Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 45
Filter
1.
J Med Chem ; 63(6): 2877-2893, 2020 03 26.
Article in English | MEDLINE | ID: mdl-32084316

ABSTRACT

As an FDA-approved drug, teniposide, was utilized in cancer treatment but was accompanied by a strong side effect in long-term clinical trials. This work discovered potential candidate drugs with low toxicity by modifying the molecule structure of teniposide through a structure-guided drug design approach. The IC50 value of novel 4,6-O-thenylidene-ß-d-glucopyranoside-(2″-acetamido, 3″-acetyl-di-S-5-fluorobenzothizole/5-fluorobenzoxazole)-4'-demethylepipodophyllotoxin (compounds 15 and 16) was 120.4-125.1 µM, which was significantly improved by around 10 times more than teniposide (11.5-22.3 µM) against healthy human cells (i.e., HL-7702, H8, MRC-5, and HMEC). In vivo studies demonstrated compounds 15 and 16 significantly suppressed the tumor growth in the HepG2 cell xenograft model without exhibiting obvious toxicity (LD50 values of 208.45 and 167.52 mg/kg), which was lower than that of teniposide (LD50 = 46.12 mg/kg). Compounds 15 and 16 caused mild γH2AX phosphorylation for low DNA toxicity and less inhibition of PI3K/Akt. Compounds 15 and 16 might be potential antitumor drugs with low toxicity.


Subject(s)
Antineoplastic Agents/chemistry , Antineoplastic Agents/pharmacology , Podophyllotoxin/analogs & derivatives , Teniposide/analogs & derivatives , Teniposide/pharmacology , Animals , Antineoplastic Agents/toxicity , Cell Line , DNA Damage/drug effects , Hep G2 Cells , Humans , Male , Mice, Inbred C57BL , Molecular Docking Simulation , Neoplasms/drug therapy , Neoplasms/genetics , Neoplasms/metabolism , Phosphatidylinositol 3-Kinases/metabolism , Phosphoinositide-3 Kinase Inhibitors/chemistry , Phosphoinositide-3 Kinase Inhibitors/pharmacology , Phosphoinositide-3 Kinase Inhibitors/toxicity , Podophyllotoxin/chemistry , Podophyllotoxin/pharmacology , Podophyllotoxin/toxicity , Teniposide/toxicity
2.
BMC Res Notes ; 12(1): 446, 2019 Jul 22.
Article in English | MEDLINE | ID: mdl-31331386

ABSTRACT

OBJECTIVE: We explored detrimental effects of VCD on non-ovarian tissues such as kidneys and liver 14 days post-drug administration. Twelve rats were randomly assigned into two groups. In VCD group, rats received 160 mg/kgbw VCD intraperitoneally for 15 consequent days. Control rats were injected with VCD-free normal saline. At the respective time point, rats were euthanized, blood and tissue samples were collected. H&E staining was performed to evaluate pathological changes. Serum level of ALT, AST, creatinine and urea were also measured. RESULTS: Histological analysis revealed hyperemia and follicular atresia in the ovaries, indicating successful POF induction in rats. In renal tissue, extensive tubular necrosis, focal hemorrhage, hyaline casts, and interstitial nephritis were observed. Analysis of hepatic tissue showed numerous hemorrhagic foci, chronic cholangitis, and hepatocyte necrosis, indicating apparent VCD toxicity of both hepatic and renal tissues. The biochemical evaluation revealed a tendency of increase in ALT, AST, creatinine, and Urea in VCD-treated rats; however, the values did not reach significant level. In conclusion, the induction of POF in rats by VCD correlates with renal and hepatic damages. Commensurate with data from this study, any conclusions from experiments based on VCD-induced premature ovarian failure rats should be reported with caution.


Subject(s)
Antineoplastic Combined Chemotherapy Protocols/pharmacology , Kidney/drug effects , Liver/drug effects , Ovary/drug effects , Alanine Transaminase/blood , Animals , Antineoplastic Combined Chemotherapy Protocols/toxicity , Aspartate Aminotransferases/blood , Creatinine/blood , Cyclophosphamide/administration & dosage , Cyclophosphamide/toxicity , Dexamethasone/administration & dosage , Dexamethasone/toxicity , Female , Injections, Intraperitoneal , Kidney/pathology , Liver/pathology , Ovary/pathology , Primary Ovarian Insufficiency/chemically induced , Primary Ovarian Insufficiency/diagnosis , Random Allocation , Rats, Wistar , Teniposide/administration & dosage , Teniposide/toxicity
3.
Toxicol In Vitro ; 37: 88-96, 2016 Dec.
Article in English | MEDLINE | ID: mdl-27622579

ABSTRACT

Chemotherapy-induced neurotoxicity can reduce the quality of life of patients by affecting their intelligence, senses and mobility. Ten percent of safety-related late-stage clinical failures are due to neurological side effects. Animal models are poor in predicting human neurotoxicity due to interspecies differences and most in vitro assays cannot distinguish neurotoxicity from general cytotoxicity for chemotherapeutics. We developed in vitro assays capable of quantifying the paediatric neurotoxic potential for cytotoxic drugs. Mixed cultures of human fetal brain cells were differentiated in monolayers and as 3D-neurospheres in the presence of non-neurotoxic chemotherapeutics (etoposide, teniposide) or neurotoxicants (methylmercury). The cytotoxic potency towards dividing progenitors versus differentiated neurons and astrocytes was compared using: (1) immunohistochemistry staining and cell counts in monolayers; (2) through quantitative Western blots in neurospheres; and (3) neurosphere migration assays. Etoposide and teniposide, were 5-10 times less toxic to differentiated neurons compared to the mix of all cells in monolayer cultures. In contrast, the neurotoxicant methylmercury did not exhibit selectivity and killed all cells with the same potency. In 3D neurospheres, etoposide and teniposide were 24 to 10 times less active against neurons compared to all cells. These assays can be used prioritise drugs for local drug delivery to brain tumours.


Subject(s)
Antineoplastic Agents/toxicity , Astrocytes/drug effects , Etoposide/toxicity , Methylmercury Compounds/toxicity , Neurons/drug effects , Teniposide/toxicity , Astrocytes/metabolism , Brain/cytology , Cell Count , Cell Culture Techniques , Cell Survival/drug effects , Cells, Cultured , Fetus , Glial Fibrillary Acidic Protein/metabolism , Humans , Neurons/metabolism , Neurotoxicity Syndromes , Oxazines/metabolism , Oxidation-Reduction , Spheroids, Cellular/drug effects , Spheroids, Cellular/metabolism , Tubulin/metabolism , Xanthenes/metabolism
4.
Mutagenesis ; 26(4): 533-43, 2011 Jul.
Article in English | MEDLINE | ID: mdl-21430063

ABSTRACT

The intention of the present study was to answer the question whether the catalytic topoisomerase-II inhibitor, dexrazoxane, can be used as a modulator of teniposide-induced DNA damage and programmed cell death (apoptosis) in the bone marrow cells in vivo. The alkaline single cell gel electrophoresis, scoring of chromosomal aberrations, micronuclei and mitotic activity were undertaken in the current study as markers of DNA damage. Apoptosis was analysed by the occurrence of a hypodiploid DNA peak and caspase-3 activity. Oxidative stress marker such as intracellular reactive oxygen species production, lipid peroxidation, reduced and oxidised glutathione were assessed in bone marrow as a possible mechanism underlying this amelioration. Dexrazoxane was neither genotoxic nor apoptogenic in mice at the tested dose. Moreover, for the first time, it has been shown that dexrazoxane affords significant protection against teniposide-induced DNA damage and apoptosis in the bone marrow cells in vivo and effectively suppresses the apoptotic signalling triggered by teniposide. Teniposide induced marked biochemical alterations characteristic of oxidative stress including accumulation of intracellular reactive oxygen species, enhanced lipid peroxidation, accumulation of oxidised glutathione and reduction in the reduced glutathione level. Prior administration of dexrazoxane ahead of teniposide challenge ameliorated these biochemical alterations. It is thus concluded that pretreatment with dexrazoxane attenuates teniposide-induced oxidative stress and subsequent DNA damage and apoptosis in bone marrow cells. Based on our data presented, strategies can be developed to decrease the teniposide-induced DNA damage in normal cells using dexrazoxane. Therefore, dexrazoxane can be a good candidate to decrease the deleterious effects of teniposide in the bone marrow cells of cancer patients treated with teniposide.


Subject(s)
Apoptosis/drug effects , Bone Marrow Cells/cytology , Bone Marrow Cells/drug effects , DNA Damage , Razoxane/pharmacology , Teniposide/toxicity , Animals , Bone Marrow Cells/enzymology , Caspase 3/metabolism , Cell Cycle/drug effects , Chromosomes, Mammalian/metabolism , DNA Breaks/drug effects , Flow Cytometry , Glutathione/metabolism , Lipid Peroxidation/drug effects , Male , Mice , Micronuclei, Chromosome-Defective/drug effects , Reactive Oxygen Species/metabolism
5.
Mol Pharmacol ; 79(4): 735-41, 2011 Apr.
Article in English | MEDLINE | ID: mdl-21252292

ABSTRACT

Nuclear factor (NF)-YB, a subunit of the transcription factor nuclear factor Y (NF-Y) complex, binds and activates CCAAT-containing promoters. Our previous work suggested that NF-YB may be a mediator of topoisomerase IIα (Top2α), working through the Top2α promoter. DNA topoisomerase II (Top2) is an essential nuclear enzyme and the primary target for several clinically important anticancer drugs. Our teniposide-resistant human lymphoblastic leukemia CEM cells (CEM/VM-1-5) express reduced Top2α protein compared with parental CEM cells. To study the regulation of Top2α during the development of drug resistance, we found that NF-YB protein expression is increased in CEM/VM-1-5 cells compared with parental CEM cells. This further suggests that increased NF-YB may be a negative regulator of Top2α in CEM/VM-1-5 cells. We asked what causes the up-regulation of NF-YB in CEM/VM-1-5 cells. We found by microRNA profiling that hsa-miR-485-3p is lower in CEM/VM-1-5 cells compared with CEM cells. MicroRNA target prediction programs revealed that the 3'-untranslated region (3'-UTR) of NF-YB harbors a putative hsa-miR-485-3p binding site. We thus hypothesized that hsa-miR-485-3p mediates drug responsiveness by decreasing NF-YB expression, which in turn negatively regulates Top2α expression. To test this, we overexpressed miR-485-3p in CEM/VM-1-5 cells and found that this led to reduced expression of NF-YB, a corresponding up-regulation of Top2α, and increased sensitivity to the Top2 inhibitors. Results in CEM cells were replicated in drug-sensitive and -resistant human rhabdomyosarcoma Rh30 cells, suggesting that our findings represent a general phenomenon. Ours is the first study to show that miR-485-3p mediates Top2α down-regulation in part by altered regulation of NF-YB.


Subject(s)
Antigens, Neoplasm/biosynthesis , CCAAT-Binding Factor/metabolism , DNA Topoisomerases, Type II/biosynthesis , DNA-Binding Proteins/biosynthesis , Etoposide/toxicity , Gene Expression Regulation, Enzymologic/drug effects , MicroRNAs/physiology , Antigens, Neoplasm/genetics , Antineoplastic Agents/toxicity , Cell Line, Tumor , DNA Topoisomerases, Type II/genetics , DNA-Binding Proteins/antagonists & inhibitors , DNA-Binding Proteins/genetics , Down-Regulation/drug effects , Drug Resistance, Neoplasm , Gene Expression Regulation, Enzymologic/physiology , Humans , Poly-ADP-Ribose Binding Proteins , Teniposide/toxicity , Up-Regulation/drug effects
6.
Z Naturforsch C J Biosci ; 56(9-10): 892-7, 2001.
Article in English | MEDLINE | ID: mdl-11724401

ABSTRACT

The combination of cis-diamminedichloroplatinum (II) (DDP, cisplatin) and topoisomerase II inhibitor teniposide (VM-26) has been shown to exert a synergistic effect in the clinical treatment of cancer. In this study, the combined effect of DDP and VM-26 on the growth and induction of apoptosis in synchronized murine erythroleukemia (MEL) cells, treated at the beginning or in the middle of S-phase of cell cycle, was examined. MEL cells, clone F4 N, were synchronized by a double thymidine block leading to accumulation of 70% of cells at the G1/S boundary. The growth-inhibitory effect of DDP and VM-26 applied alone were stronger in the middle of the S-phase than at the beginning. Morphological analysis showed that the majority of the cells revealed typical signs of apoptosis: nuclei fragmentation and appearance of apoptotic bodies. The combination of both agents at low concentrations had a synergistic effect on cytotoxicity. At higher concentrations the effect was additive. The remainder of the cells were characterized by unbalanced growth, aberrant mitosis and appearance of multinucleated cells. These processes led to delayed cell death. The appearance of aberrant mitosis was more expressed after treatment in the middle of the S-phase. It is likely that as a result of the combined action of cisplatin and VM-26, cells become supersensitive to the ability of topoisomerase II inhibitor to influence mitosis, and this increased sensitivity may contribute to the observed synergism.


Subject(s)
Antineoplastic Agents/toxicity , Cisplatin/toxicity , Leukemia, Erythroblastic, Acute/pathology , Mitosis/drug effects , Teniposide/toxicity , Animals , Clone Cells , DNA, Neoplasm/biosynthesis , Dose-Response Relationship, Drug , Kinetics , Mice , S Phase/drug effects
7.
Toxicology ; 163(2-3): 163-74, 2001 Jun 21.
Article in English | MEDLINE | ID: mdl-11516526

ABSTRACT

Alteration in the testicular weight and various germ cell populations was studied in male mice treated with different doses (0.05, 0.25, 0.5, 1.0 and 2.0 mg/kg b. wt.) of teniposide (VM-26) at various post-treatment time periods. Treatment of mice with different doses of teniposide did not significantly alter the testicular weights, irrespective of the drug dose used. Flow-cytometric analysis of germ cells of the untreated control mice testes revealed four distinct DNA peaks corresponding to elongated spermatids (HC), round spermatids (1C), spermatogonia and non-germ cells (2C) and primary spermatocytes (4C). The region between 2C and 4C peaks represents cells that are actively synthesizing DNA (S-phase cells). Treatment of mice with different doses of teniposide resulted in a significant depletion in the relative percentage of spermatogonia from day 2 to 35 post-treatment depending on the drug dose. DNA-synthesizing, i.e. S-phase, cells declined significantly at day 1 post-treatment and continued to decline up to day 70 post-treatment for all the drug doses studied, except 2 mg/kg drug dose at day 28 post-treatment. A significant decline in the relative percentage of primary spermatocytes (4C) was observed at day 7 that continued up to day 70 post-treatment depending on the drug dose. Round spermatids (1C) declined significantly at day 21 post-treatment after administration of 0.25--2.0 mg/kg VM-26. The relative percentage of elongated spermatids showed a significant decline at day 28 after 1 and 2 mg/kg drug treatment. These alterations in different germ-cell populations are reflected in the various germ-cell ratios. The 4C:2C ratio showed a significant decline at day 7 and 14 post-treatment after 1 and 2 mg/kg VM-26 treatment, while the 1C:2C ratio declined significantly at day 21 post-treatment in the mice treated with 0.5 and 2.0 mg/kg of VM-26. 4C:S-phase and 1C:4C ratios increased significantly from day 1 to 70 post-treatment, depending on the drug dose. Our study demonstrates that the treatment of mice with low doses of VM-26 exerts cytotoxic effects on various germ-cell populations.


Subject(s)
Antineoplastic Agents/toxicity , Spermatogenesis/drug effects , Spermatozoa/drug effects , Teniposide/toxicity , Animals , Body Weight , Diploidy , Dose-Response Relationship, Drug , Ethidium , Flow Cytometry , Fluorescent Dyes , Injections, Intraperitoneal , Male , Mice , Organ Size , Sperm Count , Spermatozoa/classification , Spermatozoa/pathology , Testis/drug effects , Testis/pathology , Time Factors
8.
Br J Cancer ; 79(5-6): 882-7, 1999 Feb.
Article in English | MEDLINE | ID: mdl-10070885

ABSTRACT

We conducted a phase I and pharmacokinetic study of the topoisomerase II catalytic inhibitor fostriecin. Fostriecin was administered intravenously over 60 min on days 1-5 at 4-week intervals. Dose was escalated from 2 mg m(-2) day(-1) to 20 mg m(-2) day(-1) in 20 patients. Drug pharmacokinetics was analysed with high performance liquid chromatography with UV-detection. Plasma collected during drug administration was tested in vitro for growth inhibition of a teniposide-resistant small-cell lung cancer (SCLC) cell line. The predominant toxicities were elevated liver transaminases (maximum common toxicity criteria (CTC) grade 4) and serum creatinine (maximum CTC grade 2). These showed only a limited increase with increasing doses, often recovered during drug administration and were fully reversible. Duration of elevated alanine-amino transferase (ALT) was dose-limiting in one patient at 20 mg m(-2). Other frequent toxicities were grade 1-2 nausea/vomiting, fever and mild fatigue. Mean fostriecin plasma half-life was 0.36 h (initial; 95% CI, 0-0.76 h) and 1.51 h (terminal; 95% CI, 0.41-2.61 h). A metabolite, most probably dephosphorylated fostriecin, was detected in plasma and urine. No tumour responses were observed, but the plasma concentrations reached in the patients were insufficient to induce significant growth inhibition in vitro. The maximum tolerated dose (MTD) has not been reached, because drug supply was stopped at the 20 mg m(-2) dose level. However, further escalation seems possible and is warranted to achieve potentially effective drug levels. Fostriecin has a short plasma half-life and longer duration of infusion should be considered.


Subject(s)
Antibiotics, Antineoplastic/adverse effects , Antibiotics, Antineoplastic/pharmacokinetics , Neoplasms/drug therapy , Adult , Aged , Alanine Transaminase/blood , Alkenes/adverse effects , Alkenes/pharmacokinetics , Alkenes/toxicity , Antibiotics, Antineoplastic/toxicity , Aspartate Aminotransferases/blood , Carcinoma, Non-Small-Cell Lung/drug therapy , Carcinoma, Small Cell , Colorectal Neoplasms/drug therapy , Dose-Response Relationship, Drug , Drug Resistance, Neoplasm , Female , Half-Life , Humans , Lung Neoplasms/drug therapy , Male , Metabolic Clearance Rate , Middle Aged , Neoplasms/blood , Polyenes , Pyrones , Teniposide/toxicity , Topoisomerase II Inhibitors , Tumor Cells, Cultured
9.
Toxicol Lett ; 104(3): 203-10, 1999 Feb 22.
Article in English | MEDLINE | ID: mdl-10079055

ABSTRACT

The effect of various doses (0-10 mg/kg body wt.) of teniposide (VM-26) was studied on the induction of micronuclei at 12, 24 and 36 h post-treatment. The frequency of micronuclei (MPCE and MNCE) increased in a dose-dependent manner up to a dose of 0.3125 mg/kg VM-26, where a peak frequency of micronuclei was observed. A further increase in the drug dose resulted in the reduction in micronuclei frequency in comparison with 0.3125 mg/kg drug dose reaching a nadir at 10 mg/kg. However, it was significantly higher than DDW (double distilled water) treated controls. The pattern of micronuclei induction was similar for all the post-treatment time periods. The frequency of micronuclei also increased with scoring time and the highest frequency of micronuclei was observed at 24 h post-treatment, which declined thereafter without restoration to DDW treated control level. Conversely, the PCE/NCE ratio registered a dose-dependent decline after treatment of mice with various doses of VM-26. A peak decline was observed at a dose of 0.3125 mg/kg, thereafter the decline became consistently less resulting in an elevation in the PCE/NCE ratio in comparison with 0.3125 mg/kg VM-26.


Subject(s)
Bone Marrow/drug effects , Micronuclei, Chromosome-Defective/drug effects , Teniposide/toxicity , Animals , Antineoplastic Agents/toxicity , Bone Marrow/ultrastructure , Dose-Response Relationship, Drug , Male , Mice
10.
Clin Cancer Res ; 4(6): 1367-73, 1998 Jun.
Article in English | MEDLINE | ID: mdl-9626452

ABSTRACT

Dexrazoxane (ICRF-187) is a catalytic inhibitor of the nuclear enzyme DNA topoisomerase II (topo II). It protects cells against topo II poisons, such as etoposide and teniposide, by hindering the DNA cleavage reaction of the target enzyme. We have previously shown that this antagonism also extends to an in vivo model. Thus, ICRF-187 protected mice against supralethal doses of etoposide and amsacrine, and the etoposide LD10 dose increased as much as 3.6-fold when combined with ICRF-187 (B. Holm, Cancer Chemother. Pharmacol., 38: 203-209, 1996). We describe here how scheduling of this drug combination can be optimized and used. Interestingly, ICRF-187 can protect when it is given after etoposide. Although timing is very critical here, ICRF-187 was able to completely protect when given 10 min after etoposide. This rescue principle resembles methotrexate rescue by folinic acid. We also found scheduling to be crucial because ICRF-187 did not protect when etoposide was given once a day for five days, whereas effective protection was seen when etoposide was used three times, once every four days. Similar investigations were performed with teniposide in combination with ICRF-187. The combination with ICRF-187 allowed a 3.4-fold teniposide dose escalation. Such dose escalations could be advantageous in specific situations. One such case is when the tumor is situated in a pharmacological sanctuary, e.g., in the brain. ICRF-187 is hydrophilic and does not cross the blood-brain barrier, whereas the lipophilic etoposide and teniposide do. Therefore, ICRF-187 would protect normal tissues and allow a cytotoxic dose of etoposide to reach the central nervous system (CNS). We therefore studied the combinations using L1210 or EHR2 cells inoculated into the CNS of mice. L1210 presented a leukemic CNS model with leptomeningeal spread and infiltration of liver, spleen, and lymph nodes, whereas EHR2 cells acted as a solid tumor with no evidence of extracerebral disease. In all experiments, the combination of high-dose etoposide and ICRF-187 was significantly superior to an equitoxic dose of etoposide alone. Such superiority was also seen when treatment was given on days 4, 8, and 12 after tumor inoculation. Here etoposide alone resulted in a mean increased life span of 12.3%, whereas the rescue regimen yielded an increase of 47% (P < 0.0001). In conclusion, DNA topo II rescue by catalytic inhibitors is a new strategy enabling significant epipodophyllotoxin dose escalations; in this study, we have demonstrated the superiority of this strategy in two in vivo CNS tumor models. This concept is now being tested in a clinical trial.


Subject(s)
Antineoplastic Agents/therapeutic use , Antineoplastic Combined Chemotherapy Protocols/therapeutic use , Brain Neoplasms/drug therapy , Etoposide/therapeutic use , Razoxane/therapeutic use , Teniposide/therapeutic use , Topoisomerase II Inhibitors , Amsacrine/therapeutic use , Amsacrine/toxicity , Animals , Drug Administration Schedule , Etoposide/administration & dosage , Etoposide/toxicity , Female , Mice , Mice, Inbred Strains , Survival Rate , Teniposide/administration & dosage , Teniposide/toxicity
11.
Mol Pharmacol ; 51(6): 1053-9, 1997 Jun.
Article in English | MEDLINE | ID: mdl-9187272

ABSTRACT

The role of the site selectivity of topoisomerase II poisoning in the cytotoxic activity of anthracyclines has not been established. In this article, we have thus studied the levels and persistence of double-stranded DNA breaks (DSB) along with the cytotoxic activity in human leukemic HL60 cells of seven anthracyclines, including doxorubicin, daunorubicin, and idarubicin, as well as sugar-modified analogues characterized by an altered sequence specificity. Epimerization at the 3' position of the sugar moiety markedly affected the biological activity; indeed, a dramatic reduction of drug effects was evident for 3'-deamino-3'-epi-hydroxy-4'-deoxy-4'-amino-daunorubicin. The studied analogues could be gathered into three groups based on the DSB/cytotoxicity ratio. At equitoxic concentrations: (a) parent drugs and 3'-deamino-3'-epi-hydroxy-4'-deoxy-4'-amino-daunorubicin endowed with the same sequence specificity stimulated low DSB levels; (b) 3'-epi-daunorubicin and 3'-deamino-4'-deoxy-4'-epi-amino-idarubicin, which have a different sequence specificity, and teniposide (a structurally unrelated poison) stimulated higher amounts of DSB; and (c) 4-demethoxy-3'-deamino-3'-hydroxy-4'-epi-doxorubicin stimulated the highest DSB levels. For the last agent, a faster rate of cleavage resealing, which is consistent with a reduced DNA binding affinity, could account for the increased DSB/cytotoxicity ratio compared with parent drugs. However, for other analogues, the observed differences in DSB persistence/resealing could not completely explain the different DSB/cytotoxicity ratios. The results thus suggest that the cytotoxic potency of anthracyclines may be the result of an interplay of the level, the persistence, and the genomic localization of topoisomerase II-mediated DNA cleavage.


Subject(s)
Antibiotics, Antineoplastic/toxicity , DNA Damage , DNA Topoisomerases, Type II/metabolism , DNA, Neoplasm/drug effects , DNA, Neoplasm/metabolism , Antineoplastic Agents, Phytogenic/toxicity , Apoptosis/drug effects , Cell Death/drug effects , HL-60 Cells/drug effects , HL-60 Cells/metabolism , HL-60 Cells/pathology , Humans , Kinetics , Nucleic Acid Synthesis Inhibitors/pharmacology , Substrate Specificity , Teniposide/toxicity
12.
Mutat Res ; 362(2): 181-91, 1996 Feb 15.
Article in English | MEDLINE | ID: mdl-8596537

ABSTRACT

To study the involvement of DNA topoisomerases in recombination in mammalian cells, we used gene transfer assays to examine the effects of DNA topoisomerase inhibitors on nonhomologous (illegitimate) and homologous recombination. The assays were performed by transfecting adenine phosphoribosyltransferase-deficient (APRT-) CHO cells with plasmids carrying the wild-type or mutant aprt genes and by treating the cells with the inhibitors, followed by subsequent cultivation to select for APRT-positive (APRT+) colonies. Treatments with DNA topoisomerase II inhibitors such as VP-16, VM-26, ICRF-193 resulted in a 3- to 5-fold stimulation of integration of both closed-circular and linearized plasmids carrying the wild-type aprt gene into the recipient genome through nonhomologous recombination. The same treatments also increased 6- to 9-fold and 3-fold the number of APRT+ recombinant colonies that were generated by cotransfecting two closed-circular plasmids with nonoverlapping defective aprt genes and their linearized equivalents, respectively. However, this cotransfection assay involved intrinsically nonhomologous recombination processes; normalization of the frequencies by dividing them with those of the above nonhomologous recombination revealed 2-fold enhancement of homologous recombination events between the circular mutant genes but not between the linear ones. In contrast, DNA topoisomerase I inhibitor, camptothecin, showed no such effect on either recombination. From these results, we discuss the function of DNA topoisomerases on recombination in mammalian cells.


Subject(s)
DNA Topoisomerases, Type II/metabolism , Enzyme Inhibitors/toxicity , Recombination, Genetic/drug effects , Topoisomerase II Inhibitors , Adenine Phosphoribosyltransferase/genetics , Animals , Antineoplastic Agents, Phytogenic/toxicity , CHO Cells/drug effects , CHO Cells/enzymology , Camptothecin/toxicity , Chromosome Aberrations/genetics , Cricetinae , DNA Topoisomerases, Type I/metabolism , Diketopiperazines , Dose-Response Relationship, Drug , Etoposide/toxicity , Mutagenesis , Mutagenicity Tests , Nucleic Acid Heteroduplexes , Piperazines/toxicity , Teniposide/toxicity , Topoisomerase I Inhibitors , Transfection
13.
Clin Cancer Res ; 1(1): 49-56, 1995 Jan.
Article in English | MEDLINE | ID: mdl-9815886

ABSTRACT

Topoisomerase II (Topo II) is an essential enzyme that catalyzes the breakage of double-strand DNA and is the target of several effective anticancer drugs, including the epipodophyllotoxins. The regulatory subunits of the cyclic AMP-dependent protein kinase are differentially expressed in normal and cancer cells. The RIalpha subunit is overexpressed in cells transformed by transforming growth factor-alpha (TGF-alpha) or Ha-ras oncogene. It has been shown that murine cells transformed by Ha-ras become hypersensitive to Topo II-targeting anticancer drugs. In this report we have tested whether any correlation exists between the expression of RIalpha protein and cellular sensitivity of Topo II-targeting drugs. Normal human breast MCF-10A cells and their derivatives overexpressing TGF-alpha, Ha-ras, or the different protein kinase subunits were treated with either Topo II inhibitors, such as etoposide, teniposide, or amsacrine, or with drugs which act independently of Topo II, such as bleomycin. Here we show that MCF-10A TGF-alpha and MCF-10A Ha-ras cells overexpress the RIalpha protein and become hypersensitive to epypodophyllotoxins and amsacrine but not to bleomycin. Direct introduction of the RIalpha gene into MCF-10A induces hypersensitivity to Topo II inhibitor drugs. In contrast, the overexpression of the other protein kinase subunits, RIIbeta or Calpha, does not modify the drug sensitivity of MCF-10A cells. No differences in the mRNA/protein content or in the activity of Topo II were found between hypersensitive cells and parental MCF-10A cells, suggesting that RIalpha may influence drug sensitivity via modulation of events downstream of the Topo II-DNA cleavable complex.


Subject(s)
Amsacrine/toxicity , Bleomycin/toxicity , Cyclic AMP-Dependent Protein Kinases/genetics , Etoposide/toxicity , Genes, ras , Teniposide/toxicity , Topoisomerase II Inhibitors , Transforming Growth Factor alpha/physiology , Breast Neoplasms , Cell Division/drug effects , Cell Survival/drug effects , Cyclic AMP-Dependent Protein Kinase Type II , Cyclic AMP-Dependent Protein Kinases/metabolism , DNA Topoisomerases, Type II/genetics , Female , Gene Expression Regulation, Enzymologic/drug effects , Gene Expression Regulation, Neoplastic/drug effects , Humans , Recombinant Proteins/metabolism , Transfection , Transforming Growth Factor alpha/genetics , Tumor Cells, Cultured
14.
J Cell Sci ; 107 ( Pt 10): 2887-98, 1994 Oct.
Article in English | MEDLINE | ID: mdl-7876355

ABSTRACT

The relative content of topoisomerase II (topo II) and the induction of topo-II-mediated DNA damage and cellular abnormalities have been characterized in developing spermatogenic cells of Xenopus laevis to gain an insight into the role of topo II during spermatogenesis. Decatenation assays identified topo II activity in nuclear extracts from spermatocytes and pre-elongate spermatids, but not in extracts from elongate spermatids or sperm. Extracts from early-mid spermatids contained 14% (per cell) of the decatenation activity found in spermatocyte extracts. Immunoblots of SDS extracts from whole cells and nuclei from both spermatocytes and pre-elongate spermatids, but not elongate spermatids or sperm, resolved a 180 kDa polypeptide that reacts with polyclonal antisera to Xenopus oocyte topo II, an antipeptide antibody (FHD29) to human topo II alpha and beta, and an antipeptide antibody to human topo II alpha, suggesting homology between Xenopus spermatogenic cell topo II and mammalian topo II alpha. Immunofluorescence microscopy of topo II in testis cryosections revealed the presence of topo II in nuclei of all spermatogenic stages, but not in sperm. The relative levels of topo II estimated from fluorescence intensity were highest in spermatogonia and spermatocytes, then early-mid spermatids, followed by elongate spermatids and somatic cells. Incubation of isolated spermatogenic cells with teniposide (VM-26), a topo II-targetted drug, resulted in a dose-dependent induction of DNA breaks in all spermatocytes and spermatid stages to nuclear elongation stages, as analyzed by alkaline single cell gel electrophoresis. Addition of 0.5-50 microM VM-26 to spermatogenic cell cultures for 27 hours resulted in stage-dependent abnormalities. Mid-late spermatid stages were relatively resistant to VM-26-induced damage. In contrast, meiotic division stages were arrested and spermatogonia B were killed by VM-26, and VM-26 induced abnormal chromosome condensation in pachytene spermatocytes. The results of these studies show that cellular levels of topo II are stage-dependent during spermatogenesis, that most spermatogenic stages are sensitive to topo II-mediated DNA damage, and that spermatogonia B, meiotic divisions and pachytene spermatocytes are particularly sensitive to induction of morphological abnormalities and cell death during acute exposure to topo II-targetted drugs.


Subject(s)
DNA Damage , DNA Topoisomerases, Type II/biosynthesis , Spermatogenesis/drug effects , Teniposide/toxicity , Animals , Cells, Cultured , Fluorescent Antibody Technique , Immunoblotting , Male , Spermatogenesis/genetics , Xenopus laevis
15.
Anal Biochem ; 220(1): 103-14, 1994 Jul.
Article in English | MEDLINE | ID: mdl-7526729

ABSTRACT

The studies described below were carried out to analyze the damage induced by DNA active drugs to episomal (Epstein-Barr virus, EBV) DNA in the Raji Burkitt's lymphoma cell line. This work: (i) applies pulsed-field gel electrophoresis (PFGE) techniques to quantify DNA damage on a large (approximately 180 kbp), circular target, (ii) investigates the DNA strand-scission behavior of different classes of drugs on the EBV episome, and (iii) compares EBV episomal damage to that generated in genomic DNA in the Raji cell line. Cells were treated with ionizing radiation to induce random strand scission, and the migration of topological forms of EBV was measured using PFGE. DNA damage induced in the episome by DNA active drugs was then assayed. Three drugs, acting by different types of DNA interactive mechanisms, were used: bleomycin, an intercalative DNA strand-scission agent; and amsacrine (mAMSA) and teniposide (VM26), intercalative and nonintercalative topoisomerase II active drugs, respectively. Rad equivalency of damage was determined by comparing the drug-induced change in percentage of Forms I and III to that generated by ionizing radiation. Additionally, single- and double-strand scission induced in genomic (total cellular) DNA by X-rays, bleomycin, amsacrine, and teniposide were assayed by high-sensitivity alkaline and neutral filter elution techniques. We demonstrate that pulsed-field gel electrophoresis is a useful technique for measuring form conversion in large episomal DNA. While all three drugs effect both episomal and genomic DNA strand scission, bleomycin appears to preferentially damage the EBV episome. The topoisomerase II active drugs mAMSA and VM26 show no evidence of episome-directed damage in this system and, in fact, damage genomic DNA at somewhat higher rates.


Subject(s)
Antineoplastic Agents/toxicity , Burkitt Lymphoma/drug therapy , DNA Damage , DNA, Viral/drug effects , Electrophoresis, Gel, Pulsed-Field , Herpesvirus 4, Human/genetics , Amsacrine/toxicity , Bleomycin/toxicity , Burkitt Lymphoma/genetics , Centrifugation, Density Gradient , Plasmids , Teniposide/toxicity , Tumor Cells, Cultured
16.
Mutat Res ; 312(2): 67-78, 1994 Apr.
Article in English | MEDLINE | ID: mdl-7510833

ABSTRACT

Frameshift mutations induced by acridines in bacteriophage T4 have been shown to be due to the ability of these mutagens to cause DNA cleavage by the type II topoisomerase of T4 and the subsequent processing of the 3' ends at DNA nicks by DNA polymerase or its associated 3' exonuclease followed by ligation of the processed end to the original 5' end. An analysis of the ability of nick-processing models is presented here to test the ability of nick processing to account for the DNA sequences of duplications and deletions induced in the aprt gene of CHO cells by teniposide (VM-26) [Han et al. (1993) J. Mol. Biol., 229, 52]. Although teniposide is not an acridine, it induces topoisomerase II-mediated DNA cutting in aprt sequences in vitro and mutagenesis in vivo. Although the previous study noted a correlation between mutation sites and nearby DNA discontinuities induced by the enzyme in vitro, neither the nick-processing model responsible for T4 mutations, nor double-strand break models alone were able to account for most of the mutant sequences. Thus, no single model explained the correlation between teniposide-induced DNA cleavage and mutagenic specificity. This report describes an expanded analysis of the ways that nick-processing models might be related to mutagenesis and demonstrates that a modified nick-processing model provides a biochemical rationale for the mutant specificities. The successful nick-processing model proposes that either 3' ends at nicks are elongated by DNA polymerase and/or that 5' ends of nicks are subject to nuclease activity; 3'-nuclease activity is not implicated. The mutagenesis model for nick-processing of teniposide-induced nicks in CHO cells when compared to the mechanism of nick-processing in bacteriophage T4 at acridine-induced nicks provides a framework for considering whether the differences may be due to cell-specific modes of DNA processing and/or due to the precise characteristics of topoisomerase-DNA intermediates created by teniposide or acridine that lead to mutagenesis.


Subject(s)
DNA Topoisomerases, Type II/metabolism , DNA/metabolism , Mutagenesis , Mutagens/toxicity , Phosphoric Diester Hydrolases/metabolism , Teniposide/toxicity , Acridines/metabolism , Acridines/toxicity , Adenine Phosphoribosyltransferase/genetics , Animals , Bacteriophage T4/genetics , Bacteriophage T4/metabolism , Base Sequence , CHO Cells/drug effects , Cricetinae , DNA/drug effects , DNA Damage , DNA Restriction Enzymes/metabolism , DNA, Bacterial/drug effects , DNA, Bacterial/metabolism , DNA-Directed DNA Polymerase/metabolism , Exodeoxyribonucleases/metabolism , Frameshift Mutation , Models, Genetic , Molecular Sequence Data , Mutagens/metabolism , Phosphodiesterase I , Repetitive Sequences, Nucleic Acid , Sequence Deletion , Substrate Specificity , Teniposide/metabolism , Topoisomerase II Inhibitors
17.
Cancer Chemother Pharmacol ; 34 Suppl: S127-33, 1994.
Article in English | MEDLINE | ID: mdl-8070021

ABSTRACT

Teniposide was introduced into clinical trials prior to etoposide, but its role was not defined because interest shifted early on to etoposide. However, long-term encouraging results obtained in acute leukemia treated with teniposide have rekindled interest in this compound. In addition to pharmacokinetic differences, teniposide has greater CNS penetrance and is more lipophilic. Its greater potency is related to enhanced intracellular uptake. Although its antitumor spectrum of activity appears to be very similar to that of etoposide, a search for some differences might prove instructive.


Subject(s)
Leukemia/drug therapy , Lymphoma/drug therapy , Neoplasms/drug therapy , Teniposide/toxicity , Teniposide/therapeutic use , Acquired Immunodeficiency Syndrome/complications , Adult , Brain Neoplasms/drug therapy , Carcinoma, Non-Small-Cell Lung/drug therapy , Carcinoma, Small Cell/drug therapy , Clinical Trials as Topic , Clinical Trials, Phase I as Topic , Etoposide/therapeutic use , Etoposide/toxicity , Female , Genital Neoplasms, Female/drug therapy , Hodgkin Disease/drug therapy , Humans , Lung Neoplasms/drug therapy , Lymphoma, Non-Hodgkin/drug therapy , Sarcoma, Kaposi/drug therapy , Sarcoma, Kaposi/etiology
18.
Cancer Chemother Pharmacol ; 34 Suppl: S14-8, 1994.
Article in English | MEDLINE | ID: mdl-8070022

ABSTRACT

We review herein current work on the cytotoxic and cellular actions of two classes of inhibitors of DNA topoisomerase II: one represented by etoposide and teniposide, which stabilize DNA-protein complexes, and another represented by merbarone and aclarubicin, which do not stabilize such complexes. We discuss current concepts of protooncogene activation and cell cycle perturbations by some of these inhibitors and summarize recent findings of novel actions of the latter compounds in tumor cells that express a mutant topoisomerase II.


Subject(s)
Antineoplastic Agents/toxicity , Drug Resistance , Topoisomerase II Inhibitors , Animals , Cell Line , Etoposide/toxicity , Humans , Models, Biological , Teniposide/toxicity , Tumor Cells, Cultured
19.
Cancer Chemother Pharmacol ; 34 Suppl: S89-95, 1994.
Article in English | MEDLINE | ID: mdl-8070034

ABSTRACT

We reported marked biologic activity with the epipodophyllotoxins in phase I/II studies of childhood cancer conducted in the 1970s. We have since extensively used the combination of teniposide and ara-C in the treatment of acute lymphoblastic leukemia (ALL). Initially we treated patients with refractory disease and found that the combination lacked clinical cross-resistance with standard antileukemic drugs. This formed a rationale to move teniposide and/or etoposide to front-line therapy of childhood ALL. The superior results projected for our last trial, an overall cure rate of about 75%, are attributable in part to early use of epipodophyllotoxins. This class of agents is also used extensively in the treatment of newly diagnosed childhood solid tumors, including neuroblastoma, medulloblastoma, rhabdomyosarcoma, and germ-cell tumors. Secondary leukemias following treatment with epipodophyllotoxins have been reported in a small subset of patients. Current data show that the most important risk factor is the schedule of drug delivery, which has led to appropriate protocol modifications.


Subject(s)
Etoposide/therapeutic use , Neoplasms/drug therapy , Precursor Cell Lymphoblastic Leukemia-Lymphoma/drug therapy , Teniposide/therapeutic use , Antineoplastic Combined Chemotherapy Protocols/therapeutic use , Child , Clinical Trials, Phase I as Topic , Clinical Trials, Phase II as Topic , Cytarabine/therapeutic use , Etoposide/administration & dosage , Etoposide/toxicity , Humans , Leukemia, Myeloid, Acute/drug therapy , Lymphoma, Non-Hodgkin/drug therapy , Neoplasms, Second Primary/chemically induced , Neuroblastoma/drug therapy , Teniposide/toxicity
20.
Mutat Res ; 267(1): 1-17, 1992 May.
Article in English | MEDLINE | ID: mdl-1373845

ABSTRACT

Various compounds were evaluated for their ability to induce prophage lambda in the Escherichia coli WP2s(lambda) microscreen assay. The inability of a DNA gyrase subunit B inhibitor (novobiocin) to induce prophage indicated that inhibition of the gyrase's ATPase was insufficient to elicit the SOS response. In contrast, poisons of DNA gyrase subunit A (nalidixic acid and oxolinic acid) were the most potent inducers of prophage among the agents examined here. This suggested that inhibition of the ligation function of subunit A, which also has a DNA nicking activity, likely resulted in DNA breaks that were available (as single-stranded DNA) to act as strong SOS-inducing signals, leading to prophage induction. Agents that both intercalated and produced reactive-oxygen species (the mammalian DNA topoisomerase II poisons, adriamycin, ellipticine, and m-AMSA) were the next most potent inducers of prophage. Agents that produced reactive-oxygen species only (hydrogen peroxide and paraquat) were less potent than adriamycin and ellipticine but more potent than m-AMSA. Agents that intercalated but did not generate reactive-oxygen species (actinomycin D) or that did neither (teniposide) were unable to induce prophage, suggesting that intercalation alone may be insufficient to induce prophage. These results illustrate the variety of mechanisms (and the relative effectiveness of these mechanisms) by which agents can induce prophage. Nonetheless, these agents may induce prophage by producing essentially the same type of DNA damage, i.e., DNA strand breaks. The potent genotoxicity of the DNA gyrase subunit A poisons illustrates the genotoxic consequences of perturbing an important DNA-protein complex such as that formed by DNA and DNA topoisomerase.


Subject(s)
Bacteriophage lambda/genetics , DNA Damage , Lysogeny/drug effects , Oxygen/toxicity , Topoisomerase II Inhibitors , Virus Replication , Amsacrine/toxicity , Azides/toxicity , Dactinomycin/toxicity , Dose-Response Relationship, Drug , Doxorubicin/toxicity , Ellipticines/toxicity , Escherichia coli , Free Radicals , Hydrogen Peroxide/toxicity , Nalidixic Acid/toxicity , Novobiocin/toxicity , Oxolinic Acid/toxicity , Paraquat/toxicity , SOS Response, Genetics , Sodium Azide , Teniposide/toxicity
SELECTION OF CITATIONS
SEARCH DETAIL
...