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.
In vitro cultures of Crithidia sp. were exposed to various concentrations of hydroxyurea (HU) during the logarithmic phase. In the presence of 5 × 10?2M HU, cell division was completely blocked after an initial increase in cell numbers by about 20%. Inhibition of incorporation of 3H-thymidine into acid-insoluble material was effective within 1 hr of exposure to the drug (5 × 10?2M) and it reached a level of 80% after 8 hr. At lower concentrations (5 × 10?4M ? 1 × 10?3M), however, incorporation of 3H-thymidine was remarkably increased while cell division remained unaffected indicating that the increase in incorporation was not due to increased DNA synthesis in preparation for cell division.  相似文献   

2.
Intact and excised cultured pea roots (Pisum sativum L. cv Alaska) were treated with chlorsulfuron at concentrations ranging from 2.8 ×10?4 M to 2.8×10?6 M. At all concentrations this chemical was demonstrated to inhibit the progression of cells from G2 to mitosis (M) and secondarily from G1 to DNA synthesis (S). The S and M phases were not directly affected, but the transition steps into those phases were inhibited. Total protein synthesis was unaffected by treatment of intact roots with 2.8×10?6 M chlorsulfuron. RNA synthesis was inhibited by 43% over a 24-h treatment period. It is hypothesized that chlorsulfuron inhibits cell cycle progression by blocking the G2 and G1 transition points through inhibition of cell cycle specific RNA synthesis.  相似文献   

3.
The influence of 2‐methoxyestradiol (2‐ME) was investigated on cell numbers, morphology, cell cycle progression, and apoptosis induction in an oesophageal carcinoma cell line (WHCO3). Dose‐dependent studies (1 × 10?9M–1 × 10?6M) revealed that 2‐ME significantly reduced cell numbers to 60% in WHCO3 after 72 h of exposure at a concentration of 1 × 10?6M compared to vehicle‐treated cells. Morphological studies entailing light‐, fluorescent‐, as well as transmission electron microscopy (TEM) confirmed 2‐ME's antimitotic effects. These results indicated hallmarks of apoptosis including cell shrinkage, hypercondensation of chromatin, cell membrane blebbing, and apoptotic bodies in treated cells. Flow cytometric analyses demonstrated an increase in the G2/M‐phase after 2‐ME exposure; thus preventing cells from proceeding through the cell cycle. β‐tubulin immunofluorescence revealed that 2‐ME caused spindle disruption. In addition, increased expression of death receptor 5 protein was observed further supporting the proposed mechanism of apoptosis induction via the extrinsic pathway in 2‐ME‐exposed oesophageal carcinoma cells. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
The effect of cytokinin, kinetin, on abscisic acid (dormin) inhibition of α-amylase synthesis and growth in intact barley seed was investigated. Abscisic acid at 5 × 10?5M nearly completely inhibited growth response and α-amylase synthesis in barley seed. Kinetin reversed to a large extent abscisic acid inhibition of α-aniylase synthesis and coleoptile growth. The response curves of α-amylase synthesis and coleoptile growth in presence of a fixed amount of abscisic acid (6 × l0?6M) and increasing concentrations of kinetin (from 5 × l0?7M to 5 × 10?5 M) showed remarkable similarity. Kinetin and abscisic acid caused synergistic inhibition of root growth. Gibberellic acid was far less effective than kinetin in reversing abscisic acid inhibition of α-amylase synthesis and coleoptile growth. A combination of kinetin and gibberellic acid caused nearly complete reversal of abscisic acid inhibition of α-amylase synthesis but not the abscisic acid inhibition of growth. The results suggest that factors controlling α-amylase synthesis may not have a dominant role in all growth responses of the seed. Kinetin possibly acts by removing the abscisic acid inhibition of enzyme specific sites thereby allowing gibberellic acid to function to produce α-amylase.  相似文献   

5.
The cytotoxic action of 2-thiouracil, 2-thiocytosine, 2-thiouridine and 4-thiouridine was studied in cultures of a clone of Chinese hamster cells with a generation time of 16 hours (S — 8 hours, G2 — 2 hours, and G1 plus M — 6 hours). The cells were synchronized at metaphase by the method of reversal of colcemid inhibition and cell survival was measured by their colony-forming ability. The four analogs induced cytotoxic effects which increased with the concentration of the chemical and the length of the exposure time. Exposure to 4 × 10?4 M 2-thiocytosine, 2-thiouridine or 4-thiouridine for a period of 20 hours reduced cell survival to less than 10% of the controls. The other analog (2-thiouracil) was less effective when tested at similar concentrations and time of exposure and decreased the survival to only 35% of the controls. Short periods of treatment (one hour) produced little effect at concentrations of 4 × 10?5M, and affected the survival of cells differently when 4 × 10?4 M were administered at different stages of the cell cycle. Two peaks of maximum sensitivity, one at late G1 and the other at G2 were observed. These peaks correspond to the peaks of maximum RNA synthesis described for synchronized mammalian cells. Therefore, it is likely that the cytotoxic effects of thiopyrimidine analogs are related to interference with RNA synthesis.  相似文献   

6.
The effects of 5-HT and glutamate on dopamine synthesis and release by striatal synaptosomes were investigated and compared with the action of acetylcholine, which acts presynaptically on this system. 5-HT inhibited (28%) synthesis of [14C]dopamine from L-[U-14C]tyrosine, at 10-5M and above. This contrasts with the action of acetylcholine, which stimulated [14C]-dopamine synthesis by 24% at 10-4 M. Tissue levels of GABA were unaffected by either 5-HT or acetylcholine up to concentrations of 10-4 M. The inhibitory action of 5-HT (5 × 10?5 M and 2 × 10?4 M) on [19C]dopamine synthesis was completely abolished by methysergide (2 × 10?6 M). Higher concentrations of methysergide (10?4 M) or cyproheptadine (10?5 M) inhibited [14C]dopamine synthesis by 28% and 25%, respectively, when added alone to synaptosomes. However, only methysergide prevented the further inhibition of synthesis caused by 5-HT. At concentrations of 2 × 10?5 M and above, 5-HT stimulated [14C]dopamine release. This releasing action differed from that of acetylcholine, which occurred at lower concentrations (e.g., 10?6 M). Methysergide (up to 10?4 M) or cyproheptadine (2 × 10?4 M) did not reduce the 5-HT (5 × 10?5 M)-induced release of [14C]dopamine, but methysergide (10?4 M) showed a potentiation (49%) of this increased release. The stimulatory effects of 5-HT (2 × 10?5 M) and K+ (56 mM) on [14C]dopamine release were additive, indicating that two separate mechanisms were involved. However, when both agents were present the stimulatory effect of K+ (56 mM) on [14C]dopamine synthesis was not seen above the inhibitory effect of 5-HT. Glutamate (0.1-5 mM) did not affect [4C]dopamine release or its synthesis from L-[U-14C]tyrosine. It is concluded that 5-HT modulates the synthesis of dopamine in striatal nerve terminals through a presynaptic receptor mechanism, an action antagonised by methysergide. The releasing action of 5-HT apparently occurs through a separate mechanism which is also distinct from that involved in the response to K+ depolarisation.  相似文献   

7.
The effects of retinoic acid (RA) on cell proliferation, activity of acid phosphatase, protein synthesis and methionine uptake were studied in transformed murine LPA cells. Early inhibition of protein synthesis was demonstrated under experimental conditions in which the rate of cell proliferation was diminished and non-specific effects of vitamin action could be excluded. Measurements of l-methionine uptake revealed a decrease to approximately one-half of that in control cultures after treatment with RA at the concentrations of 5 × 10?5 M and 10?5 M.  相似文献   

8.
Inhibition of bovine erythrocyte acetylcholinesterase (free and immobilized on controlled pore glass) by separate and simultaneous exposure to malathion and malathion transformation products which are generally formed during storage or through natural or photochemical degradation was investigated. Increasing concentrations of malathion, its oxidation product malaoxon, and its isomerisation product isomalathion inhibited free and immobilized AChE in a concentration-dependent manner. KI, the dissociation constant for the initial reversible enzyme inhibitor-complex, and k3, the first order rate constant for the conversion of the reversible complex into the irreversibly inhibited enzyme, were determined from the progressive development of inhibition produced by reaction of native AChE with malathion, malaoxon and isomalathion. KI values of 1.3 × 10? 4 M? 1, 5.6 × 10? 6 M? 1 and 7.2 × 10? 6 M? 1 were obtained for malathion, malaoxon and isomalathion, respectively. The IC50 values for free/immobilized AChE, (3.7 ± 0.2) × 10? 4 M/(1.6 ± 0.1) × 10? 4, (2.4 ± 0.3) × 10? 6/(3.4 ± 0.1) × 10? 6 M and (3.2 ± 0.3) × 10? 6 M/(2.7 ± 0.2) × 10? 6 M, were obtained from the inhibition curves induced by malathion, malaoxon and isomalathion, respectively. However, the products formed due to photoinduced degradation, phosphorodithioic O,O,S-trimethyl ester and O,O-dimethyl thiophosphate, did not noticeably affect enzymatic activity, while diethyl maleate inhibited AChE activity at concentrations > 10 mM. Inhibition of acetylcholinesterase increased with the time of exposure to malathion and its inhibiting by-products within the interval from 0 to 5 minutes. Through simultaneous exposure of the enzyme to malaoxon and isomalathion, an additive effect was achieved for lower concentrations of the inhibitors (in the presence of malaoxon/isomalathion at concentrations 2 × 10? 7 M/2 × 10? 7 M, 2 × 10? 7 M/3 × 10? 7 M and 2 × 10? 7 M/4.5 × 10? 7 M), while an antagonistic effect was obtained for all higher concentrations of inhibitors. The presence of a non-inhibitory degradation product (phosphorodithioic O,O,S-trimethyl ester) did not affect the inhibition efficiencies of the malathion by-products, malaoxon and isomalathion.  相似文献   

9.
Prostaglandin D2 was found to be a potent inhibitor of B-16 melanoma cell replication in vitro. The inhibition was dose-dependent between 3×10?9M and 3×10?6M (IC50~ 0.3 μM after 6 days). On a molar basis, PGD2 was a better inhibitor than PGA2 or 16,16-dimethyl-PGE2-methyl ester (di-M-PGE2) and in higher concentrations (10?6?10?7M), comparable to retinoic acid. In higher concentrations, PGD2 inhibited DNA, RNA and protein synthesis. The B-16 melanoma cell line which we used synthesized arachidonic acid metabolites which comigrated with PGA2, PGD2, PGE2 and PGF on a thin layer chromatography system.  相似文献   

10.
The reaction kinetics of APS-kinase from Chlamydomonas reinhardii showed that the enzyme formed PAPS from APS upon the addition of ATP. Evidence for a 35S-labelled protein intermediate between APS and PAPS has been obtained. The APS-kinase activity could only be measured in the presence of low concentrations of APS (20 ± 10 μM) and of ATP (0.2 ± 0.05 mM) due to substrate inhibition. The inhibition was partially overcome by low concentrations of 3′,5′-PAP (10,μM). The rates of PAPS formation obtained with cell extracts from the alga varied from 2 to 6 nM PAPS/mg protein/min (33–100 × 10?12 kat/mg).  相似文献   

11.
The inhibition of growth by different concentrations of CdCl2 in the range 4,5 × 10?7 to 5.6 × 10?7M was studied in the green alga Coelastrum proboscideum Bohlin in inorganic media at pH 4.3, 5.3 and 6.2. The factorial destgn of the experiments was evaluated as an analysis of 22 factors. Below pH 4.0 and above pH 6.5 growth was depressed without adding Cd. Cd concentrations exceeding 5.6 × 10?8M reduced algal growth significantly with a 50% inhibition at 5.6 × 10?7M Cd. The Cd concentration of 5.6 × 10?7M was less toxic at pH 6.2 than at pH 5.3 and 4.3, thus revealing a negative interaction between protons and Cd.  相似文献   

12.
Abstract: Pridefine (AHR-1118) is a pyrrolidine derivative with clinically established antidepressant efficacy. Previous work from this laboratory indicates that pridefine is a reuptake blocker of catecholamines and serotonin with weak releasing activity. This study characterized the mode of amine uptake inhibition by pridefine as noncompetitive. The uptake experiments were performed utilizing ouabain instead of zero-degree controls to differentiate between the passive and active components of uptake. Furthermore, the passive component was resolved into diffusion and binding of substrate. Correction was made for the effects of ouabain on binding. Kinetic constants determined from Lineweaver-Burk plots were: Km= 3 × 10?7 M for NE, Km= 9 × 10?8 M for DA, and Km= 3 × 10?8 M for 5-HT. Dixon analyses of uptake at various pridefine concentrations indicated noncompetitive inhibition with Ki= 2.5 × 10?6 M for NE uptake, Ki= 2.0 × 10?6 M for DA uptake, and Ki= 1 × 10?5 M for 5-HT uptake. These constants compare well with IC50 values for the same transmitters: NE, IC50= 2.4 × 10?6 M; DA, IC50= 2.8 × 10?6 M; 5-HT, IC50= 1.0 × 10?5 M. The in vitro results indicate that pridefine is relatively specific as a catecholamine uptake blocker. It differs from tricyclic antidepressants which are reportedly competitive inhibitors of monoamine uptake. The possible mechanisms by which pridefine acts as a noncompetitive inhibitor are discussed.  相似文献   

13.
The activity of concentrated exsheathing fluid of Haemonchus contortus against isolated sheaths was not inhibited by ethylenediamine tetra-acetic acid (EDTA), 10?2 M, even when the concentrations of Mg and Mn were < 4 × 10?4 M and < 0·9 × 10?6 M respectively. Purified or diluted solutions of exsheathing fluid, even in the presence of Mg2+, 10?3 M, were inhibited. Leucine aminopeptidase (LAP) in exsheathing fluid was active even at concentrations of Mg < 1·3 × 10?5M. Concentrated solutions were partially inhibited by EDTA, 10?2 M, at low concentrations of Mg; inhibition was increased in diluted and purified preparations.1,10-phenanthroline (Ophen) strongly inhibited exsheathing activity (Zn < 1 × 10?6 M). When Zn2+, 10?3 M was added, the inhibition was abolished. The hydrolysis of l-leucinamide was greatly increased in the presence of Ophen, 10?4 M; this effect was abolished by adding Zn2+, 10?3 M.It is suggested that exsheathing fluid from at least some ‘strains’ of H. contortus contains a Zn metallo-enzyme, probably LAP, which is involved in the process of exsheathment.  相似文献   

14.
We investigated the effects of zinc or lead on growth and on exudation of fluorescent dissolved organic matter (FDOM) by the marine toxic dinoflagellate Alexandrium catenella (Whedon & Kofoid) Balech. The species was exposed to increasing free zinc (1.34 × 10?7 M–3.98 × 10?6 M) or lead (5.13 × 10?9 M–1.82 × 10?7 M) concentra‐tions. Low metal levels ([Zn2+] = 1.34 × 10?7 M; [Pb2+] = 5.13 × 10?9 M) had no effect on cell growth. Toxic effects were observed from higher metal contamination ([Zn2+] = 3.98 × 10?6 M; [Pb2+] = 6.54 × 10?8 M), as a conversion of vegetative cells into cysts. Analysis of the released FDOM by three‐dimensional (3‐D) fluorescence spectroscopy was achieved, using the parallel factor analysis (PARAFAC). The PARAFAC modeling revealed four components associated with two contributions: one related to the biological activity; the other linked to the organic matter decomposition in the culture medium. The C1 component combined a tryptophan peak and characteristics of humic substances, whereas the C2 component was considered as a tryptophan protein fluorophore. The two others C3 and C4 components were associated with marine organic matter production. Relea‐sed fluorescent substances were induced by low ([Zn2+]= 1.34 × 10?7 M; [Pb2+] = 5.13 × 10?9 M) and moderate ([Zn2+] = 6.21 × 10?7 M; [Pb2+] = 2.64× 10?9 M) metal concentrations, suggesting the activation of cellular mechanisms in response to metal stress, to exudate FDOM that could complex metal cations and reduce their toxicity toward A. catenella cells.  相似文献   

15.
Subacute methyl mercury (MeHg) intoxication was induced in adult rats following the daily intragastric administration of 1 mg MeHg/100 g body weight. Decreased [14C]leucine incorporation into cerebral and cerebellar slice protein was found. Weight loss occurred during the latent and neurotoxic phases but pair feeding did not reveal a significant defect in amino acid incorporation into slice protein. There was no decline in synaptosome protein synthesis in vitro during the latent phase but a significant decline in cerebellar and cerebral synaptosome synthesis was found during the neurotoxic phase. MeHg in vitro inhibited cerebral slice and synaptosome protein synthesis at half maximal concentrations of 7.5 and 12.5 μM respectively. Inhibition of synthesis in synaptosomes was non-competitive with K1 of 4 × 10?6M. MeHg had no effect on [14C]leucine or [14C]proline uptake into synaptosomes. There was no significant inhibition of synaptosome basal ATPase or Na + K ATPase at concentrations of MeHg (12 μM) giving half maximal inhibition of protein synthesis. No preferential inhibition of the chloramphenicol (55S) or cycloheximide sensitive components of synaptosome fraction protein synthesis was found, suggesting that the inhibition is common to both mitochondrial and extramitochondrial protein synthesizing systems. Addition of nucleotides and/or atractylate failed to influence protein synthesis and did not reverse the MeHg inhibition. Mannitol, as a replacement for the predominant cation species of the incubation medium, gave 40% inhibition of protein synthesis in the control but protected against further inhibition by MeHg.  相似文献   

16.
The effects of colchicine and 2-Br-α-ergocryptine-methane-sulfonate (CB 154) on the release of prolactin and growth hormone have been studied in a clonal strain of rat pituitary tumor cells (GH3) in monolayer culture. These cultures produce both prolactin and growth hormone and release both proteins spontaneously into the medium without storing them in large amounts. Immunological methods were used to measure both intracellular and extracellular concentrations of the hormones. Colchicine (5 × 10?6 M for 3 hours) caused a 2- to 3-fold increase in intracellular concentrations of prolactin and growth hormone but, under basal conditions, had little or no measurable effect on the amounts of hormone accumulated in the medium during the course of the standard three hour treatment period. This latter finding evidently is due to a lag in the onset of drug action. Colchicine had little or no effect on accumulation of extracellular prolactin during the first two hours of treatment whereas such accumulation was depressed by over 60% during the third hour of treatment. Previous studies have shown that treatment of GH3 cells with thyrotropin releasing hormone (TRH) and hydrocortisone (HC) increases both intra and extracellular levels of prolactin and growth hormone, respectively. In cultures treated with TRH (5 × 10?8 M), colchicine (5 × 10?6 M for 3 hours) increased intracellular prolactin by about 70% and decreased extracellular hormone by 10%. In cultures treated with HC (3 × 1O?6 M), colchicine increased intracellular growth hormone by more than 100% and decreased medium concentrations of the hormone by 15%. Colchicine did not significantly alter total hormone (intracellular + extracellular) accumulation, cellular uptake of 3H-amino acids, or total cell protein synthesis. The synthetic ergot alkaloid, CB 154, (3.3 × 10?6 M for 3 hours) caused an 80% increase in intracellular, and a nearly 50% decrease in extracellular, prolactin without affecting the accumulation of growth hormone, the uptake of 3H-labeled amino acids, or overall protein synthesis in the cultures. Elevation of medium potassium concentration from a basal value of 5.3 mM to 3–5 × 10?2 M (by addition of KCl) decreased intracellular levels of prolactin by 85% and growth hormone by 55%. These effects of high potassium were blocked by colchicine and by CB 154. We conclude that colchicine, after a lag period of two hours, acts to inhibit the release of prolactin and growth hormone from GH3 cells. By the end of three hours of treatment, this inhibition is over 60% complete in the case of prolactin. The qualitatively different effects of colchicine and CB 154 on prolactin and growth hormone release suggest that these two secretory blocking agents probably act on GH3 cells by different mechanisms.  相似文献   

17.
SYNOPSIS. Continuous growth of one cell line (UCI variant) of Leishmania tarentolae was achieved in the absence of organic sulfur. These cells were able to use sodium sulfate, and, to a limited extent, sodium sulfite as their sole sulfur source and could utilize methionine sulfoxide in place of L-methionine. A related cell line (RU variant) was unable to grow in organic sulfate-free media nor could these cells utilize methionine sulfoxide. UCI promastigotes incorporated significant amounts of 35S sodium sulfate; killed cells did not take up the label. 35S incorporation was inhibited by sodium molybdate (5 × 10?4 M), sodium arsenite (5 × 10?4 M), 2,4-dinitrophenol (1 × 10?4 M), or KCN (5 × 10?4 M). RU promastigotes did not incorporate significant amounts of 35S sodium sulfate. Thin layer chromatographs of protein hydrolysates from UCI cells incubated in 35S sodium sulfate revealed several radio opaque spots, one of which had chromatographic properties of cystine. UCI variants of L. tarentolae were therefore capable of assimilatory sulfate reduction whereas RU cells lacked this ability.  相似文献   

18.
Previous studies from this laboratory have indicated that tricyclohexyltin hydroxide (Plictran) is a potent inhibitor of both basal- and isoproterenol-stimulated cardiac sarcoplasmic reticulum (SR) Ca2+-ATPase, with an estimated IC-50 of 2.5 × 10?8M. The present studies were initiated to evaluate the mechanism of inhibition of Ca2+-ATPase by Plictran. Data on substrate and cationic activation kinetics of Ca2+-ATPase indicated alteration of Vmax and Km by Plictran (1 and 5×10?8M), suggesting a mixed type of inhibition. The beta-adrenergic agonist isoproterenol increased Vmax of both ATP- and Ca2+-dependent enzyme activities. However, the Km of enzyme was decreased only for Ca2+ Plictran inhibited isoproterenol-stimulated Ca2+-ATPase activity by altering both and Vmax and Km of ATP as well as Ca2+-dependent enzyme activities, suggesting that after binding to a single independent site, Plictran inhibits enzyme catalysis by decreasing the affinity of enzyme for ATP as well as for Ca2+ Preincubation of enzyme with 15 μM cAMP or the addition of 2mM ATP to the reaction mixture resulted in slight activation of Plictran-inhibited enzyme. Pretreatment of SR with 5 × 10?7M propranolol and 5 × 10?8M Plictran resulted in inhibition of basal activity in addition to the loss of stimulated activity. Preincubation of heart SR preparation with 5 × 10?5M coenzyme A in combination with 5 × 10?8M Plictran partly restored the beta-adrenergic stimulation. These results suggest that some critical sites common to both basal- and beta-adrenergic-stimulated Ca2+-ATPase are sensitive to binding by Plictran, and the resultant conformational change may lead to inhibition of beta-adrenergic stimulation.  相似文献   

19.
The effects of calcium ions and of the calcium channel blockers verapamil, diltiazem and nifedipine on galvanotaxis in Chlamydomonas have been investigated using a fully automated and computerized population system. Galvanotaxis is a function of the voltage applied to the cell population. However, the galvanotactic orientation also depends on the external calcium concentration. In a calcium-deprived nutrient medium which still contains 6 × 10?7M calcium, galvanotactic orientation is about 20% of orientation at optimal calcium concentration of 10?4 M at 9 V. The higher the external calcium concentration is, the lower is the voltage necessary for optimal galvanotactic orientation. The calcium channel blockers diltiazem and nifedipine likewise inhibit galvanotaxis of Chlamydomonas very specifically without impairing motility. Verapamil is effective, but also inhibits motility by causing detachment or shortening of the flagella. Nevertheless, inhibition of galvanotaxis by verapamil is not the only result of decreased motility, because the galvanotactic orientation is impaired to a greater extent than motility. The effectiveness of the three blockers tested in inhibiting galvanotaxis depends on the concentration and on the voltage applied. At 10?5 M, verapamil causes maximal inhibition of galvanotaxis at 9 V. At increasing concentrations up to 10?4 M, diltiazem inhibits galvanotaxis more strongly than the other blockers. If the voltage is varied at a constant blocker concentration of 2 × 10?5 M, nifedipine causes maximal inhibition at 3 V–6 V, diltiazem at 9 V and verapamil above 12 V.  相似文献   

20.
14α-Hydroxymethyl-5α-cholest-7-en-3β-ol (I) and 14α-hydroxymethyl-5α-cholest-6-en-3β-ol (II) have been prepared by chemical synthesis from 3β-acetoxy-7α,32-epoxy-14α-methyl-5α-cholestane. Compound I, previously shown to be efficiently convertible to cholesterol upon incubation with rat liver homogenate preparations, has been found to be a potent inhibitor of sterol synthesis in animal cells in culture. Compound I caused a 50% reduction of the levels of HMG-CoA reductase activity in cultures of L cells and fetal liver cells at concentrations of 3 × 10?6 M and 8 × 10?6 M, respectively. Compound II, the Δ6-analogue of I, caused a 50% suppression of the enzyme activity in the two cell types at even lower concentrations, 5 × 10?7 M and 2 × 10?6 M, respectively. Concentrations of I and II required to specifically inhibit sterol synthesis from acetate were similar to those required to suppress the levels of HMG-CoA reductase activity.  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