首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Energy-dependent drug efflux is believed to be a major factor in cellular resistance to doxorubicin (DOX). However, recent studies have shown that decreased retention alone cannot account for anthracycline resistance, and possibly other factors, such as drug metabolism, free radical scavengers, and altered DNA damage/repair, may be involved. We have measured DOX-induced DNA damage and its repair in P388 cells sensitive (P388/S) and resistant (P388/R) to DOX. Our studies show 2- to 5-fold less DNA damage, measured as protein-associated single-strand DNA breaks, in P388/R cells when compared to similarly treated P388/S cells. The repair of DNA in whole cells, expressed as percent DNA rejoined, was complete in 4 hr in P388/R, whereas no repair was seen in P388/S cells until 20 hr. No difference in repair of DNA lesions was observed when nuclei were used in repair experiments. The absence of repair in sensitive whole cells may be due to high retention or slow drug efflux. Increase of cellular DOX retention by exposure of cells to trifluoperazine (TFP) or verapamil (VPL) did not result in the increase of DNA damage in P388/R cells. DOX analogs, N-trifluoroacetyladriamycin-14-valerate (AD 32), 4'-O-tetrahydropyranyladriamycin (THP-adriamycin), and N-benzyladriamycin-14-valerate (AD 198), induced 2- to 4-fold more DNA damage than DOX in resistant cells. There was no difference in the poly(ADP-ribose) synthesis of P388/S and P388/R cells exposed to DOX or AD 32. Since ADP-ribose polymer synthesis is associated with free radical-induced DNA damage and is indicative of DNA repair by an excision-repair mechanism, data from these studies suggest that DNA breaks in anthracycline-exposed cells may not be due to free radical production but rather to other mechanisms, such as inhibition of DNA topoisomerase II activity. The present studies, in addition to emphasizing the role of DNA damage in resistance, also underscore the relative importance of DNA topoisomerase II function in anthracycline cytotoxicity.  相似文献   

2.
Energy-dependent drug efflux is a major factor in cellular resistance of P388/R84 mouse leukemic cells to anthracyclines such as doxorubicin (DOX), and blocking of efflux increases sensitivity. However, efflux does not play a significant role in resistance to N-trifluoroacetyladriamycin-14-valerate (AD 32), a DOX analog. Since drug efflux alone cannot account for resistance to anthracyclines, we have, in the present study, measured cellular glutathione (GSH) content and activity of GSH cycle related enzymes to determine their role in resistance. Cellular GSH content was similar in DOX-sensitive and -resistant mouse leukemic cells (P388 and P388/R84). GSH peroxidase, glucose-6-phosphate dehydrogenase and glutathione reductase activities were 1.36-, 1.58- and 1.14-fold higher in P388/R84 cells. Incubation of P388/R84 cells with 100 microM buthionine-S,R-sulfoximine (BSO) for 24 hr reduced cellular GSH content to 6% of control and reduced their resistance to DOX [dose modification factor (DMF) 3.9]. GSH depletion had no significant effect on the cytotoxicity of AD 32 (DMF 1.5). Exposure of P388/R84 cells to BSO (for GSH depletion) and trifluoperazine (for efflux blocking) further reduced their resistance to DOX (DMF 14). These results indicate that DOX resistance in P388/R84 cells is multifactorial and that changes in GSH cycle related enzymes such as GSH peroxidase may also contribute to their resistance.  相似文献   

3.
We have isolated an 84-fold adriamycin resistant subline, P388/R84, from mouse leukemia P388 cells by serial cultivation in methylcellulose in the presence of increasing drug concentrations. Electrophoresis of detergent soluble fractions of radiolabeled sensitive and resistant cells suggested marked alterations in the protein fractions of 160, 100, 60, 45, and 30 kd. In resistant clones labeled with 1251 an increase in 160 and 100 kd proteins was accompanied by concomitant reduction in the 60, 45, and 30 kd proteins. In 35S methionine-labeled resistant cells, similar increases in the 160 and 100 kd components were observed but in contrast to 125I-labeled cells the 30 kd component was also higher. Alterations in surface proteins were confirmed in experiments where the cell extracts were adsorbed to concanavalin A polymers and extracted with 0.26 M methyl-α-D-mannopyranoside. Our data confirm earlier reported observations on cell-surface protein changes in cells resistant to anthracyclines and alkaloids.  相似文献   

4.
The intracellular uptake, retention and cytotoxicity of pirarubicin, an anthracycline derivative, combined with chlorpromazine were investigated in doxorubicin-resistant mouse P388 leukemia (P388/DXR) cells. The number of viable cells was determined by the dye exclusion method. Chlorpromazine increased the cytotoxicity of pirarubicin in a dose-related manner in P388/DXR cells. A similar dose-response relationship was observed for chlorpromazine in increasing net intracellular pirarubicin accumulation. The accumulation was based on block of enhanced pirarubicin efflux from resistant cells by chlorpromazine. However, chlorpromazine did not affect cytotoxicity or transport of pirarubicin in the drug-sensitive cell line. The possible mechanisms of the restoration of pirarubicin sensitivity in P388/DXR cells by chlorpromazine are discussed.  相似文献   

5.
1. Overexpression of glutathione S-transferases (GST) in breast cancer cells is hypothesized to be a component of the multifactorial doxorubicin-resistant phenotype. 2. We have characterized the expression of GST enzymes at the catalytic activity, protein and mRNA levels in wild-type MCF-7 (MCF-7/WT) human breast cancer cells and a line selected for resistance to doxorubicin (MCF-7/ADR), with the goal of modulating GST activity to overcome resistance. 3. The MCF-7/ADR cells were 30-65-fold more resistant to doxorubicin than the MCF-7/WT cells. 4. Total cytosolic GST catalytic activity was elevated 23-fold in the MCF-7/ADR cells as compared with the MCF-7/WT cells, and the MCF-7/ADR cells also showed 3-fold increases in catalytic activity toward GST mu and alpha class-selective substrates. Neither cell line showed detectable catalytic activity with a GST mu class-selective substrate. 5. MCF-7/ADR cells showed pronounced overexpression of GST mu protein and GST P1 mRNA in comparison with the wild-type cell line. Neither cell line displayed detectable GST alpha or mu at the protein level. 6. A glutathione analogue that functions as a selective GST alpha inhibitor was more potent at inhibiting total cytosolic GST catalytic activity in the MCF-7/ADR cell line than GST alpha and mu class-selective inhibitory glutathione analogues and the non-selective GST inhibitor ethacrynic acid. 7. The multidrug resistance-associated protein, which can function as a glutathione-conjugate transporter, appeared weakly overexpressed in the MCF-7/ADR cells in comparison with the MCF-7/WT cells.  相似文献   

6.
We studied the effects of etoposide on the influx, intracellular accumulation and efflux of ara-C in P388 leukemic cells. Etoposide inhibited the active influx of ara-C in a dose dependent manner, and the inhibition was reversible. Etoposide also affected the intracellular accumulation of ara-C, though its inhibitory effect for the intracellular accumulation was weaker than that for the influx of ara-C. It was also shown that etoposide inhibited the active efflux of ara-C. Furthermore, etoposide inhibited the active transport of ara-C bidirectionally but the insensitive route of etoposide existed especially at high concentration of ara-C. The effect of inhibition of the transport of ara-C on the accumulation of ara-C was more significant at lower concentration of ara-C. Judging from the drug concentrations used in this study, an interaction between etoposide and ara-C could occur in the clinical treatment.  相似文献   

7.
8.
The distribution in Japanese monkey tissues of glutathione S-transferase activity toward some aromatic nitro compounds was examined by measuring the release of the nitro group as nitrite ion. The activity was especially high in liver, kidney and small intestine when compounds such as 4-nitroquinoline N-oxide, 5-nitrofurfural diacetal and o-dinitrobenzene were used as substrates. The nitrite-releasing activity of the major enzyme purified from rhesus monkey liver was also tested on fifty-two nitro compounds including nineteen nitrofuran derivatives. Among the thirty-three nitro compounds other than the nitrofuran derivatives tested as substrates, the purified enzyme showed activity only toward o-dinitrobenzene, 4-nitroquinoline N-oxide, 3,4-dinitrobenzoic acid, p-dinitrobenzene, 2,5-dinitrobenzoic acid, 2,5-dinitrophenol, tetra-chloronitrobenzene and 2,4-dinitrobenzoic acid. The crude supernatant fraction of rhesus monkey liver showed activity in substrate specificity roughly similar to that of the purified enzyme. On the other hand, among at least ten carcinogenic 2-substituted 5-nitrofran derivatives tested, 4,6-diamino-2-(5-nitro-2-furyl)-s-triazine, 5-nitro-2-furaldehyde semicarbazone, N-[[3-(5-nitro-2-furyl)-1,2,4-oxadiazol-5-yl]methyl] acetamide, and N-[5-(5-nitro-2-furyl)-1-3,4-thiadiazol-2-yl)acetamide were shown to be enzymatically conjugated with reduced glutathione. Among the other nine 2-substituted 5-nitrofuran derivatives tested, six compounds could be the substrates of the enzyme, and 5-nitrofurfural and 5-nitrofurfural diacetal were especially good substrates. There was, however, little apparent correlation between their carcinogenicity and susceptibility to glutathione S-transferase. The bulky substituents at position 2 appeared to decrease the susceptibility of these nitrofuran derivatives to the enzyme. Both Vmax and Km values of the purified enzyme varied greatly among the substrates, and the optimum pH fell between 7.5 and 9.0 in most cases.  相似文献   

9.
10.
11.
Male Sprague-Dawley rats were exposed in inhalation chambers to halothane and enflurane in concentrations from 50 ppm-1000 ppm (0.0025-0.05 minimum alveolar concentration; MAC) 6 h a day for 3-9 days. Repeated subanaesthetic concentrations were used to avoid effects of general anaesthesia and to increase the metabolized fraction of the inhaled anaesthetics. Exposure to 0.05 MAC of halothane (500 ppm) and enflurane (1000 ppm) for 9 days reduced the activity of glutathione S-transferases. A decrease in liver concentration of reduced glutathione (GSH) was observed after inhalation of enflurane, probably caused by metabolic release of inorganic fluoride. The results indicate a decreased detoxifying capacity of rat liver under the given conditions. Inhalation of occupational related concentrations of the anaesthetics (50 ppm) did neither affect the activity of the transferases nor the concentration of GSH in rat liver.  相似文献   

12.
We evaluated in vitro the antiproliferative activity of DPR (Fig. 1), a new cisplatin-derived compound, in combination with five conventional anticancer drugs: the antimetabolites 5-fluorouracil (5FU) and methotrexate (MTX), the alkylating agent mitomycin C (MMC), the antimicrotubule agent taxol (TAX) and the intercalating agent of the antracycline group doxorubicin (DOX), against murine P388 leukemic cells. MTT assay was used to determine growth inhibition after incubation of cells for 72 hours in the presence of single or combined drugs. The additive, synergistic or antagonistic nature of the combined drug effect was determined using the isobole method. In our cellular model, synergism was the prevailing result observed when DPR was combined with MMC. Conversely, antagonism was observed when DPR was combined with TAX. When DPR was administered together with the other antineoplastic drugs, the final effect was dependent on the concentrations of single agents. The study in vitro of the association between DPR and MMC was extended in vivo in BDF-1 female mice bearing i.p. P388 leukemic cells. Our data in vivo confirmed those obtained in vitro, demonstrating the therapeutic advantage of the association of ineffective doses of DPR (2 and 7 mg/kg) and MMC (3.2 mg/kg) over the administration of MMC alone.  相似文献   

13.
We evaluated in vitro the antiproliferative activity of DPR (Fig. 1), a new cisplatin-derived compound, in combination with five conventional anticancer drugs: the antimetabolites 5-fluorouracil (5FU) and methotrexate (MTX), the alkylating agent mitomycin C (MMC), the antimicrotubule agent taxol (TAX) and the intercalating agent of the antracycline group doxorubicin (DOX), against murine P388 leukemic cells. MTT assay was used to determine growth inhibition after incubation of cells for 72 hours in the presence of single or combined drugs. The additive, synergistic or antagonistic nature of the combined drug effect was determined using the isobole method.In our cellular model, synergism was the prevailing result observed when DPR was combined with MMC. Conversely, antagonism was observed when DPR was combined with TAX. When DPR was administered together with the other antineoplastic drugs, the final effect was dependent on the concentrations of single agents.The study in vitro of the association between DPR and MMC was extended in vivo in BDF-1 female mice bearing ip P388 leukemic cells. Our data in vivo confirmed those obtained in vitro, demonstrating the therapeutic advantage of the association of ineffective doses of DPR (2 and 7 mg/kg) and MMC (3.2 mg/kg) over the administration of MMC alone.  相似文献   

14.
15.
16.
Cytotoxic effect in vitro and antitumor effect in vivo of doxorubicin (DOX) combined with cepharanthine were investigated on DOX-resistant murine P388 leukemia (P388/R) cells. Cepharanthine was minimally cytotoxic in the cell line, but reversed DOX-resistance in a dose-related manner in P388/R cells. The administration of cepharanthine to mice bearing the P388 leukemia enhanced the antitumor activity of DOX. These results indicate that cepharanthine is an effective agent to reverse DOX-resistant cells.  相似文献   

17.
Atrazine is one of the most widely used herbicides in the United States and has been detected, occasionally, at low levels in drinking water sources. The biotransformation of atrazine in humans has not been fully characterized. Rodent studies suggest Phase I-dominated biotransformation with minor Phase II-mediated biotransformation by glutathione S-transferase(s) (GST). In human urine, mercapturates of atrazine are significant metabolites, yet the specific GST form(s) responsible for glutathione (GSH) conjugation have not been identified. Using recombinant alpha, mu, pi and theta class human GSTs, we demonstrated that only hGSTP1-1 displays significant activity toward atrazine (7.1 nmol/min/mg protein). We also confirmed that mouse GST Pi (pi) protein is responsible for the GSH-dependent biotransformation of atrazine in mouse liver; recombinant mGSTP1-1 had a specific activity of 7.3-nmol/min/mg protein. Furthermore, cytosolic fractions from mouse and human liver conjugated atrazine with glutathione at rates of 282.3 and 3.0 pmol/min/mg, respectively. Docking studies of the atrazine-GST conjugate in the hGSTP1-1 substrate-binding site were used to elucidate a basis for the dramatic difference in activity between mouse GSTP1-1 and GSTP2-2 (7.14 versus 0.02 nmol/min/mg protein, respectively). The inactivity of mGSTP2-2 appears to be attributable to an indirect structural disruption of the G-site by Pro12. Possible effects of the hGSTP1 polymorphisms were investigated. No significant differences in catalytic-specific activity were noted among purified proteins corresponding to the four hGSTP1 variants: hGSTP1(*)A (most common form), hGSTP1(*)B (Ile105Val), hGSTP1(*)C (Ile105Val, Ala114Val), and hGSTP1(*)D (Ala114Val). Overall, this work supports a physiological role for GSTs in atrazine biotransformation and indicates a novel diagnostic substrate for human and mouse GSTP1-1 proteins.  相似文献   

18.
月腺大戟水提物诱导P388白血病细胞的凋亡   总被引:1,自引:0,他引:1  
目的:研究月腺大戟水提物抗R388白血病及对瘤细胞凋亡的作用。方法:以一级清洁DBA/2型小鼠为实验对象,腹腔接种R388白血病细胞,胃饲10,2.5g·L^-1的月腺大戟溶液0.3mL,连续6d,检测外周血白细胞总数和腹腔瘤细胞总数,分离外周血淋巴细胞,观察瘤细胞及凋亡特征,计算细胞凋亡率,并检测bcl-2和bax凋亡相关基因的表达情况。结果:月腺大戟高剂量组外周血白细胞数量显著低于模型对照组(P〈0.01)。月腺大戟低剂量组外周血白细胞虽低于模型对照组,但差异无显著性(P〉0.05)。两剂量组的细胞凋亡率均明显高于模型对照组,两剂量组间的细胞凋亡率差异显著(P〈0.01)。凋亡细胞中可见核质浓缩、数个圆形凋亡小体位于胞浆,或呈半月形、块状靠近胞膜边缘,还可见核质松散呈网状,细胞膜不完整的退化淋巴细胞,表明高剂量明显诱导细胞凋亡(P〈0.01),低剂量虽能诱导细胞凋亡,但差异无显著性(P〉0.05)。结论:高剂量月腺大戟可通过诱导细胞凋亡明显抑制R388白血病瘤细胞的恶性增殖。  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号