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.
Isolated rat hepatocytes posses a saturable glucocorticoid uptake system with high affinity (Kd value = 2.8 ± 0.7 × 10−8 M; 318,000 ± 80,000 binding sites per cell; 317 fmol/mg protein). The initial rates of uptake decrease by about 30–40% if the cells are incubated simultaneously with [3H]corticosterone and either SH-reagents (N-ethylmaleimide and p-chloromercuriphenylsulphonate, 1 mM), metabolic inhibitors (2,4-dinitrophenol, 1 mM; and antimycin, 0.1 mM) or the Na+/K+-ATPase-inhibitors, ouabain and quercetine. These Na+/K+-ATPase-blockers exert half-maximal inhibition at 3 × 10−7 and 3 × 10−6 M, respectively. A slight increase in K+ concentration and a corresponding decrease in Na+ in the medium leads to a significant reduction in the initial uptake rate. The uptake system from the rat hepatocytes shows a clear steroid specificity, being different from the intracellular receptor. Corticosterone and progesterone are the strongest competitors, cortisol, 5- and 5β-dihydrocorticosterone, 11-deoxycorticosterone, cortisone and testosterone have an intermediate effect and only weak competition is exerted by dexamethasone and by the mineralocorticoid, aldosterone. Estradiol and estrone sulphate as well as the synthetic glucocorticoid triamcinolone acetonide are unable to inhibit initial corticosterone uptake.  相似文献   

2.
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   

3.
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   

4.
Aqueous solutions of fractions of an extracellular linear mannan formed by Rhodotorula rubra yeast have been investigated by hydrodynamic methods (high-speed sedimentation, translation isothermic diffusion and viscometry). The molecular weight was determined according to Svedberg ( ) and the polydispersity parameters of the initial sample were also determined (Mw/Mn = 1·20 and Mz/Mw = 1·21). Relationships between the molecular weight (M) and so, Do and [η] in the range were: [η] = 2·33 × 10−2 M0.75, Do = 1·65 × 10−4 M0·58, so = 2·24 × 10−15 M0·43. The equilibrium rigidity and hydrodynamic diameter of chains representing mannan molecules were evaluated.  相似文献   

5.
High-pressure liquid-chromatography and microcalorimetry have been used to determine equilibrium constants and enthalpies of reaction for the disproportionation reaction of adenosine 5′-diphosphate (ADP) to adenosine 5′-triphosphate (ATP) andadenosine 5′-monophosphate (AMP). Adenylate kinase was used to catalyze this reaction. The measurements were carried out over the temperature range 286 to 311 K, at ionic strengths varying from 0.06 to 0.33 mol kg−1, over the pH range 6.04 to 8.87, and over the pMg range 2.22 to 7.16, where pMg = -log a(Mg2+). The equilibrium model developed by Goldberg and Tewari (see the previous paper in this issue) was used for the analysis of the measurements. Thus, for the reference reaction: 2 ADp3− (ao) AMp2− (ao)+ ATp (ao), K° = 0.225 ± 0.010, ΔG° = 3.70 +- 0.11 kJ mol −1, ΔH° = −1.5 ± 1. 5 kJ mol −1, °S ° = −17 ± 5 J mol−1 K−1, and ACPp°≈ = −46 J mo1l−1 K−1 at 298.15 K and 0.1 MPa. These results and the thermodynamic parameters for the auxiliary equilibria in solution have been used to model the thermodynamics of the disproportionation reaction over a wide range of temperature, pH, ionic strength, and magnesium ion morality. Under approximately physiological conditions (311.15 K, pH 6.94, [Mg2+] = 1.35 × 10−3 mol kg−1, and I = 0.23 mol kg−1) the apparent equilibrium constant (KA′ = m(ΣAMP)m(ΣATP)/[ m(ΣADP)]2) for the overall disproportionation reaction is equal to 0.93 ± 0.02. Thermodynamic data on the disproportionation reaction and literature values for this apparent equilibrium constant in human red blood cells are used to calculate a morality of 1.94 × 10−4 mol kg−1 for free magnesium ion in human red blood cells. The results are also discussed in relation to thermochemical cycles and compared with data on the hydrolysis of the guanosine phosphates.  相似文献   

6.
1. The alteration of the Ca2+ requirements of the ATPase activity of fibrils from rabbits and crabs at varying ionic strength, pH and concentration of MgATP (i.e. MgATP2− + MgHATP) was investigated.

2. Under physiological conditions, it was found that the ATPase activity of rabbit and crab fibrils after an initial increase decreased steeply when the Ca2+ concentration is raised above 1×10−4 M. This is a primary effect of the over-optimal Ca2+ concentration and not a secondary one caused by the influence of accompanying ions.

3. The Ca2+ requirements for ATP splitting by rabbit fibrils remain constant at an ionic strength from 0.1 to 0.2 and for a MgATP concentration in the range from 0.5 to 10 mM. At I = 0.05 it is about 5 times smaller than at 0.1. When the pH is decreased from 8 to 7, the Ca2+ requirements are increased some 10 times but only 3 times when the pH is varied between 7 and 6.

4. In crab fibrils, there is no alteration of the Ca2+ requirements when the ionic strength is varied between 0.05 and 0.2, but a reduction of the pH from 8.0 to 6.0 raises the Ca2+ requirements for half activation and for threshold by a factor of 10. Changing the MgATP concentration increases the Ca2+ requirements only in the range from 1 to 5 mM, while the concentration required in 0.5 mM is identical with that at 1 mM, and 10 mM corresponds to 5 mM.

5. It can be deduced from the experimental results that at a pH above 6.0 maximal activation is always obtained if the Ca2+ concentration is 5×10−5 M. By contrast, relaxation is only achieved when the Ca2+ concentration is below 1×10−7 M for pH 7.0 and I > 0.1 or below 1×10−8 for pH > 7.0 or I < 0.1.

6. To achieve complete relaxation, an ethyleneglycoldiaminotetraacetate (EGTA) concentration of 1 mM is sufficient, even when there is a large degree of contamination by Ca2+ as long as the pH stays above 6.5.  相似文献   


7.
Guar gum has been modified by graft copolymerization with acrylic acid in aqueous medium using vanadium (V)–mercaptosuccinic acid redox system. The optimum reaction conditions affording maximum grafting ratio, efficiency, add on and conversion have been determined. The grafting parameters have been found to increase with increase in vanadium (V) concentration upto 1.0 × 10−2 mol dm−3, but these parameters decrease on further increasing the vanadium (V) concentration. On increasing the mercaptosuccinic acid concentration from 1.0 × 10−2 to 4.0 × 10−2 mol dm−3 grafting ratio, efficiency and add on increase up to 2.0 × 10−2 mol dm−3 but decrease with further increase in mercaptosuccinic acid concentration. On varying the acrylic acid concentration from 5.0 × 10−2 to 30.0 × 10−2 mol dm−3, maximum grafting ratio, efficiency and add on have been obtained at 20.0 × 10−2 mol dm−3. The grafting ratio, add on and conversion increase, on increasing the H+ ion concentration from 1.5 × 10−1 to 6.0 × 10−1 mol dm−3. On increasing the guar gum concentration the grafting parameters increase. The grafting ratio, add on and conversion have been found to increase with time period while efficiency started decreasing after 120 min. It has been observed that %G increases on increasing the temperature up to 35 °C. The graft copolymer has been characterized by IR spectroscopy and thermogravimetric analysis.  相似文献   

8.
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   

9.
Differential UV spectroscopy and thermal denaturation were used to study the Mg2+ ion effect on the conformational equilibrium in poly A · 2 poly U (A2U) and poly A · poly U (AU) solutions at low (0.01 M Na+) and high (0.1 M Na+) ionic strengths. Four complete phase diagrams were obtained for Mg2+–polynucleotide complexes in ranges of temperatures 20–96 °C and concentrations (10−5–10−2) M Mg2+. Three of them have a ‘critical’ point at which the type of the conformational transition changes. The value of the ‘critical’ concentration ([Mgt2+]cr=(4.5±1.0)×10−5 M) is nearly independent of the initial conformation of polynucleotides (AU, A2U) and of Na+ contents in the solution. Such a value is observed for Ni2+ ions too. The phase diagram of the (A2U+Mg2+) complex with 0.01 M Na+ has no ‘critical’ point: temperatures of (3→2) and (2→1) transitions increase in the whole Mg2+ range. In (AU+Mg2+) phase diagram at 0.01 M Na+ the temperature interval in which triple helices are formed and destroyed is several times larger than at 0.1 M Na+. Using the ligand theory, a qualitative thermodynamic analysis of the phase diagrams was performed.  相似文献   

10.
1. Fluoride is a mixed-type inhibitor of the cytochrome c oxidase activity with a Ki for the free enzyme of 10 mM and a Ki for the cytochrome c-complexed enzyme of 35 mM.

2. Fluoride shifts the γ-band of the enzyme from 423 to 421 nm and the -band from 597 to 598 nm. The difference spectrum (oxidized enzyme in the presence of fluoride minus oxidized enzyme) has peaks at 400, 453, 482, 605 and 638 nm and troughs at 430, 520, 552 and 674 nm. The changes in absorbance are small (about 3% at absorbance maxima) with respect to those of other hemoproteins.

3. On addition of fluoride to isolated cytochrome c oxidase 3 reactions can be distinguished: (I) a bimolecular binding reaction (Kon = 4 M−1 · s−1 and koff = 2.9 · 10−2s−1 at 25 °C, pH 7.4) contributing at 638 nm and 430 nm; (II) a first-order reaction (k = 2.4 · 10−2) s−1 at 22 °C, pH 7.2) visible mainly at 430 nm and (III) a very slow reaction with a half-time in the order of 10 min.

4. The spectroscopic dissociation constants for the fluoride binding, determined from Hill plots using the absorbance changes at 638 and 430 nm, are similar (7 and 10 mM, respectively, at 22 °C, pH 7.2).

5. A mechanism for the reaction is discussed in which the bimolecular binding reaction is followed by a conformational change of the enzyme-fluoride complex.  相似文献   


11.
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   

12.
We used a direct polymerase chain reaction (PCR) method for quantification of HPRT exons 2+3 deletions and t(14;18) translocations as a measure of illegitimate V(D)J recombination. We determined the baseline frequencies of these two mutations in mononuclear leukocyte DNA from the umbilical cord blood of newborns and from the peripheral blood of adults. In an initial group of 21 newborns, no t(14;18) translocations were detected (<0.049×10−7). The frequency of HPRT exons 2+3 deletions was 0.10×10−7 per mononuclear leukocyte, lower than expected based on the T-cell proportion of this cell fraction (55%–70%) and previous results using the T-cell cloning assay (2–3×10−7 per clonable T-cell). Phytohemagglutinin (PHA), as used in the T-cell cloning assay, was examined for its effect on the frequencies of these mutation events in mononuclear leukocytes from an additional 11 newborns and from 12 adults. There was no significant effect of PHA on t(14;18) translocations which were rare among the newborns (1 detected among 2.7×108 leukocytes analyzed), and which occurred at frequencies from <1×10−7 (undetected) to 1.6×10−4 among the adults. The extremely high frequencies of t(14;18)-bearing cells in three adults were due mainly to in vivo expansion of two to six clones. However, PHA appeared to stimulate a modest (although not significant) increase in the frequency of HPRT exons 2+3 deletions in the leukocytes of the newborns, from 0.07×10−7 to 0.23×10−7. We show that both the direct PCR assay and the T-cell cloning assay detect similar frequencies of HPRT exons 2+3 deletions when calculations are normalized to blood volume, indicating that the apparent discrepancy is probably due to the different population of cells used in the assays. This direct PCR assay may have utility in characterizing the effects of environmental genotoxic agents on this clinically important recombination mechanism.  相似文献   

13.
The enzymatic activity of mushroom tyrosinase was investigated using catechin as substrate in selected organic solvent media. The results showed that optimal tyrosinase activity was obtained at pH 6.2, 6.6, 6.0 and 6.2 in the organic solvent media of heptane, toluene, dichloromethane, and dichloroethane, respectively, and at a temperature between 25°C and 27.5°C. In addition, the kinetic studies showed that the Km values were 5.38, 1.03, 2.52 and 4.03 mM, for the tyrosinase-catechin biocatalysis in the reaction media of heptane, toluene, dichloromethane, and dichloroethane, respectively, while the corresponding Vmax values were 1.22×10−3, 0.33×10−3, 1.47×10−3 and 1.20×10−3 δA per μg protein per second, respectively. The use of acetone as co-solvent for the tyrosinase-catechin biocatalysis showed that acetone concentrations ranging from 5% to 30% (v/v) in the heptane reaction medium produced a decrease of 4.3% to 96.7% in tyrosinase activity. The results also indicated that the presence of 12.5% acetone in the reaction medium of dichloromethane, and 22.0% in those of toluene and dichloroethane produced a maximal increase of 42.6%, 92.1% and 71.8%, respectively, in tyrosinase activity. However, the overall findings indicated that additional increases in acetone concentration resulted in an inhibition of tyrosinase activity.  相似文献   

14.
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   

15.
Recently, we demonstrated that angiotensin-(1–7) (Ang-(1–7)) stimulates the Na+-ATPase activity through a losartan-sensitive angiotensin receptor, whereas bradykinin inhibits the enzyme activity through the B2 receptor [Regul. Pept. 91 (2000) 45; Pharmacol. Rev. 32 (1980) 1]. In the present paper, the effect of bradykinin (BK) on Ang-(1–7)-stimulated Na+-ATPase activity was evaluated. Preincubation of Na+-ATPase with 10−9 M Ang-(1–7) increases enzyme activity from 7.9±0.9 to 14.1±1.5 nmol Pi mg−1 min−1, corresponding to an increase of 79% (p<0.05). This effect is reverted by bradykinin in a dose-dependent manner (10−14–10−8 M), reaching maximal inhibitory effect at 10−9 M. Des-Arg9 bradykinin (DABK), an agonist of B1 receptor, at the concentrations of 10−9–10−7 M, does not mimic the BK inhibitory effect, and des-Arg9-[Leu8]-BK (DALBK), a B1 receptor antagonist, at the concentrations of 10−10–10−7 M, does not prevent the inhibitory effect of BK on Ang-(1–7)-stimulated enzyme. On the other hand, HOE 140, an antagonist of B2 receptor, abolishes the inhibitory effect of BK on the Ang-(1–7)-stimulated enzyme in a dose-dependent manner, reaching maximal effect at 10−7 M. Taken together, these data indicate that stimulation of B2 receptors by BK can counteract the stimulatory effect of Ang-(1–7) on the proximal tubule Na+-ATPase activity.  相似文献   

16.
Reactions of cis-diamminedichloroplatinum(II) with phosphonoformic acid (PFA), phosphonoacetic acid (PAA), and methylenediphosphonic acid (MDP) yield various phosphonatoplatinum(II) chelates which were characterized by phosphorus-31 NMR spectroscopy. The P-31 resonances for the chelates appear at 6–12 ppm downfield as compared to the uncomplexed ligands. All complexes exhibit monoprotic acidic behavior in the pH range 2–10. The chemical shift-pH profiles yielded acidity constants, 1.0 × 10−4, 1.5 × 10−4, and 1.3 × 10−6 M−1, for the PFA, PAA, and MDP chelates. In addition to the monomeric chelate, MDP formed a bridged diplatinum(II,II) complex when it reacted with cis-Pt (NH3)2(H2O)22+. The P-31 resonance for this binuclear complex appears at 22 ppm downfield from the unreacted ligand.

Rate data for the complexation reactions of the phosphonate ligands with the dichloroplatinum complex are consistent with a mechanism in which a monodentate complex is formed initially through rate-limiting aquation process of the platinum complex, followed by a rapid chelation. For the PFA and PAA complexes, initial binding sites are the carboxylato oxygens. Implications of the various binding modes of the phosphonates in relationship to their antiviral activities are discussed.  相似文献   


17.
Amperometric choline biosensors were fabricated by the covalent immobilization of an enzyme of choline oxidase (ChO) and a bi-enzyme of ChO/horseradish peroxidase (ChO/HRP) onto poly-5,2′:5′,2″-terthiophene-3′-carboxylic acid (poly-TTCA) modified electrodes (CPMEs). A sensor modified with ChO utilized the oxidation process of enzymatically generated H2O2 in a choline solution at +0.6 V. The other one modified with ChO/HRP utilized the reduction process of H2O2 in a choline solution at −0.2 V. Experimental parameters affecting the sensitivity of sensors, such as pH, applied potential, and temperature were optimized. A performance comparison of two sensors showed that one based on ChO/HRP/CPME had a linear range from 1.0×10−6 to 8.0×10−5 M and the other based on ChO/CPME from 1.0×10−6 to 5.0×10−5 M. The detection limits for choline employing ChO/HRP/CPME and ChO/CPME were determined to be about 1.0×10−7 and 4.0×10−7 M, respectively. The response time of sensors was less than 5 s. Sensors showed good selectivity to interfering species. The long-term storage stability of the sensor based on ChO/HRP/CPME was longer than that based on ChO/CPME.  相似文献   

18.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  相似文献   


19.
It is well accepted that estradiol (E2) plays an important role in the genesis and evolution of breast cancer. Quantitative evaluation indicates that in human breast tumor, estrone sulfate (E1S) ‘via sulfatase’ is a much more likely precursor for E2 than is androstenedione ‘via aromatase’. In previous studies, it was demonstrated that in isolated MCF-7 and T-47D breast cancer cell lines, estradiol can block estrone sulfatase activity. In the present study, the effect of E2 was explored using total normal and cancerous breast tissues. This study was carried out with post-menopausal patients with breast cancer. None of the patients had a history of endocrine, metabolic or hepatic diseases or had received treatment in the previous 2 months. Each patient received local anaesthetic (lidocaine 1%) and two regions of the mammary tissue were selected: (A) the tumoral tissue and (B) the distant zone (glandular tissue) which was considered as normal. Samples were placed in liquid nitrogen and stored at –80 °C until enzyme activity analysis. Breast cancer histotypes were ductal and post-menopausal stages were T2. Homogenates of tumoral or normal breast tissues (45–75 mg) were incubated in 20 mM Tris–HCl, pH 7.2 with physiological concentrations of [3H]-E1S (5 × 10−9 M) alone or in the presence of E2 (5 × 10−5 to 5 × 10−7 M) during 30 min or 3 h. E1S, E1 and E2 were characterized by thin layer chromatography and quantified using the corresponding standard. The sulfatase activity is significantly more intense with the breast cancer tissue than normal tissue, since the concentration of E1 was 3.20 ± 0.15 and 0.42 ± 0.07 pmol/mg protein, respectively after 30 min incubation. The values were 27.8 ± 1.8 and 3.5 ± 0.21 pmol/mg protein, respectively after 3 h incubation. Estradiol at the concentration of 5 × 10−7 M inhibits this conversion by 33% and 31% in cancerous and normal breast tissues, respectively and by 53% and 88% at the concentration of 5 × 10−5 M after 30 min incubation. The values were 24% and 18% for 5 × 10−7 M and 49% and 42% for 5 × 10−5 M, respectively after 3 h incubation. It was observed that [3H]-E1S is only converted to [3H]-E1 and not to [3H]-E2 in normal or cancerous breast tissues, which suggests a low or no 17β-hydroxysteroid dehydrogenase (17β-HSD) Type 1 reductive activity in these experimental conditions. In conclusion, estradiol is a strong anti-sulfatase agent in cancerous and normal breast tissues. This data can open attractive perspectives in clinical trials using this hormone.  相似文献   

20.
Biological properties of amino-terminal PTHrP analogues modified in the region 11–13 were examined using ROS 17/2.8 cells. [Leu11,D-Trp12,Arg13,Tyr36]PTHrP(1–36)amide had a 17-fold lower binding affinity for the receptor (apparent Kd: 5 × 10−8 M) than [Tyr36]PTHrP(1–36)amide or [Arg11,13,Tyr36]PTHrP(1–36)amide (apparent Kd for both: 2 × 10−9 M). Moreover, it is only a weak partial agonist despite completely inhibiting radioligand binding. [Leu11,D-Trp12,Arg13,Tyr36,Cys38]PTHrP(7–38) and PTHrP(7–34)amide had similar receptor affinities (apparent Kds: 5 × 10−8 M and 8 × 10−8 M), while that of [Nle8,18,Tyr34]bPTH(7–34)amide was more than 10-fold lower (apparent Kd: 2 × 10−6 M). These changes in biological properties suggest that high affinity receptor binding requires both amino- and carboxyl-terminal domains of the PTHrP(1–36) sequence and/or intramolecular interactions which are impaired by the D-Trp substitution for Gly12.  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