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.
Four 14C-labelled amphotericin B (Am B) derivatives with different net electric charges were examined: zwitterionic N-fructosyl Am B, positively charged N-fructosyl Am B methyl ester, negatively charged N-acetyl Am B and neutral N-acetyl Am B methyl ester. The binding of these four derivatives to human red cells and their octanol-water partition coefficients were measured. Simple partitioning between red cells and buffer was found for the four compounds, regardless of concentration, within a range of 10(-8) and 10(-4) M. This indicates the absence of cooperativity and saturability of binding at least in this concentration range. The constant partition coefficients were found to be three to five times higher for the two methyl ester derivatives than for the two non-esterified compounds. All partition coefficients were proportional to those found for the octanol-water system. Efficiency in inducing K+ leak from red cells was measured during the binding experiments. Despite the higher partition coefficients of the two methyl ester derivatives, they were found to have much lower ionophoric efficiency than the two non-esterified compounds. These results are discussed in terms of the mechanism of permeability pathway formation by polyene antibiotics.  相似文献   

2.
The relationship between polyene antibiotic binding to red cells and their membrane permeabilization was studied using two 14C-labelled amphotericin B (AmB) derivatives: N-fructosyl AmB and N-acetyl methyl ester AmB. The binding kinetics of both derivatives were determined on whole red cells and ghosts. The resulting experimental points were well fitted by monoexponential functions, and the characteristic t1/2 for both derivatives were calculated from these functions. At 2 X 10(-5) M, the half time t1/2 for N-acetyl methyl ester AmB (30.2 min) which forms aqueous aggregates was longer than the t1/2 for the more soluble species N-fructosyl AmB (4.5 min). At lower concentrations (10(-7) M), the t1/2 for N-acetyl methyl ester AmB (6.3 min) in a more solubilized form was close to that of N-fructosyl AmB (7.9 min). These results suggest that only solubilized species bound to red cell membranes and that disaggregation of aggregates is the limiting step in the binding process. The permeabilization of red cell membranes by N-fructosyl AmB, measured as intracellular K+ leak, was not instantaneous and at 10 degrees C external K+ was only detected 20 min after antibiotic addition. In contrast, binding occurs without lag time. Consequently, different mecanisms underlie binding and K+ permeability inducement. Absorption spectroscopy data showed that bound antibiotic is located in the hydrophobic membrane interior and that this penetration of the membrane by AmB derivatives occurs without lag time. Consequently, the lag time occurring for K+ permeability inducement would be due to some steps subsequent to binding and probably located in the hydrophobic membrane interior. This statement is further supported by the observation that the lag time is sensitive to changes in membrane fluidity as shown here by the break between 20 and 30 degrees C in the slope of the Arrhenius plot for the lag time, coinciding with the phase transition in red cell membranes.  相似文献   

3.
The activity of membrane-bound alkaline phosphatase (ALP) expressed on the external surface of cultured murine P19 teratocarcinoma and human HL-60 myeloblastic leukemia cells was studied at physiological pH using p-nitrophenylphosphate (pNPP) as substrate. The rate of substrate hydrolysis catalyzed by intact viable cells remained constant for eight successive incubations of 30 min and was optimal at micromolar substrate concentrations over the pH range 7.4-8.5. The value of apparent K(m) for pNPP in P19 and HL-60 cells was 120 microM. Hydrolytic activity of the ecto-enzyme at physiological pH decreased by the addition of levamisole, a specific and noncompetitive inhibitor of ALP (K(i) P19 = 57 microM; K(i) HL-60 = 50 microM). Inhibition of hydrolysis was reversed by removal of levamisole within 30 min. Retinoic acid (RA), which promotes the differentiation of P19 and HL-60 cells, induced levamisole-sensitive ecto-phosphohydrolase activity at pH 7.4. After its autophosphorylation by ecto-kinase activity, a 98-kDa membrane protein in P19 cells was found to be sensitive to ecto-ALP, and protein dephosphorylation increased after incubation of cells with RA for 24 h and 48 h. Orthovanadate, an inhibitor of all phosphatase activities, blocked the levamisole-sensitive dephosphorylation of the membrane phosphoproteins, while (R)-(-)-epinephrine reversed the effect by complexation of the inhibitor. The results demonstrate that the levamisole-sensitive phosphohydrolase activity on the cell surface is consistent with ecto-ALP activity degrading both physiological concentrations of exogenously added substrate and endogenous surface phosphoproteins under physiological pH conditions. The dephosphorylating properties of ecto-ALP are induced by RA, suggesting a specific function in differentiating P19 teratocarcinoma and HL-60 myeloblastic leukemia cells.  相似文献   

4.
Kim DS  Kim SH  Song JH  Chang YT  Hwang SY  Kim TS 《Life sciences》2007,81(25-26):1638-1644
Differentiation-inducing therapy by agents such as 1,25-dihydroxyvitamin D(3) [1,25-(OH)(2)D(3)] represents a useful approach for the treatment for cancer, including acute myeloid leukemia (AML). Recent studies demonstrated that the combined administration of 1,25-(OH)(2)D(3) and differentiation-enhancing agents could alleviate the side effects of 1,25-(OH)(2)D(3) and improve the rate of long term survival. In this study, we determined the enhancing activities of ceramide derivatives on 1,25-(OH)(2)D(3)-induced differentiation of human myeloid leukemia HL-60 cells. Importantly, some of these derivatives -- namely, A2, B3, and H9 -- enhanced the 1,25-(OH)(2)D(3)-induced differentiation of HL-60 cells in a concentration-dependent manner. In addition, the morphologic studies using Giemsa staining and flow cytometric analysis demonstrated that the combined treatment of 1,25-(OH)(2)D(3) with one of the three analogues, A2, B3, and H9, directed the HL-60 cells into monocytic lineage, but not into granulocytic lineage. The inhibition studies demonstrated that A2, B3, and H9, enhanced 1,25-(OH)(2)D(3)-induced differentiation of HL-60 cells via the PI3-K/PKC/JNK/ERK pathways. The ability of ceramide derivatives to enhance the differentiation-inducing potential of 1,25-(OH)(2)D(3) may contribute to an effective therapy for AML.  相似文献   

5.
We have previously reported that Brazilian propolis extracts inhibited growth of HL-60 human myeloid leukemia cells, which is partly attributed to the induction of apoptosis associated with granulocytic differentiation. In this study, we isolated three compounds which induce granulocytic differentiation evaluated by nitroblue tetrazolium (NBT)-reducing assays from the water extract of propolis and identified as 4,5-di-O-caffeoylquinic, 3,5-di-O-caffeoylquinic, and 3,4-di-O-caffeoylquinic acids by NMR analysis. Cell growth inhibitory activity of these caffeoylquinic acids was found in HL-60 cell, which was mainly attributed to the induction of apoptosis. Furthermore, the potency of caffeoylquinic acid derivatives to induce granulocytic differentiation was examined in HL-60 cells. Caffeic, quinic, and chlorogenic acids had no effects on the NBT-reducing activity, while 3,4,5-tri-O-caffeoylquinic acid induced more than 30% of NBT-positive cells. These results suggest that the number of the caffeoyl groups bound to quinic acid plays an important role in the potency of the caffeoylquinic acid derivatives to induce granulocytic differentiation. This is the first report demonstrating that the caffeoylquinic acid derivatives induce granulocytic differentiation of HL-60 cells.  相似文献   

6.
Fluoroquinolones and propionic acid derivatives are widely used antibacterials and non-steroidal anti-inflammatory drugs, respectively, which have been reported to frequently trigger drug hypersensitivity reactions. Such reactions are induced by inflammatory mediators such as cytokines and chemokines. The present study investigated whether levofloxacin, a fluoroquinolone, and loxoprofen, a propionic acid derivative, have the potential to induce immune-related gene expression in dendritic cell-like cell lines such as HL-60, K562, and THP-1, and immortalized keratinocytes such as HaCaT. The expression of IL-8, MCP-1, and TNFα messenger RNA (mRNA) was found to increase following treatment with levofloxacin or loxoprofen in HL-60 cells. In addition, these drugs increased the mRNA content of annexin A1, a factor related to keratinocyte necroptosis in patients with severe cutaneous adverse reactions. Inhibition studies using specific inhibitors of mitogen-activated protein (MAP) kinases and NF-κB suggest that the extracellular signal-regulated kinase (ERK) pathway is the pathway principally involved in the induction of cytokines and annexin A1 by levofloxacin, whereas the involvement of MAP kinases and NF-κB in the loxoprofen-induced gene expression of these factors may be limited. Fluoroquinolones and propionic acid derivatives that are structurally related to levofloxacin and loxoprofen, respectively, were also found to induce immune-related gene expression in HL-60 cells. Collectively, these results suggest that fluoroquinolones and propionic acid derivatives have the potential to induce the expression of immune-related factors and that an in vitro cell-based assay system to detect the immune-stimulating potential of systemic drugs might be useful for assessing the risk of drug hypersensitivity reactions.  相似文献   

7.
Activation of neutrophils leading to superoxide production is accompanied by cytoplasmic alkalinization, which results from stimulation of the Na+/H+ exchanger. Since the exchanger undergoes permanent alterations during neutrophilic maturation of HL-60 cells (Costa-Casnellie et al.: Journal of Biological Chemistry 263:11851-11855, 1988), we investigated whether its response to external stimuli such as phorbol esters or osmotic shock also was modified during cell maturation. Mature HL-60 cells produce superoxide in response to active phorbol esters, whereas immature HL-60 cells do not. Stimulation of the exchanger by active phorbol esters (phorbol 12-myristate 13-acetate or phorbol 12,13-dibutyrate) was observed in mature neutrophilic HL-60 cells but not in their immature counterparts. Inactive 4-alpha phorbol had no effect in either cell population. Compound H7 inhibited phorbol ester activation by 65%. In mature neutrophilic cells activation of the exchanger by phorbol esters caused two novel changes of its properties: 1) its apparent Km for Na+ transport increased 2-fold; 2) its Vmax increased 2.6-fold. Phorbol esters also caused a shift in pH dependence of activation similar to that induced in other cells. Osmotic shock, a different method known to activate the exchanger of other cells, induced activation in mature neutrophilic cells but not in immature cells. Thus, the response of the exchanger to external stimuli is affected by alterations occurring in association with cell maturation.  相似文献   

8.
Many sesquiterpene lactone compounds either induce or enhance the cell differentiation of human leukemia cells. However, we reported in a previous study that santonin, a eudesmanolide sesquiterpene lactone, exerts no effects on the differentiation of leukemia cells. In this report, to evaluate the possibility of chemically modifying santonin into its derivatives with differentiation inducing activity, we synthesized a series of santonin derivatives, and determined their effects on cellular differentiation in the human promyelocytic leukemia HL-60 cell system. A diacetoxy acetal derivative of santonin (DAAS) was found to induce significant HL-60 cell differentiation in a dose-dependent manner, whereas santonin in its original form did not. The HL-60 cells were differentiated into a granulocytic lineage when exposed to DAAS. In addition, the observed induction in cell differentiation closely correlated with the levels of NF-kappaB DNA binding activity inhibited by DAAS. Both Western blot analyses and kinase inhibitor studies determined that protein kinase C, ERK, and phosphatidylinositol 3-kinase were upstream components of the DAAS-mediated inhibition of NF-kappaB binding activity in HL-60 leukemia cells. The results of this study indicate that santonin can, indeed, be chemically modified into a derivative with differentiation inducing abilities, and suggest that DAAS might prove useful in the treatment of neoplastic diseases.  相似文献   

9.
The HL-60 cell line, established from a patient with acute promyelocytic leukemia, can be induced to undergo differentiation along the granulocyte or monocyte/macrophage line, depending on the particular inducer that is used. In this communication we provide evidence that HL-60 cells also have B lymphoid characteristics because by flow cytometry and clonal excess calculations, these cells are found to express immunoglobulin (Ig) lambda light chains on their surface. Furthermore, HL-60 cells contain poly(A)+ RNA that hybridizes with a DNA fragment encoding the constant region of Ig lambda chains and comigrates with lambda mRNA on RNA blots. Treatment of HL-60 cells with a phorbol ester that induces monocyte/macrophage differentiation resulted in the loss of surface Ig lambda chains and lambda RNA.  相似文献   

10.
The appearance of the arachidonic acid metabolic pathway in human promyelocytic leukemia (HL-60) cells was investigated during 1 alpha,25-dihydroxyvitamin D-3-induced monocytic differentiation. 1 alpha,25-Dihydroxyvitamin D-3-treated HL-60 cells acquired the ability to convert [1-14C]arachidonic acid to thromboxane B2 and prostaglandin E2 during monocytic differentiation. The major cyclooxygenase product synthesized by the HL-60 cells after 3-4 days exposure to 1 alpha,25- dihydroxyvitamin D-3 (48 nM) was thromboxane B2 and its production was about 19-25-times higher than that of untreated HL-60 cells. The percent conversion of thromboxane B2 from [1-14C]arachidonic acid in the 1 alpha,25-dihydroxyvitamin D-3 (48 nM, 3 day exposure)-treated HL-60 cells was about 4.4% as compared to that (about 0.3%) of the untreated cells, whereas the percent conversion of thromboxane B2 from [1-14C]prostaglandin H2 in the 1 alpha,25-dihydroxyvitamin D-3-treated cell homogenate was about 22.4% as compared to that (about 13.6%) of the untreated cell homogenate. The stimulatory effect of 1 alpha,25-dihydroxyvitamin D-3 on thromboxane B2 production from [1-14C]arachidonic acid and from [1-14C]prostaglandin H2 in HL-60 cells was inhibited by the addition of cycloheximide (1 microgram/ml). However, 1 alpha,25-dihydroxyvitamin D-3 (48 nM) did not significantly stimulate the arachidonic acid release either in HL-60 cells or in 1 alpha,25-dihydroxyvitamin D-3-induced cells. These results suggest that the stimulatory effect of 1 alpha,25-dihydroxyvitamin D-3 on the thromboxane production in HL-60 cells was not due to the activation of phospholipase A2 but due to the induction of fatty acid cyclooxygenase and thromboxane synthetase activities. Thromboxane A2 actively produced during the monocytic differentiation of HL-60 cells could influence the cell adhesiveness of the monocyte-macrophage-differentiated cells.  相似文献   

11.
Human promyelocytic leukemia cell line (HL-60) has been shown to be induced to the terminal differentiation into macrophage-like cells by a tumor promoter, 12-O-tetradecanoyl-phorbol-13-acetate (TPA). The present studies describe the effects of TPA on the phosphorylation of HL-60 cell proteins. A rapid decrease in the phosphorylation of a 75 kD protein was observed within a few minutes after treatment with TPA. On the other hand, TPA treatment of HL-60 cells caused rapid increase in the phosphorylation of a 67 kD protein and other minor proteins. Phorbol and 4α-phorbol-12,13-dodecanoate, both of which are biologically inactive derivatives of TPA, failed to cause any changes in protein phosphorylation in HL-60 cells. These results suggest that changes in protein phosphorylation are involved in mechanisms of the differentiation in HL-60 cells induced by TPA. Cell fractionation experiments revealed that 67K protein was located in cytosol. Though 75K protein also seemed to be located in cytosol, the phosphate moiety of 75K protein was almost lost during cell fractionation, suggesting that the phosphorylation of 75K protein was specifically regulated in HL-60 cells. Dimethyl sulfoxide (DMSO), retinoic acid (RA) and 1,25-dihydroxy-vitamin D3, all of which induce the differentiation in HL-60 cells, did not cause any changes in protein phosphorylation. These results suggest that the changes in protein phosphorylation are specific for TPA. The possible mechanisms of changes in protein phosphorylation by TPA were discussed.  相似文献   

12.
Huang ST  Yang RC  Chen MY  Pang JH 《Life sciences》2004,75(3):339-351
Phyllanthus urinaria (P. urinaria), a widely used herb medicine, was tested for the anticancer effect on human myeloid leukemia cells in this study. The water extract of P. urinaria induced the apoptosis of HL-60 cells as demonstrated by morphological change, DNA fragmentation and increased caspase-3 activity. However, normal human peripheral mononuclear cells remained viable under the same treatment. The P. urinaria-induced apoptosis of HL-60 cells was associated with the increased Bax gene expression and decreased Bcl-2 gene expression. In addition, the gene expressions of Fas receptor and Fas ligand, but not p53, were also induced in HL-60 cells dose- and time-dependently. The inhibitor of ceramide synthase, fumonisin B1, completely suppressed the apoptosis induced by P. urinaria and this inhibitory effect of fumonisin B1 could be eliminated by the addition of ceramide. It indicated that the activity of ceramide synthase is critical for the P. urinaria-induced apoptosis in HL-60 cells. The P. urinaria-induced apoptosis in HL-60 cells is mediated through a ceramide-related pathway.  相似文献   

13.
Cytokinins, N6-substituted adenine derivatives, are plant hormones playing important roles in various processes in plant development. Furthermore, cytokinins and their derivatives are able to control mammalian cell apoptosis and differentiation. The aim of our study was the synthesis of 7-azaindole derivatives as cytokinin analogues with the Hartwig-Buchwald coupling reaction in order to evaluate their biological properties on human myeloblastic leukaemia cells (HL-60 cell line). All these compounds presented a cytotoxic activity on HL-60 cells especially the 4-phenylaminopyrrolo[2,3-b]pyridine and the 4-phenethylaminopyrrolo[2,3-b]pyridine.  相似文献   

14.
Myeloid cells, including granulocytes and monocyte/macrophages, are important in disease-associated inflammatory reactions. These cells come from a common progenitor, the promyelocyte. The human promyelocytic cell line, HL-60, can be induced to terminally differentiate into granulocytes or monocyte/macrophages in a controlled fashion providing a model to study various aspects of myelomonocytic differentiation. The expression of several ion channels is controlled in HL-60 cells in a differentiation specific pattern. The purpose of this study was to determine if lineage-specific ion channel expression during HL-60 differentiation resulted in differences in functional responses to external stimuli. This was investigated by examining transmembrane potential responses in HL-60 promyelocytes, HL-60-derived polymorphonuclear cells (PMNs), and monocytes to various stimuli using the transmembrane potential sensitive dye, diSBAC2-(3). Exposure of HL-60 promyelocytes to ionomycin or ATP produced a membrane hyperpolarization. Studies using ion substitutions and ion channel blockers indicate that the hyperpolarization was mediated by KCa channels. During HL-60 promyelocyte differentiation to PMNs, the membrane potential response to ionomycin and ATP shifted from a hyperpolarization to a depolarization over 7 days. Conversely, HL-60-derived monocytes exhibited a membrane hyperpolarization in response to ionomycin and ATP. HL-60-derived monocytes also exhibit a Cl conductance specifically induced by ATP. Lineage-specific expression of ion channels during HL-60 cell differentiation is important in determining the transmembrane potential response of these cells. This may be translated into functional responses of various myelomonocytic cells during disease-associated inflammatory reactions. © 1996 Wiley-Liss, Inc.  相似文献   

15.
Addition of the polyene antibiotic amphotericin B or tissue culture medium to nondifferentiated HL-60 cells in the presence of luminol induces a chemiluminescence signal that reaches a peak value within a few seconds and decays exponentially in less than a minute. The kinetics of the signal and its modulation by superoxide dismutase, catalase, and horseradish peroxidase are consistent with a series of solution biochemical processes with a rate-determining step corresponding to the disproportionation of a luminol-superoxide complex. The effects of the enzymes demonstrate that superoxide is a precursor to the rate-determining intermediate and that both catalase and peroxide enhance a reaction that competes with the rate-limiting process.  相似文献   

16.
Km values for H2O2 and Vmax values for three types of myeloperoxidase (MPO) from human leukocytes (MPO-I, -II, and -III) and four types from human myeloid leukemia HL-60 cells (MPO-IA, -IB, -II, and -III) were determined. Km values of human leukocyte MPOs decreased with increasing pH from 4.4 to 6.2 and increased with increasing NaCl concentration from 0.025 to 0.14 M. There was no significant difference among Km values of leukocyte MPO-I, -II, and -III. NaBr also showed a tendency similar to that of NaCl with regard to the effects of pH and halide concentration on Km values. However, Km values in the presence of NaBr were lower than those in the presence of NaCl. Effects of pH and NaCl concentration on Vmax values of MPO-I, -II, and -III were also examined. Vmax values of MPO-I, -II, and -III were higher at pH 4.9 and 5.4 and increased with increasing NaCl concentration. In addition, no difference was observed between Km values of leukocyte and those of HL-60 cells. MPO-IB, the half-molecular-weight enzyme of HL-60 cells, also had the same Km values as the others. Furthermore, inhibition of the activities of seven MPOs of leukocytes and HL-60 cells by H2O2 was similarly observed at concentrations above 1 mM at pH 5.4 but not at pH 4.4. These results indicate that there is no difference in the affinity to H2O2 among leukocyte MPO-I, -II, and -III and HL-60 cell MPO-IA, -IB, -II, and -III.  相似文献   

17.
Exposure of HL-60 cells for 6 days to a combination of 1.25% (v/v) dimethyl sulfoxide and 10 microM dexamethasone induces myeloid differentiation which results in a cell with many of the characteristics of a mature granulocyte. At 4 degrees C myeloid differentiated, but not undifferentiated, monocytic differentiated or eosinophilic differentiated HL-60 cells display marked specific leukotriene B4 binding. Leukotriene B4 binding at 4 degrees C reaches a maximum within 10 min, is readily reversed by unlabeled leukotriene B4, and is stereospecific. Only molecules with structural and biological similarity to leukotriene B4 can competitively inhibit leukotriene B4 binding. Scatchard analysis at 4 degrees C in differentiated cells shows two classes of binding sites. The high affinity sites have a Kd of 0.27 nM and a Bmax of 14.8 fmol/10(7) cells; the low affinity sites have a Kd of 0.58 microM and a Bmax of 2453 fmol/10(7) cells. The appearance of specific leukotriene B4 binding sites in the myeloid differentiated cells correlates with their ability to chemotax in response to leukotriene B4. Undifferentiated cells do not chemotax to leukotriene B4. At 37 degrees C leukotriene B4 is incorporated into phospholipid and triglyceride species in both undifferentiated and myeloid differentiated HL-60 cells making binding studies at 37 degrees C in intact cells impossible. No evidence of omega-hydroxylase activity was found in HL-60 cells. These data suggest that the HL-60 cell may be an excellent model system for the study of leukotriene B4 receptor binding, processing, and gene expression.  相似文献   

18.
The properties of the Na+/H+ exchange system in the glial cell lines C6 and NN were studied from 22Na+ uptake experiments and measurements of the internal pH (pHi) using intracellularly trapped biscarboxyethyl-carboxyfluorescein. In both cell types, the Na+/H+ exchanger is the major mechanism by which cells recover their pHi after an intracellular acidification. The exchanger is inhibited by amiloride and its derivatives. The pharmacological profile (ethylisopropylamiloride greater than amiloride greater than benzamil) is identical for the two cell lines. Both Na+ and Li+ can be exchanged for H+. Increasing the external pH increases the activity of the exchanger in the two cell lines. In NN cells the external pH dependence of the exchanger is independent of the pHi. In contrast, in C6 cells, changing the pHi value from 7.0 to 6.5 produces a pH shift of 0.6 pH units in the external pH dependence of the exchanger in the acidic range. Decreasing pHi activates the Na+/H+ exchanger in both cell lines. Increasing the osmolarity of the external medium with mannitol produces an activation of the exchanger in C6 cells, which leads to a cell alkalinization. Mannitol action on 22Na+ uptake and the pHi were not observed in the presence of amiloride derivatives. Mannitol produces a modification of the properties of interaction of the antiport with both internal and external H+. It shifts the pHi dependence of the system to the alkaline range and the external pH (pHo) dependence to the acidic range. It also suppresses the interdependence of pHi and pHo controls of the exchanger's activity. NN cells that possess an Na+/H+ exchange system with different properties do not respond to mannitol by an increased activity of the Na+/H+ exchanger. The action of mannitol on C6 cells is unlikely to be mediated by an activation of protein kinase C.  相似文献   

19.
Summary Ascorbate is stabilized in the presence of HL-60 cells. Our results showed that cAMP derivatives and agents that increase cAMP stimulate the ability of HL-60 cells to stabilize ascorbate. On the other hand, tunicamycin, a glycosilation-interfering agent, inhibited this ability. The ascorbate stabilization in the presence of HL-60 cells has been questioned as a simple chemical effect. Further properties and controls about the enzymatic nature of this stabilization are described and discussed. This data, together with hormonal regulation, support the hypothesis that an enzymatic redox system located at the plasma membrane is responsible of the extracellular ascorbate stabilization by HL-60 cells.Abbreviations AFR ascorbate free radicals - FCS fetal calf serum - Sp-cAMPS Sp-cyclic adenosine monophosphothionate - Rp-cAMPS Rp-cyclic adenosine monophosphothionate  相似文献   

20.
33 polycyclic derivatives have been studied and tested on Leishmania donovani and L. major promastigotes. Their antileishmanial activity was assessed in vitro and an assay of their cytotoxicity was realized on human myelomonocytic cell line. The reference molecules used in the assays were amphotericin B and pentamidine. Among the compounds tested, 29 possess an antileishmanial activity; 25 of those were more active against L. donovani than amphotericin B, and nine were as effective as amphotericin B against L. major. Many synthesized derivatives were more active against L. donovani than against L. major. The cytotoxicity studies have shown that among the thirty-three derivatives tested, 12 molecules have an IC50 towards THP-1 cells about equal than that reference drugs, the 21 other derivatives are much less toxic. A 3D QSAR study was undertaken and has permitted to predict activity against L. donovani and L. major and to highlight critical area to optimize activity against the two species.  相似文献   

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