 38.0 ± 1.9% for 24, 48, 72, and 96 h treatment, respectively. The recorded data for 2 µM cisplatin were 89.0 ± 2.2%, 74.0 ± 4.0%, 52.0 ± 3.9% and 41.0 ± 2.8% for 24, 48, 72, and 96 h treatment, respectively. The cell viability percentages for 3 µM cisplatin were 84.0 ± 3.1%, 66.0 ± 1.8%, 48.0 ± 3.3%, and 34.0 ± 3.6% for 24, 48, 72, and 96h treatment, respectively. Compared to the 1 µM treatment; cell survival rates decreased by about 8%, 15%, 11%, and 4% in 3 µM cisplatin, for 24, 48, 72, and 96h treatment, respectively. Overall, significant dose- and time-dependent decreases (p<0.05) were observed in cisplatin-treat cells compared to control cells.

    Cisplatin elevates SOD and catalase activities in HL60 cells

    SOD and catalase activities were estimated in untreated or treated HL60 cells with 1, 2, or 3 µM cisplatin for 24, 48, 72, and 96 h time periods. The results were presented in Figures 2 and 3 for SOD and catalase, respectively. SOD and catalase activities were increased as the cisplatin dose or time period increases. At the lowest concentration, 1 µM cisplatin, SOD activity was increased by 23%, 30%, 62%, and 82%, while catalase activity was increased by 29%, 33%, 41%, and 58% compared to their respective controls at 24, 48, 72, and 96 h of exposure, respectively. In 2 µM cisplatin-treated cells, SOD activity was increased by 41%, 47%, 67%, and 86% while catalase activity was increased by 37%, 39%, 40%, and 57% compared to their respective controls, at 24, 48, 72, and 96 h of exposure, respectively. At the highest concentration, 3 µM cisplatin, SOD activity was increased by 61%, 68%, 73%, and 110% while catalase activity was increased by 41%, 45%, 49%, and 75% compared to their respective controls at 24, 48, 72, and 96h of exposure, respectively. The increases in SOD and catalase activities in cisplatin-treated cells were significantly different from the respective controls (p<0.05).

    Figure 1: Cytotoxic effect of cisplatin on HL-60 cells. Cells were treated with 1, 2 and 3 µM of cisplatin for 24h, 48h, 72h and 96h. Cell viability was tested by cell proliferation assay using a spectrophotometer at 490 nm. Data wasnormalized with control to one. Results were expressed as means of three independent experiments ± standard deviations. p-values less than 0.05 were considered statistically significant.

    Figure 2: Effect of cisplatin on SOD activity in HL-60 cells. SOD activity was measured for the HL-60 cells challenged with 1, 2 and 3 µM of cisplatin for 24h, 48h, 72h and 96h. Post treatment, cells were lysed and SOD activity was quantified. The absorbance of each sample was measured at 450 nm and results were presented in the graph. Data was normalized with control to 1 and expressed as means of three independent experiments ± standard deviations. p-values less than 0.05 were considered statistically significant.

    Figure 3: Effect of cisplatin on catalase activity in HL-60 cells. HL-60 cells were incubated with 1, 2, and 3 µM of cisplatin for 24h, 48h, 72h, and 96h. After each treatment, catalase activity was estimated using spectrophotometer at 570 nm. Results were expressed in nmol/min/ml. Data was normalized with control to 1. Data were expressed as mean of three independent experiments ± standard deviations. p-values less than 0.05 were considered statistically significant.

    Cisplatin induces lipid peroxidation in HL60 cells

    One of the best methods for assessing oxidative stress is the estimation of MDA, a byproduct of lipid peroxidation. The MDA concentrations were estimated in HL-60 cells treated with cisplatin. The data are presented in Figure 4. As shown on this figure, the estimated MDA levels gradually increased in a dose- and time- response manner. After 96 h of exposure HL-60 cells treated with 1, 2 and 3 µM cisplatin induced 39, 48, and 64% more MDA compared to the respective controls. A similar trend was also found for any other exposure time periods. Hence, the increased MDA levels for all the tested concentrations and time periods were significantly higher than the respective controls (p<0.05).

    Figure 4: Effect of cisplatin on lipid peroxidation in HL-60 cells. HL-60 cells were treated with different doses of cisplatin or left untreated for 24h, 48h, 72h, and 96h. MDA levels were estimated in the treated and untreated cells. The amount of MDA (nmol/ml) was measured at 532 nm. Data was normalized with control to 1. The results were expressed as mean of three independent experiments ± standard deviations. p-values less than 0.05 were considered statistically significant.

    Cisplatin induces DNA damage in HL60 cells

    Treatment of HL60 cells with cisplatin at doses 1, 2 or 3 µM for a period of 24, 48, 72, and 96 hour shows dose- and time-response relationships with regard to DNA damage. Figure 5 presents the representative Comet assay images of HL-60 cells showing substantial increases in DNA damage at higher levels of cisplatin exposure. These qualitative data were analyzed quantitatively based on three replicates using the DNA damage analysis software (Loats Associates Inc., USA), The quantified DNA damage data presented in Figure6 indicated that HL-60 cells treated with 3 µM cisplatin for 96 h induced the highest level of DNA damage, whereas those treated with 1 µM cisplatin for 24 h induced lowest level of DNA damage. At 1 µM, the percentages of DNA damage were1.6 ± 0.12%, 2.8 ± 0.51%, 3.6 ± 0.47%, and 5.7 ± 0.23% for 24, 48, 72, and 96 h exposure, respectively. At 2 µM these percentages were 3.4 ± 0.19%, 3.60 ± 0.38%, 6.1 ± 0.46% and 7.1 ± 0.28% for 24, 48, 72, and 96 h exposure, respectively. At the highest dose, 3 µM of cisplatin, the percentages of DNA damage were 5.9 ± 0.09%, 6.5 ± 0.12%, 6.7 ± 0.39%, and 10.5 ± 0.16% for 24, 48, 72, and 96 h exposure, respectively. The induced DNA damage is significantly higher for all the tested concentrations compared to the respective controls (p<0.05).

    Discussion

    The present study was carried out to investigate the cytotoxicity, oxidative stress, and genotoxicity of cisplatin in HL60 cells. The main aim of this study was to research the anti-cancer potential of low doses of cisplatin, in order to minimize effect on non-cancer cells. We evaluated cell viability for cytotoxicity, anti-oxidants levels and lipid peroxidation for oxidative stress, and comet assay for genotoxicity.

    Cisplatin is a neutral inorganic chemical compound that induces cytotoxicity by characteristic inhibition of cell cycle progression as a result of DNA adducts formation and subsequent induction of apoptosis. To exhibit toxicity, cisplatin has to hydrolyze with water molecules to become an active molecule that interacts with various macro molecules in the cell [17,24].The reactive cisplatin activity is dependent on the endogenous nucleophiles including glutathione, methionine, metallothionein, and other cellular components [2,25]. The endogenous nucleophiles act as a defense mechanism to counter attack cisplatin induced toxicity at lower levels of exposure.

    Figure 5: Representative Comet assay images of HL-60 cells exposed to cisplatin 0, 1, 2, or 3 µM for 24, 48, 72, and 96 h. The treated and untreated cells were subjected to single cell gel electrophoresis, stained with SBGR, and analyzed as described in the Materials and Methods section.

    In this study, HL60 cells were treated with 1, 2, or 3 µM cisplatin for various time periods. Study results show a significant cytotoxic effect in all tested doses (Figure1). The lowest tested dose, 1 µM cisplatin, inhibited almost 60% of cell proliferation compared to the controls. The anti-proliferative or anti- cancer properties of cisplatin are in agreement with the previous reports that cisplatin inhibited cell proliferation rate in various cancers [26-28].

    The eventual inhibition of cell proliferation or cytotoxicity in cells could be the result of collective mechanisms of cisplatin-induced oxidative stress, genotoxicity and other cellular responses. Hence, we determined whether oxidative stress plays a key role in cisplatin-induced toxicity in HL60 cells. We found that the activities of antioxidant enzymes such as SOD and catalase significantly increased in cisplatin-treated cells compared to control cells. These increases in enzymatic activities were both dose- and time-dependent (Figures 2 and 3).

    As a part of defense mechanism, SOD catalyzes the dismutation of superoxide anion to hydrogen peroxide (H2 O2 ) and the H2 O2 become the substrate for the catalase. The induction of these two enzymes in the present study can be considered as an adaptive mechanism to combat the reactive oxygen species induced by cisplatin exposure. The dose-and time-dependent increases in antioxidant levels in the present study are consistent with the previous reports on cisplatin-induced reactive oxygen formation in a concentration- and time-dependent manner [15]. As a result of elevated antioxidants in response to oxidative stress, we observed a significant induction of MDA, alipid peroxidation byproduct, in HL60 cells treated with cisplatin (Figure 4). It has been previously reported that the release lipid peroxidation products increases carbonylation of proteins, induces oxidative damage of cell membranes, and those events may lead to the initiation of cell death [29-33].

    In addition to oxidative stress, cisplatin induces genotoxocity and it is probably the most effective way to inhibit cell proliferation or cell cycle progression. Reactive cisplatin has been reported to react with functional residues such as sulfhydryl groups in proteins and DNA, specifically, nucleophilic sites of purines in DNA, leading to the formation of DNAprotein and DNA-DNA inter strand or intra strand cross links [34-37]. In the present study, the reported DNA damage was dose- and timedependent (Figures 5 and 6). For 96 h treatment, DNA damage reported for 1 µM is almost 5% whereas for 3 µM is almost 10% higher than the respective controls (Figure 6). These findings are in agreement with cisplatin induced oxidative stress (Figures 2 and 3) and lipid peroxidation (Figure 4) in a dose-and time-dependent fashion. The induced DNA damage could be result of inter and intra strand cross links or culmination of both oxidative stress and DNA cross links. The current findings are consistent with previous reports that cisplatin interacts with purine bases of DNA and form DNA adducts to induce toxicity [8,38].

    Cisplatin has also been reported to induce apoptosis and changes in cell cycle progression in wild type p53 and p53 deleted hepatoma cell lines [39]. The HL60 cells lack p53 protein, while cells with wild type p53 have been reported to show more sensitivity to cisplatin compared to p53-deficient cells [40]. These studies suggest that p53 activation could sensitize the cells to cisplatin toxicity. It has been reported that cisplatin induces apoptosis through both p53 dependent and independent pathways [39,41,42]. In consistence with the results of our study, it has been pointed out that cisplatin inhibits cell proliferation irreversibly by arresting the cells in G1 phase of cell cycle [43]. However, additional studies are required to elucidate the molecular mechanisms of cisplatininduced growth inhibition, oxidative stress, cell cycle modulation, and genotoxicity in cancer cells.

    Conclusion

    In this study, we investigated cisplatin-induced cytotoxicity, oxidative stress and genotoxicity at low levels of exposure (1, 2, and 3 µM) and at various time periods (24, 48, 72 and 96h). Cisplatin significantly inhibited cell proliferation, and induced antioxidants levels and lipid peroxidation even at lowest dose of 1 µM. Also, cisplatin significantly induced DNA damage in HL60 cells at all treatment doses. Taken together, low doses of cisplatin show a significant activity against HL-60 cells, and hence, may be used in combination with arsenic trioxide (ATO) to improve treatment and reduce potential side effects in acute promyelocytic leukemia patients. However, additional research is required to study the combined effect of cisplatin and ATO in order to determine their potential for use in APL chemotherapy.

    Figure 6: DNA damage in HL-60 cells treated with 0, 1, 2, or 3 µM cisplatin for 24, 48, 72, and 96 h. Data represent the percentages of DNA damageexpressed as means of three independent experiments ± standard deviations. p-values less than 0.05 were considered statistically significant.

    Acknowledgements

    The research described in this publication was made possible in part by a grant from the National Institutes of Health (NIMHD-G12MD007581) through the RCMI-Center for Environmental Health at Jackson State University, and in part by a grant from the Mississippi INBRE (NIGMSP20GM103476).

    References
    1. Dasari S, Tchounwou PB (2014) Cisplatin in cancer therapy: molecular mechanisms of action. Eur J Pharmacol 740: 364-378. [Ref.]
    2. Siddik ZH (2003) Cisplatin: mode of cytotoxic action and molecular basis of resistance. Oncogene 22: 7265-7279. [Ref.]
    3. Desoize B, Madoulet C (2002) Particular aspects of platinum compounds used at present in cancer treatment. Crit Rev Oncol Hematol 42: 317-325. [Ref.]
    4. Goodsell DS (2006) The molecular perspective: Cisplatin. Stem Cells 24: 514-515. [Ref.]
    5. Weiss RB, Christian MC (1993) New cisplatin analogues in development. A review. Drugs 46: 360-377. [Ref.]
    6. Cepeda V, Fuertes MA, Castilla J, Alonso C, Quevedo C, et al. (2007) Biochemical mechanisms of cisplatin cytotoxicity. Anticancer Agents Med Chem 7: 3-18. [Ref.]
    7. Dobrzyńska I, Skrzydlewska E, Figaszewski ZA (2014) Effects of Novel Dinuclear Cisplatinum(II) Complexes on the Electrical Properties of Human Molt-4 Leukemia Cells. Cell Biochem Biophys. [Ref.]
    8. Alakananda B, Soumya K (2010) Cellular Responses to CisplatinInduced DNA Damage. J Nucleic Acids 2010: 201367. [Ref.]
    9. Gupta SC, Hevia D, Patchva S, Park B, Koh W, et al. (2012) Upsides and downsides of reactive oxygen species for cancer: the roles of reactive oxygen species in tumorigenesis, prevention, and therapy. Antioxid Redox Signal 16: 1295-322. [Ref.]
    10. Saad SY, Najjar TA, Alashari M (2004) Role of non-selective adenosine receptor blockade and phosphodiesterase inhibition in cisplatininduced nephrogonadal toxicity in rats. Clin Exp Pharmacol Physiol 31: 862-867. [Ref.]
    11. Hirata J, Kikuchi Y, Kita T, Imaizumi E, Tode T, et al. (1993) Modulation of sensitivity of human ovarian cancer cells to cisdiamminedichloroplatinum(II) by 12-O-tetradecanoylphorbol-13- acetate and D,L-buthionine-S,R-sulphoximine. Int J Cancer 55: 521- 527. [Ref.]
    12. Isonishi S, Andrews PA, Howell SB (1990) Increased sensitivity to cis-diamminedichloroplatinum(II) in human ovarian carcinoma cells in response to treatment with 12-O-tetradecanoylphorbol 13-acetate. J Biol Chem 265: 3623-3627. [Ref.]
    13. Cuadrado A, Lafarga V, Cheung PC, Dolado I, Llanos S, et al. (2007) A new p38 MAP kinase-regulated transcriptional coactivator that stimulates p53-dependent apoptosis. EMBO J 26: 2115-2126. [Ref.]
    14. Jones EV, Dickman MJ, Whitmarsh AJ (2007) Regulation of p73- mediated apoptosis by c-Jun N-terminal kinase. Biochem J 405: 617- 623. [Ref.]
    15. Brozovic A, Ambriović-Ristov A, Osmak M (2010) The relationship between cisplatin-induced reactive oxygen species, glutathione, and BCL-2 and resistance to cisplatin. Crit Rev Toxicol 40: 347-359. [Ref.]
    16. Ozben T (2007) Oxidative stress and apoptosis: impact on cancer therapy. J Pharm Sci 96: 2181-2196. [Ref.]
    17. Kelland LR (2000) A new resistance mechanism to cisplatin. Drug Resist Updat 3: 139-141. [Ref.]
    18. Yoshida M, Khokhar AR, Siddik ZH (1994) Biochemical pharmacology of homologous alicyclic mixed amine platinum(II) complexes in sensitive and resistant tumor cell lines. Cancer Res 13: 3468-3473. [Ref.]
    19. Bernabei PA, Santini V, Silvestro L, Dal Pozzo O, Bezzini R, et al. (1989) In vitro chemosensitivity testing of leukemic cells: development of a semiautomated colorimetric assay. Hematol Oncol 7: 243-253. [Ref.]
    20. An CH, Wang XM, Lam HC, Ifedigbo E, Washko GR, et al. (2012) TLR4 deficiency promotes autophagy during cigarette smoke-induced pulmonary emphysema. Am J Physiol Lung Cell Mol Physiol 303: L748-57. [Ref.]
    21. Choudhery MS, Badowski M, Muise A, Pierce J, Harris DT (2014) Donor age negatively impacts adipose tissue-derived mesenchymal stem cell expansion and differentiation. J Transl Med 12:8. [Ref.]
    22. Salcher S, Hagenbuchner J, Geiger K, Seiter MA, Rainer J, et al. (2014) C10ORF10/DEPP, a transcriptional target of FOXO3, regulates ROS-sensitivity in human neuroblastoma. Mol Cancer 13:224. [Ref.]
    23. Yedjou CG, Tchounwou PB (2007) In-vitro cytotoxic and genotoxic effects of arsenic trioxide on human leukemia (HL-60) cells using the MTT and alkaline single cell gel electrophoresis (Comet) assays. Mol Cell Biochem 301 : 123-130. [Ref.]
    24. el-Khateeb M, Appleton TG, Gahan LR, Charles BG, Berners-Price SJ, et al. (1999) Reactions of cisplatin hydrolytes with methionine, cysteine, and plasma ultrafiltrate studied by a combination of HPLC and NMR techniques. J Inorg Biochem 77: 13-21. [Ref.]
    25. Goto S, Yoshida K, Morikawa T, Urata Y, Suzuki K, et al. (1995) Augmentation of transport for cisplatin-glutathione adduct in cisplatinresistant cancer cells. Cancer Res 55 : 4297-4301. [Ref.]
    26. Dhar S, Kolishetti N, Lippard SJ, Farokhzad OC (2001) Targeted delivery of a cisplatin prodrug for safer and more effective prostate cancer therapy in vivo. Proc Natl Acad Sci U S A 108: 1850-1855. [Ref.]
    27. Go RS, Adjei AA (1999) Review of the comparative pharmacology and clinical activity of cisplatin and carboplatin. J Clin Oncol 17: 409-422. [Ref.]
    28. Tsimberidou AM, Braiteh F, Stewart DJ, Kurzrock R (2009) Ultimate fate of oncology drugs approved by the us food and drug administration without a randomized Trial. J Clin Oncol 27: 6243-6250. [Ref.]
    29. Chen Z, Seimiya H, Naito M, Mashima T, Kizaki A, et al. (1999) ASK1 mediates apoptotic cell death induced by genotoxic stress. Oncogene 18: 173-180. [Ref.]
    30. Dalle-Donne I, Giustarini D, Colombo R, Rossi R, Milzani A (2003) Protein carbonylation in human diseases. Trends Mol Med 9: 169-176. [Ref.]
    31. Previati M, Lanzoni I, Corbacella E, Magosso S, Guaran V, et al. (2006) Cisplatin-induced apoptosis in human promyelocytic leukemia cells. Int J Mol Med 18: 511-516. [Ref.]
    32. Song BJ, Soh Y, Bae M, Pie J, Wan J, et al. (2001) Apoptosis of PC12 cells by 4-hydroxy-2-nonenal is mediated through selective activation of the c-Jun N-terminal protein kinase pathway. Chem Biol Interact 130-132: 943-954. [Ref.]
    33. Zanke BW, Boudreau K, Rubie E, Winnett E, Tibbles LA, et al. (1996) The stress-activated protein kinase pathway mediates cell death following injury induced by cis-platinum, UV irradiation or heat. Curr Biol 6: 606-613. [Ref.]
    34. Beck DJ, Brubaker RR (1973) Effect of cis-platinum(II) diamminodichloride on wild type and deoxyribonucleic acid repair deficient mutants of Escherichia coli. J Bacteriol 116: 1247-1252. [Ref.]
    35. Eastman A (1987) The formation, isolation and characterization of DNA adducts produced by anticancer platinum complexes. Pharmacol Ther 34: 155-166. [Ref.]
    36. Fraval HN, Roberts JJ (1978) Effects of cis-platinum(II) diamminedichloride on survival and the rate of DNA synthesis in synchronously growing HeLa cells in the absence and presence of caffeine. Chem Biol Interact 23: 111-119. [Ref.]
    37. Pascoe JM, Roberts JJ (1974) Interactions between mammalian cell DNA and inorganic platinum compounds. I. DNA interstrand crosslinking and cytotoxic properties of platinum(II) compounds. Biochem Pharmacol 23: 1359-1365. [Ref.]
    38. Yousef MI, Saad AA, El-Shennawy LK (2009) Protective effect of grape seed proanthocyanidin extract against oxidative stress induced by cisplatin in rats. Food Chem Toxicol 47: 1176-1183. [Ref.]
    39. Qin LF, Ng IO (2002) Induction of apoptosis by cisplatin and its effect on cell cycle-related proteins and cell cycle changes in hepatoma cells. Cancer let 175: 27-38. [Ref.]
    40. Bressac B, Galvin KM, Liang TJ, Isselbacher KJ, Wands JR, et al. (1990) Abnormal structure and expression of p53 gene in human hepatocellular carcinoma. Proc Natl Acad Sci U S A 87: 1973-1977. [Ref.]
    41. Fajac A, Da Silva J, Ahomadegbe JC, Rateau JG, Bernaudin JF, et al. (1996) Cisplatin-induced apoptosis and p53 gene status in a cisplatinresistant human ovarian carcinoma cell line. Int J Cancer 68: 67-74. [Ref.]
    42. Burger H, Nooter K, Boersma AW, Kortland CJ, Stoter G (1998) Expression of p53, Bcl-2 and Bax in cisplatin-induced apoptosis in testicular germ cell tumour cell lines. Br J Cancer 77: 1562-1567. [Ref.]
    43. Qu K, Lin T, Wei J, Meng F, Wang Z, et al. (2013) Cisplatin induces cell cycle arrest and senescence via upregulating P53 and P21 expression in HepG2 cells. Nan Fang Yi Ke Da Xue Xue Bao 33: 1253-1259. [Ref.]

    Download Provisional pdf here