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 31 毫秒
1.
Static light scattering of high amylopectin waxy maize starch gently dispersed in 90% dimethyl sulfoxide–water yielded a weight average molecular weight Mw and radius of gyration Rg of 560×106 g/mol and 342 nm, respectively. To obtain an independent hydrodynamic characterization of these solutions, we measured the sedimentation coefficient for the main component in an analytical ultracentrifuge. The value of s0, the infinite dilution sedimentation coefficient, was 199 S. The translational diffusion coefficient D0 in very dilute solutions was measured by dynamic light scattering at 90° and found to be 2.33×10−9 cm2/s. An effective hydrodynamic radius Rh was calculated from this diffusion constant using the Stokes–Einstein equation and found to be 348 nm. The structure-related parameter ρ=Rg/Rh was calculated to be 0.98. The weight average molecular weight calculated from the Svedberg equation using the values measured for s0 and D0 was 593×106 g/mol. This result is in reasonable agreement with the light scattering results. As light scattering results are subject to experimental errors due to the possibility of dust contamination, the presence of microgel or aggregates, and the questionable applicability of light scattering theory to interpret results for macromolecular sizes approaching the wave length of light used as a source for scattering, it is advisable to have corroborating hydrodynamic data when possible to further validate light scattering results in this very high molecular weight range.  相似文献   

2.
The structure of the 7S globulin from Phaseoulus vulgaris L in dilatue solutions has been studied by small angle X-ray scattering (SAXS), by quasi-elastic light scattering (Q ELS), by circular dichroism spectroscopy (c.d.), and by precise density measurements. The molar mass, the radius of gyration, the volume, the maximum dimension and the diffusion coefficient were determined as M = 1.45 × 105 g mol−1, RG = 4.05 nm, V = 300- nm3, L = 13.0 nm and D20,w0 = 4.5 × 10−7 cm2 s−1, respectively. The molecule has an asymmetrical shape with the dimensions 12.5 × 12.5 × 3.75 nm. The secondary structure of the 7S globulin is characterized by a small portion of -helical structure (14%) and a marked content of β-structure (18%).  相似文献   

3.
Studies on the molecular chain morphology of konjac glucomannan   总被引:2,自引:0,他引:2  
The chain geometry and parameters of konjac glucomannan were studied by using laser light scatter (LLS), gel permeation chromatography (GPC) and viscosimetry. The weight-average molecular weight (Mw), root-mean-square ratio of gyration (S21/2), second viral coefficient (A2) and polydispersity index (Mw/Mn) were 1.036×106, 105±0.9 nm, (−1.587±0.283)×10−3  mol ml g−2 and 1.015±0.003 respectively. Mark-Houwink equation was established as , and the molecular chain parameters were as follows: ML=982.82 nm−1, q=27.93 nm, d=0.74 nm, h=0.26 nm, L=1054.11 nm. To confirm the above results, konjac glucomannan was observed by using atomic force microscopy (AFM) and transmission electron microscope (TEM). The physical image showed directly that the konjac glucomannan molecule was an extending semi-flexible linear chain without branches, and than the molecular dimension also conformed to the parameters above. Therefore the image of molecular chain geometry confirmed the deduction drawn by Mark-Houwink equation and molecular chain parameters magnificently.  相似文献   

4.
C. Görlach  M. Wahl 《Peptides》1996,17(8):1373-1378
Ring segments of rat middle cerebral artery (MCA) were prepared for measurement of isometric force and precontracted with 10−4 M uridine triphosphate (UTP). Concentration-effect curves (CEC) were constructed for bradykinin (BK, 10−8–10−5 M) in segments with functionally intect (E+) or denuded (E−) endothelium. E− segments did not dilate to BK. The BK receptor was characterized by application of specific B1 or B2 antagonists [des-Arg9-Leu8] BK (10−5 M) and [ -Arg0-Hyp3-Thi5- -Tic7-Oic8] BK (HOE140,3 × 10−7 M), respectively, or B1 agonist [des-Arg9] BK (10−8–10−4 M). Involvement of nitric oxide (NO) was tested with NG-nitro- -arginine (LNNA, 10−4 M). BK induced concentration-dependent relaxation with a maximal effect (Emax) of 40.86 ± 1.50% at 10−6 M and a pD2 (−log10 EC50) of 6.818 ± 0.044. This relaxation could be prevented with HOE140 or LNNA, but was not influenced by [des-Arg9-Leu8] BK. [des-Arg9] BK did not induce any effect. These results demonstrate that BK induced relaxation via endothelial B2 receptors and release of NO in isolated rat MCA.  相似文献   

5.
The interaction between sialosyl cholesterol (- or neuraminyl cholesterol, - or β-SC) and the plasma membrane of astrocytes was investigated by the use of 14C-labeled - or β-SC. Both - and β-SC were dose-dependently and time-dependently bound to rat astrocytes. The Scatchard plot analyses showed that rat astrocytes bound apparently 9.69 × 109 molecules of both -SC/cell (apparent Kd = 2.29 × 10−5 M) and β-SC/cell (apparent Kd = 5.39 × 10−5 M) at 37°C. Both the binding of -SC to astrocytes and the subsequent inhibition of DNA synthesis were decreased at the low temperature (4°C), and also suppressed by serum proteins including albumin. One molecule of bovine serum albumin (BSA) bound 2.3 molecules of -SC with the slightly lower Kd-value (8.03 × 10−6 M) than that for the binding site on astrocytes. BSA not only suppressed the -SC-binding to astrocytes but also increased its release from the cells to the culture media. Gangliosides such as GM1 and GM3 unaffected the -SC-binding, promoted the small release of -SC from the cell surface, and inhibited the morphological changes of astrocytes induced by -SC. The mechanism of -SC-binding to cultured astrocytes with reference to the effects of serum or gangliosides is discussed.  相似文献   

6.
Dextran is currently being considered as a carrier biopolymer system for drug targeting. Di-iodotyrosine (DIT), and its radioactive derivative are useful markers for identifying the in-vivo route of these substrates. The performance of these substances in vivo is closely linked to their physico-chemical properties in solution. These properties have been investigated on a series of gel permeation chromatography fractionated DIT-dextrans using a combination of sedimentation, viscometric and laser-light scattering techniques. Weight average molecular weights, Mw, from sedimentation equilibrium and light scattering range from 0·0195 to 0·054 × 106. Double logarithmic representations of Mw versus sedimentation coefficient, so20.w, and intrinsic viscosity [η] suggest that these labelled dextrans have a more compact conformation in solution than do their unlabelled analogues — which behave as random coils. The implications of this for drug delivery are indicated.  相似文献   

7.
The hydrodynamic characteristics of the polysaccharide pullulan (polymaltotriose) in water have been investigated and its molecular characteristics have been determined. Experimental values varied over the following ranges: velocity sedimentation coefficient (S): 0.9 < S < 11.2, translational diffusion coefficient (107 cm2 s−1): 1.1 < D < 14.7 and intrinsic viscosity (cm3 g−1): 6.7 < [η] < 164, which corresponds to a change in molecular weight (× 103) in the range 3.9 < MSD < 644. On the basis of analysis of the literature and our experimental data, excluded volume effects have been shown to have a prevailing influence on the chain length of these polysaccharides. The equilibrium rigidity and hydrodynamic chain diameter of pullulan were evaluated on the basis of the theory of hydrodynamic properties of a wormlike necklace, taking into account excluded volume effects. At low M (< 30 × 103) the translation friction data (in contrast to viscometric data) cannot be described in the framework of the theory of linear molecules.  相似文献   

8.
The aim of our study was to determine whether a meal modifies the antisecretory response induced by PYY and the structural requirements to elicit antisecretory effects of analogue PYY(22–36) for potential antidiarrhea therapy. The variations in short-circuit current (Isc) due to the modification of ionic transport across the rat intestine were assessed in vitro, using Ussing chambers. In fasted rats, PYY induced a dose- and time-dependent reduction in Isc, with a sensitivity threshold at 5 × 10−11 M (ΔIsc −2 ± 0.5 μA/cm2). The reduction was maximal at 10−7 M (Isc −23 ± 2 μA/cm2), and the concentration producing half-maximal inhibition was 10−9 M. At 10−7 M, reduction of Isc by PYY reached 90% of response to 5 × 10−5 M bumetanide. The PYY effect was partly reversed by 10−5 M forskolin (Isc +13.43 ± 2.91 μA/h·cm2, p < 0.05) or 10−3 M dibutyryl adenosine 3′,5′ cyclic monophosphate (Isc +12 ± 1.69 μA/cm2, p < 0.05). Naloxone and tetrodotoxin did not alter the effect of PYY. In addition, PYY and its analogue P915 reduced net chloride ion secretion to 2.85 and 2.29 μEq/cm2 (p < 0.05), respectively. The antisecretory effect of PYY was accompanied by dose- and time-dependent desensitization when jejunum was prestimulated by a lower dose of peptide. The antisecretory potencies exhibited by PYY analogues required both a C-terminal fragment (22–36) and an aromatic amino acid residue (Trp or Phe) at position 27. At 10−7 M the biological activity of PYY was lower in fed than fasted rats (p < 0.001). Our results confirm the antisecretory effect of PYY, but show that the fed period is accompanied by desensitization, similar to the transient desensitization observed in the fasted period with cumulative doses. This suggests that PYY may act as a physiological mediator that reduces intestinal secretion.  相似文献   

9.
A method for determination of carboxymethyl cellulose (CMC) molecular weight (MW) and chemical heterogeneity (degree of oxidation (DO)) using a bi-detector HPSEC (UV-detector online with refractometer) has been developed. It has been found that the use of 0.5 N NaOH or 0.4 M acetate buffer as the eluent ensures CMC separation according to MW. It has been revealed that the universal calibration for the polyelectrolyte CMC and the neutral polymer dextran is valid under the conditions applied. The Mark–Houwink equations for CMC in 0.5 N NaOH and 0.4 M acetate buffer have been estimated to be [η]=5.37×10−4 MW0.73 and [η] =2.24×10−4 MW0.83 dl g−1, respectively. The equation log K=1.64−4.00 ml g−1 for CMC has been estimated. An approach for determining DO from adsorption at 290 or 313 nm has been developed.  相似文献   

10.
Oxygenation of [CuII(fla)(idpa)]ClO4 (fla=flavonolate; IDPA=3,3′-iminobis(N,N-dimethylpropylamine)) in dimethylformamide gives [CuII(idpa)(O-bs)]ClO4 (O-bs=O-benzoylsalicylate) and CO. The oxygenolysis of [CuII(fla)(idpa)]ClO4 in DMF was followed by electronic spectroscopy and the rate law −d[{CuII(fla)(idpa)}ClO4]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2] was obtained. The rate constant, activation enthalpy and entropy at 373 K are kobs=6.13±0.16×10−3 M−1 s−1, ΔH=64±5 kJ mol−1, ΔS=−120±13 J mol−1 K−1, respectively. The reaction fits a Hammett linear free energy relationship and a higher electron density on copper gives faster oxygenation rates. The complex [CuII(fla)(idpa)]ClO4 has also been found to be a selective catalyst for the oxygenation of flavonol to the corresponding O-benzoylsalicylic acid and CO. The kinetics of the oxygenolysis in DMF was followed by electronic spectroscopy and the following rate law was obtained: −d[flaH]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2]. The rate constant, activation enthalpy and entropy at 403 K are kobs=4.22±0.15×10−2 M−1 s−1, ΔH=71±6 kJ mol−1, ΔS=−97±15 J mol−1 K−1, respectively.  相似文献   

11.
The growth of the freshwater microalga Scenedesmus obliquus was studied at 30°C in a mineral culture medium with phosphorus concentrations of between 0 and 372 μ . The values for the specific growth rates, between and , fitted a semistructured substrate-limitation model with μm1 = 0·0466 h−1, μm2 = 0·0256 h−1 and . The specific uptake rate of phosphorus reached a maximum value of qSm1 = 658·01 × 10−4 μmol P mg−1 biomass h−1.  相似文献   

12.
Tapani Kuronen  Nils Ellfolk 《BBA》1972,275(3):308-318
A procedure has been developed for purification of the cytochrome oxidase from Pseudomonas aeruginosa (EC 1.9.3.2) using DEAE- and CM-cellulose chromatography, gel filtration and crystallization. The final preparation was found to be homogeneous according to ultracentrifugal and disc electrophoretic criteria. The crystalline preparation also exhibited nitrite reductase activity. The spectrum of the enzyme characterizes it as cytochrome cd. At 280 nm E1 %1 cm was 18.5 after dry weight analysis.

The molecular weight of the cytochrome oxidase was calculated to be 119000 based on a sedimentation coefficient s° 20,w = 7.36 S, diffusion coefficient D 20,w = 5.36×10−7 cm2×s−1 and partial specific volume of 0.72 ml/g. The iron content of the enzyme (0.166 %) indicates that this entity contains four iron atoms per molecule. Succinylation of the enzyme produced two probably identical subunits containing both hemes c and d, having a sedimentation coefficient s° 20,w = 4.30 S and an approximate molecular weight of 65000. In dodecylsulphate-acrylamide gel electrophoresis the cytochrome oxidase also dissociates into two subunits with molecular weight of 63000.  相似文献   


13.
Mono- and di-manganese inclusion compounds 1 and 2 are reported. Two mono-manganese molecules Mn(bpy)2(NO3)2 (bpy=2,2′-bipyridine) and [Mn(bpy)2(NO3)(H2O)]·NO3 coexist in the mole ratio of 1:1 in the structure of 1, while two di-manganese molecules [Mn2O(bpy)2(phtha)2(H2O)2]·(NO3)2 (phtha=phthalate) and [Mn2O(bpy)2(phtha)2(NO3)(H2O)]·NO3 in the structure of 2. Refluxing Mn(NO3)2/bpy/phthalic acid reaction mixtures in CH3CN leads to the isolation of 1, further concentration of the reaction solution in raising temperature results in 2. The Mn1 and Mn2 units in the inclusion compounds 1 and 2 are similar to other reported Mn1 and Mn2 analogs, respectively. The Jahn–Teller distortion was observed to give rise to the elongation along the Oterminal---Mn---Ocarboxyl axes for all the four Mn(III) sites in 2, leading to unexpected longer Mn(III)---Oaqua than Mn(II)---Oaqua in 1. Extensive hydrogen bonding interactions among H2O, NO3 − and COOH were observed in the two inclusion compounds. Cyclic voltammetry of 2 in DMF displays two quasi-reversible redox couples at +0.10/+0.22 and −0.43/−0.36 V assigned to the Mn(III)Mn(IV)/2Mn(III) and 2Mn(III)/Mn(III)Mn(II), respectively. Variable temperature magnetic susceptibilities of 1 and 2 were measured. The data were fit to a model including axial zero-field splitting term and a good fit was found with D=1.77 cm−1, g=1.98 and F=1.48×10−5 for 1. For 2, the least-squares fitting of the experimental data led to J=2.37 cm−1, g=2.02 and D=0.75 cm−1 with R=1.45×10−3.  相似文献   

14.
To study mechanisms of aromatase inhibition in brain cells, a highly effective non-steroidal aromatase inhibitor (Fadrozole; 4-[5,6,7,8-tetra-hydroimidazo-(1,5-a)-pyridin-5-yl] benzonitrile HCl; CGS 16949A) was compared with endogenous C-19 steroids, known to be formed in the preoptic area, which inhibit oestrogen formation. Using a sensitive in vitro tritiated water assay for aromatase activity in avian (dove) preoptic tissue, the order of potency, with testosterone as substrate was: Fadrozole (Ki < 1 × 10−9 M) > 4-androstenedione 5-androstanedione > 5-dihydrotestosterone (Ki = 6 × 10−8 M) > 5β-androstanedione > 5β-dihydrotestosterone (Ki = 3.5 × 10−7 M) > 5-androstane-3, 17β-diol (Ki = 5 × 10−6 M) > 5β-androstane-3β,17β-diol. Five other steroids, 5β-androstane-3,17β-diol, 5-androstane-3β,17β-diol, progesterone, oestradiol and oestrone, showed no inhibition at 10−4 M. The kinetics indicate that endogenous C-19 steroids show similar competitive inhibition of the aromatase as Fadrozole. Mouse (BALB/c) preoptic aromatase was also inhibited by Fadrozole. We conclude that endogenous C-19 metabolites of testosterone are effective inhibitors of the brain aromatase, and suggest that they bind competitively at the same active site as Fadrozole.  相似文献   

15.
Aspergillus niger catalase has been characterized by a variety of physical techniques including gel filtration, sedimentation rate and equilibrium methods and photon correlation spectroscopy. The catalase has a sedimentation coefficient (S200) of 14.2 ± 0.08 S and diffusion coefficient (D200) of 4.14 ± 0.35 × 10−7 cm2 s−1. The average molecular weight of the catalase from all available data including current sedimentation equilibrium measurements and two previous literature values is 345 000. The frictional ratio of the molecule assuming a hydration parameter similar to that of bovine liver catalase (.3 g H2O g−1) is 1.103, suggesting that Aspergillus niger catalase has an asymmetric structure with an axial ratio of approximately 3 (the Stokes radius is 5.83 ± 0.49 nm). The titration curve and amino acid analysis indicate that in the native conformation only 23% of the ionizable amino acid residues are titratable between pH 3 and 10.5. Denaturation with sodium n-dodecylsulphate increases the number of titratable groups to 46%. The ratio of anionic to cationic amino acid residues in Aspergillus niger catalase is 2.46 and the isoelectric point is 6.5. The optimum pH for catalytic activity is approximately 7.  相似文献   

16.
Palythoa psammophilia Walsh & Bowers has a well coordinated, stereotyped feeding response, the culminating step of which is ingestion; this may be elicited by the synergistic effect of the tripeptide glutathione and the -imino acid, proline. Either activator acting separately causes responses only at high concentrations (above 10−5 M for glutathione; above 10−4 M for proline) in a reduced number of animals and at a low rate (5.00 ± 1.73 min in 5 × 10−3 M solutions of glutathione; 11.10±3.74 min in 5 × 10−3 M solutions of proline). Highest percentages of response were obtained in combinations where glutathione was at a concentration of 5 × 10−3 M and proline at 5 × 10−4 M or in combinations of glutathione at concentrations 5 × 10−6 M and proline at 5 × 10−5 M. The speed of ingestion is considerably enhanced when these activators are combined (1.17±1.18 min).  相似文献   

17.
We have investigated the genotoxic effects of 1-(2-hydroxyethyl)-1-nitrosourea (HENU). We have chosen this agent because of its demonstrated ability to produce N7-(2-hydroxyethyl) guanine (N7-HOEtG) and O6-(2-hydroxyethyl) 2′-deoxyguanosine (O6-HOEtdG); two of the DNA alkylation products produced by 1,3-bis (2-chloroethyl)-1-nitrosourea (BCNU). For these studies, we have used the Big Blue Rat-2 cell line that contains a lambda/lacI shuttle vector. Treatment of these cells with HENU produced a dose dependent increase in the levels of N7-HOEtG and O6-HOEtdG as quantified by HPLC with electrochemical detection. Treatment of Big Blue Rat-2 cells with either 0, 1 or 5 mM HENU resulted in mutation frequencies of 7.2±2.2×10−5, 45.2±2.9×10−5 and 120.3±24.4×10−5, respectively. Comparison of the mutation frequencies demonstrates that 1 and 5 mM HENU treatments have increased the mutation frequency by 6- and 16-fold, respectively. This increase in mutation frequency was statistically significant (P<0.001). Sequence analysis of HENU-induced mutations have revealed primarily G:C→A:T transitions (52%) and a significant number of A:T→T:A transversions (16%). We propose that the observed G:C→A:T transitions are produced by the DNA alkylation product O6-HOEtdG. These results suggest that the formation of O6-HOEtdG by BCNU treatment contributes to its observed mutagenic properties.  相似文献   

18.
N-Methyl-N′-nitro-N-nitrosoguanidine (MNNG) reacts with 12 nucleophilic sites in DNA to induce a variety of lesions, but O6-methylguanine (O6-MeG) and O4-methylthymine are the most effective premutagenic lesions produced, mispairing with thymine and guanine, respectively. O6-MeG is repaired by O6-alkylguanine-DNA alkyltransferase (AGT), which removes the methyl group from the O6 position and transfers it to itself, rendering the transferase inactive. When diploid human fibroblasts were exposed to 25 μM, O6-benzylguanine (O6-BzG) in the medium for 3 h, their level of AGT activity was dramatically reduced, to a level of at most 1.6% of the control. Populations of cells pretreated with this level of O6-BzG for 2 h or not pretreated, were exposed to MNNG at a concentration of 2, 4 or 6 μM in the presence or absence of O6-BzG and assayed for survival of colony-forming ability and the frequency of 6-thioguanine-resistant cells (mutations induced in the HPRT gene). O6-BzG (25 μM) was also present in the appropriate half of the cells during the 24 h immediately follwing exposure to MNNG. This 27-h exposure to O6-BzG alone had no cytotoxic or mutagenic effect on the cells but significantly increased the cytotoxicity and mutagenecity of MNNG, increasing the mutant frequency to that found previously in human cells constitutively devoid of AGT activity. At doses of 2 μM and 4 μM MNNG, the mutant frequency observed with the AGT-depleted cells was 120 × 10−6 and 240 × 10−6, respectively; in the cells with abundant AGT activity, these values were 10 × 10−6 and 20 × 10−6, respectively. DNA-sequence analysis of the coding region of the HPRT gene in 36 independent mutants obtained from MNNG-treated AGT-depleted populations and 36 from the control populations showed that even though AGT repair lowered the frequency of mutants by more than 90%, it did not affect the kinds of mutations induced by MNNG nor the strand distribution of the premutagenic guanine lesions. In mutants from the AGT-depleted cells, there were 26 base substitutions and 13 putative splice site mutations; in the control, there were 25 base substitutions and 11 splice site mutations. All but two substitutions involved G · C with 92% being G · C → A · T. In both sets, of the premutagenic lesions were located in the nontranscribed strand. Many ‘hot spots’ were seen, and there was evidence that AGT repaired more lesions from the 5′ half of the gene than from the 3′ half.  相似文献   

19.
We measured the toxicity and mutagenicity induced in human diploid lymphoblasts by various radiation doses of X-rays and two internal emitters. [125I]iododeoxyuridine ([125I]dUrd) and [3H]thymidine ([3H]TdR), incorporated into cellular DNA. [125I]dUrd was more effective than [3H]TdR at killing cells and producing mutations to 6-thioguanine resistance (6TGR). No ouabain-resistant mutants were induced by any of these agents. Expressing dose as total disintegrations per cell (dpc), the D0 for cell killing for [125I]dUrd was 28 dpc and for [3H]TdR was 385 dpc. The D0 for X-rays was 48 rad at 37°C. The slopes of the mutation curves were approximately 75 × 10−8 6TGR mutants per cell per disintegration for [125I]dUrd and 2 × 10−8 for [3H]TdR. X-Rays induced 8 × 10−8 6TGR mutants per cell per rad. Normalizing for survival, [125I]dUrd remained much more mutagenic at low doses (high survival levels) than the other two agents. Treatment of the cells at either 37°C or while frozen at −70°C yielded no difference in cytotoxicity or mutation for [125I]dUrd or [3H]TdR, whereas X-rays were 6 times less effective in killing cells at −70°C.

Assuming that incorporation was random throughout the genome, the mutagenic efficiencies of the radionuclides could be calculated by dividing the mutation rate by the level of incorporation. If the effective target size of the 6TGR locus is 1000–3000 base pairs, then the mutagenic efficiency of [125I]dUrd is 1.0–3.0 and of [3H]TdR is 0.02–0.06 total genomic mutations per cell per disintegration. 125I disintegrations are known to produce localized DNA double-strand breaks. If these breaks are potentially lethal lesions, they must be repaired, since the mean lethal dose (D0) was 28 dpc. The observations that a single dpc has a high probability of producing a mutation (mutagenic efficiency 1.0–3.0) would suggest, however, that this repair is extremely error-prone. If the breaks need not be repaired to permit survival, then lethal lesions are a subset of or are completely different from mutagenic lesions.  相似文献   


20.
It is well recognized that estradiol (E2) is one of the most important hormones supporting the growth and evolution of breast cancer. Consequently, to block this hormone before it enters the cancer cell or in the cell itself, has been one of the main targets in recent years. In the present study we explored the effect of the progestin, nomegestrol acetate, on the estrone sulfatase and 17β-hydroxy-steroid dehydrogenase (17β-HSD) activities of MCF-7 and T-47D human breast cancer cells. Using physiological doses of estrone sulfate (E1S: 5 × 10−9 M), nomegestrol acetate blocked very significantly the conversion of E1S to E2. In the MCF-7 cells, using concentrations of 5 × 10−6 M and 5 × 10−5 M of nomegestrol acetate, the decrease of E1S to E2 was, respectively, −43% and −77%. The values were, respectively, −60% and −71% for the T-47D cells. Using E1S at 2 × 10−6 M and nomegestrol acetate at 10−5 M, a direct inhibitory effect on the enzyme of −36% and −18% was obtained with the cell homogenate of the MCF-7 and T-47D cells, respectively. In another series of studies, it was observed that after 24 h incubation of a physiological concentration of estrone (E1: 5 × 10−9 M) this estrogen is converted in a great proportion to E2. Nomegestrol acetate inhibits this transformation by −35% and −85% at 5 × 10−7 M and 5 × 10−5 M, respectively in T-47D cells; whereas in the MCF-7 cells the inhibitory effect is only significant, −48%, at 5 × 10−5 M concentration of nomegestrol acetate. It is concluded that nomegestrol acetate in the hormone-dependent MCF-7 and T-47D breast cancer cells significantly inhibits the estrone sulfatase and 17β-HSD activities which converts E1S to the biologically active estrogen estradiol. This inhibition provoked by this progestin on the enzymes involved in the biosynthesis of E2 can open new clinical possibilities in breast cancer therapy.  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