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.
Solubilized sterol C24-methyltransferase (24-SMT) was purified to homogeneity from a cell extract of the yeast Candida albicans (Ca) by anion exchange chromatography, gel permeation chromatography and fast performance liquid chromatography using a Mono Q column. The purified enzyme has an apparent molecular mass of 178 kDa on gel permeation chromatography and 43 kDa on SDS/PAGE, indicating that it is composed of four identical subunits. The substrate requirement of the native enzyme has an optimal specificity for zymosterol with associated kinetic constants of Km 50 μM and kcat of 0.01 s−1. The product of the enzyme incubated with zymosterol was fecosterol. Inhibition of the catalyst was observed with substrate analogs designed as transition state analogs (25-azalanosterol, Ki = 54 nM and 24 (R,S),25-epiminolanosterol, Ki = 11 nM) or as mechanism-based inactivators (26,27-dehydrozymosterol, Ki 9 μM) and kinact = 0.03 min−1) of the C24-methylation reaction. Product analogs ergosterol and fecosterol, but neither cholesterol nor sitosterol, inhibited activity affording Ki values of 20 and 72 μM, respectively. Ammonium and thia analogs of the intermediates of the sterol C24-methyl reaction sequence were effective growth inhibitors exhibiting IC50 values that ranged from 3 to 20 μM.  相似文献   

2.
Cyclobranol 8A, an analog of the cycloartenol substrate 1A for the plant sterol C24-methyltransferase (SMT), was shown to be an acceptor of the soybean SMT1 as well as an inhibitor of enzyme action. The Km and kcat for 8A was 37 μM and 0.006 min−1, respectively. The enzyme-generated product was identified by MS and 1H NMR to be a C24, C25-doubly alkylated Δ24(28)-olefin 10A. Inhibitor treatment was concentration and time-dependent affording an apparent Ki of 25 μM, a maximum rate of inactivation of 0.15 min−1 and a partition ratio (kcat/kinact) calculated to be 0.04.  相似文献   

3.
The aggregation of proteins is believed to be intimately connected to many neurodegenerative disorders. We recently reported an “Ockham's razor”/minimalistic approach to analyze the kinetic data of protein aggregation using the Finke–Watzky (F–W) 2-step model of nucleation (A → B, rate constant k1) and autocatalytic growth (A + B → 2B, rate constant k2). With that kinetic model we have analyzed 41 representative protein aggregation data sets in two recent publications, including amyloid β, α-synuclein, polyglutamine, and prion proteins (Morris, A. M., et al. (2008) Biochemistry 47, 2413-2427; Watzky, M. A., et al. (2008) Biochemistry 47, 10790–10800). Herein we use the F–W model to reanalyze protein aggregation kinetic data obtained under the experimental conditions of variable temperature or pH 2.0 to 8.5. We provide the average nucleation (k1) and growth (k2) rate constants and correlations with variable temperature or varying pH for the protein α-synuclein. From the variable temperature data, activation parameters ΔG, ΔH, and ΔS are provided for nucleation and growth, and those values are compared to the available parameters reported in the previous literature determined using an empirical method. Our activation parameters suggest that nucleation and growth are energetically similar for α-synuclein aggregation (ΔGnucleation = 23(3) kcal/mol; ΔGgrowth = 22(1) kcal/mol at 37 °C). From the variable pH data, the F–W analyses show a maximal k1 value at pH ~ 3, as well as minimal k1 near the isoelectric point (pI) of α-synuclein. Since solubility and net charge are minimized at the pI, either or both of these factors may be important in determining the kinetics of the nucleation step. On the other hand, the k2 values increase with decreasing pH (i.e., do not appear to have a minimum or maximum near the pI) which, when combined with the k1 vs. pH (and pI) data, suggest that solubility and charge are less important factors for growth, and that charge is important in the k1, nucleation step of α-synuclein. The chemically well-defined nucleation (k1) rate constants obtained from the F–W analysis are, as expected, different than the 1/lag-time empirical constants previously obtained. However, k2 × [A]0 (where k2 is the rate constant for autocatalytic growth and [A]0 is the initial protein concentration) is related to the empirical constant, kapp obtained previously. Overall, the average nucleation and average growth rate constants for α-synuclein aggregation as a function of pH and variable temperature have been quantitated. Those values support the previously suggested formation of a partially folded intermediate that promotes aggregation under high temperature or acidic conditions.  相似文献   

4.
The synthesis and pharmacology of 15 1-deoxy-Δ8-THC analogues, several of which have high affinity for the CB2 receptor, are described. The deoxy cannabinoids include 1-deoxy-11-hydroxy-Δ8-THC (5), 1-deoxy-Δ8-THC (6), 1-deoxy-3-butyl-Δ8-THC (7), 1-deoxy-3-hexyl-Δ8-THC (8) and a series of 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=0–4, 6, 7, where n=the number of carbon atoms in the side chain−2). Three derivatives (1719) of deoxynabilone (16) were also prepared. The affinities of each compound for the CB1 and CB2 receptors were determined employing previously described procedures. Five of the 3-(1′,1′-dimethylalkyl)-1-deoxy-Δ8-THC analogues (2, n=1–5) have high affinity (Ki=<20 nM) for the CB2 receptor. Four of them (2, n=1–4) also have little affinity for the CB1 receptor (Ki=>295 nM). 3-(1′,1′-Dimethylbutyl)-1-deoxy-Δ8-THC (2, n=2) has very high affinity for the CB2 receptor (Ki=3.4±1.0 nM) and little affinity for the CB1 receptor (Ki=677±132 nM).
Scheme 3. (a) (C6H5)3PCH3+ Br, n-BuLi/THF, 65°C; (b) LiAlH4/THF, 25°C; (c) KBH(sec-Bu)3/THF, −78 to 25°C then H2O2/NaOH.  相似文献   

5.
A study is reported of complexation reactions of cyclohexaamylose (Cy) with 1-adamantanecarboxylic acid and its anion using conductometry, pH potentiometry, and 13C nmr spectrometry. Binary and ternary (2 mol Cy/mol substrate) complexes are detected with both the acid and anion, and standard entropies and enthalpies of complexation are determined from the temperature dependences of the formation constants for all except the very weak ternary complex with the anion. Both the 13C nmr results and the entropy of complexation confirm the earlier suggestion that the anion in binary complexation is structured with the adamantanyl group in proximity to, but not penetrating, the Cy cavity. However, a negative ΔSo for formation of this complex is reported which casts doubt on an earlier proposal that the adamantyl binary complex binding mode involves an “apolar” mechanism accompanied by loss of solvated water molecules. Values are also reported for pKa, ΔHo, and ΔSo for the aqueous dissociation of 1-adamantanecarboxylic acid.  相似文献   

6.
The sterol fraction from the marine worm Linneus torquatus Coe (phylum Nemertini, class Anopla, family Lineidae) has been isolated, separated by HPLC and preparative TLC on AgNO3-impregnated silica gel, and sterols identified using GC, GC-MS and NMR spectroscopy. It was shown that the fraction contains at least 12 sterols belonging mainly to Δ5,22, Δ5,24(28) and Δ5 series. The major sterol components were 24-methylcholesta-5,24(28)-dien-3β-ol, cholesta-5,22E-dien-3β-ol, 24-nor-cholesta-5,22-dien-3β-ol and cholesterol.  相似文献   

7.
To gain the structure–activity relationship of Δ1-androstenediones (Δ1-ADs) as mechanism-based inactivator of aromatase, series of 2-alkyl- and 2-alkoxy-substitiuted Δ1-ADs (6 and 9) as well as 2-bromo-Δ1-AD (14) were synthesized and tested. All of the inhibitors examined blocked aromatase in human placental microsomes in a competitive manner. In a series of 2-alkyl-Δ1-ADs (6), n-hexyl compound 6f was the most powerful inhibitor with an apparent Ki value of 31 nM. The inhibitory activities of 2-alkoxy steroids 9 decreased in relation to length of the alkyl chain up to n-hexyloxy group (Ki: 95 nM for methoxy 9a). All of the alkyl steroids 6 along with the alkoxy steroid 9, except for the ethyl and n-propyl compounds 6b and 6c, caused a time-dependent inactivation of aromatase. The inactivation rates (kinact: 0.020–0.084 min−1) were comparable to that of the parent compound Δ1-AD. The inactivation was prevented by the substrate AD, and no significant effect of l-cysteine on the inactivation was observed in each case. The results indicate that the 2-hexyl compound 6f act as the most powerful mechanism-based inactivator of aromatase among Δ1-AD analogs and may be submitted to the preclinical study in estrogen-dependent breast cancer.  相似文献   

8.
γ-Aminobutyraldehyde dehydrogenase from Escherichia coli K-12 has been purified and characterized from cell mutants able to grow in putrescine as the sole carbon and nitrogen source. The enzyme has an Mr of 195 000±10 000 in its dimeric form with an Mr of 95 000±1000 for each subunit, a pH optimum at 5.4 in sodium citrate buffer, and does not require bivalent cations for its activity. Km values are 31.3±6.8 μM and 53.8±7.4 μM for Δ-1-pyrroline and NAD+, respectively. An inhibitory capacity for NADH is also shown using the purified enzyme.  相似文献   

9.
3β-Hydroxysteroid dehydrogenase (3β-HSD)/Δ5→4-isomerase activity in steroidogenic tissues is required for the synthesis of biologically active steroids. Previously, by use of dehydroepiandrosterone (3β-hydroxy-5-androsten-17-one, DHEA) as substrate, it was established that in addition to steroidogenic tissues 3β-HSD/Δ5→4-isomerase activity also is expressed in extraglandular tissues of the human fetus. In the present study, we attempted to determine whether the C-5,C-6-double bond of DHEA serves to influence 3β-HSD activity. For this purpose, we compared the efficiencies of a 3β-hydroxy-5-ene steroid (DHEA) and a 3β-hydroxy-5α-reduced steroid (5α-androstane-3β,17β-diol, 5α-A-diol) as substrates for the enzyme. The apparent Michaelis constant (Km) for 5α-A-diol in midtrimester placenta, fetal liver, and fetal skin tissues was at least one order of magnitude higher than that for DHEA, viz the apparent Km of placental 3β-HSD for 5α-A-diol was in the range of 18 to 40 μmol/l (n = 3) vs 0.45 to 4 μmol/l for DHEA (n = 3); for the liver enzyme, 17 μmol/l for 5α-A-diol and 0.60 μmol/l for DHEA, and for the skin enzyme 14 and 0.18 μmol/l, respectively. Moreover, in 13 human fetal tissues evaluated the maximal velocities obtained with 5α-A-diol as substrate were higher than those obtained with DHEA. A similar finding in regard to Kms and rates of product formation was obtained by use of purified placental 3β-HSD with DHEA, pregnenolone, and 3β-hydroxy-5α-androstan-17-one (epiandrosterone) as substrates: the Km of 3β-HSD for DHEA was 2.8 μmol/l, for pregnenolone 1.9 μmol/l, and for epiandrosterone 25 μmol/l. The specific activity of the purified enzyme with pregnenolone as substrate was 27 nmol/mg protein·min and, with epiandrosterone, 127 nmol/mg protein·min. With placental homogenate as the source of 3β-HSD, DHEA at a constant level of 5 μmol/l behaved as a competitive inhibitor when the radiolabeled substrate, [3H]5α-A-diol, was present in concentrations of 20 to 60 μmol/l, but a lower substrate concentrations the inhibition was of the mixed type; similar results were obtained with [3H]DHEA as the substrate at variable concentrations in the presence of a fixed concentration of 5α-A-diol (40 μmol/l). These findings are indicative that both steroids bind to a common site on the enzyme, however, the binding affinity for these steroids appear to differ markedly as suggested by the respective Kms. Studies of inactivation of purified placental 3β-HSD/Δ5→4-isomerase by an irreversible inhibitor, viz 5,10-secoestr-4-yne-3,10,17-trione, were suggestive that the placental protein adopts different conformations depending on whether the steroidal substrate has a 5α-configuration, e.g. epiandrosterone, or a C-5,C-6-double bond e.g. DHEA or pregnenolone. The lower rates of product formation obtained with placenta and fetal tissues by use of 3β-hydroxy-5-ene steroids as substrates when compared with those obtained with 3β-hydroxy-5α-reduced steroids may be explained by a combination of factors, including: (i) inhibition of 3β-HSD activity by end products of metabolism of 3β-hydroxy-5-ene steroids, e.g. 4-androstene-3,17-dione formed with DHEA as substrate; (ii) higher binding affinity of the enzyme for 3β-hydroxy-5-ene steroids—and possibly for their 3-oxo-5-ene metabolites; (iii) lack of a requirement for the isomerization step with 5α-reduced steroids as substrates, and (iv) the possible presence in fetal tissues of an enzyme with 3β-HSD activity only (i.e. no Δ5→4-isomerase).  相似文献   

10.
Growth of Cryptococcus neoformans was inhibited by nine nitrogen and sulfur-containing sterols with a heteroatom positioned at C3, C7, C24, C25 or C32 in the lanostane frame. Analysis of the sterol composition of control and treated cells by GC-MS and 1H NMR has proven that the C-methylation reaction catalyzed by the sterol 24-C-methyltransferase (24-SMT) is the crucial first step in a kinetically favored pathway that fails to include obtusifoliol or zymosterol as intermediates. Cultures fed [methyl-2H3]methionine led to two deuterium atoms into each of the newly biosynthesized sterols forming a route lanosterol, eburicol (24(28)-methylene-24,25-dihydrolanosterol), 32-noreburicol and ergost-7-enol to ergosterol. Examination of the substrate specificity of a soluble 24-SMT from C. neoformans showed lanosterol to be the optimal acceptor molecule. Incubation with the test compounds generated induced amounts of lanosterol, eburicol or 32-noreburicol concurrent with a decrease of ergosterol. Among them 24(R,S),25-epiminolanosterol (inhibitor of 24-SMT) showed the most potent in vitro antifungal activity comparable to those of itraconazole (inhibitor of the 14-demethylase). Taken together, these data indicate that treatment with substrate-based inhibitors of 24-SMT, a catalyst not found in humans, can disrupt ergosterol homeostasis involved with fungal growth and therefore these compounds can provide leads for rational drug design of opportunistic pathogens.  相似文献   

11.
Lipid phase transitions in Escherichia coli membranes and in dispersions of the extracted lipids were studied using the negatively charged fluorescence probe 1-anilinonaphthalene-8-sulfonate (ANS) and the hydrophobic fluorescence probe N-phenyl-1-naphthylamine (NPN). The fluorescence change, ΔI, at the phase transition approaches a limiting value (ΔI)lim with increasing dye concentration. A comparison of the limiting values (Δ)limNPN obtained for membranes and the lipid standard allows us to estimate the lipid fraction, ρ, in the membrane that takes part in the phase transition (ρ = 80%). The same procedure carried out with ANS yields a value of 42.5% for the lipid fraction that is accessible from the aqueous phase. These values, combined with published freeze-etching data for the particle density within the fracture plane of membranes are used to quantify the Davson-Danielli-Robertson-Benson-Singer membrane model which assumes a fluid lipid bilayer with “integral” proteins embedded in the lipid matrix and surface proteins attached to the lipid head groups. It appears that on the average one “integral” membrane protein is surrounded by about 600 lipid molecules and that about 130 of these molecules are closely coupled to the protein molecule, forming an halo in which the chain-chain interaction between the lipids is disturbed. About half of the bilayer surface is covered with proteins; part of these seem to be stacked.  相似文献   

12.
We investigated the adaptative response of S. cerevisiae in sod mutants (sod1Δ, sod2Δ and sod1Δsod2Δ) after H2O2 treatment in the stationary phase. sod2Δ and sod1Δsod2Δ demonstrated the highest levels of GSH in the control, suggesting that pathways which include GSH protect these double mutants against oxidative stress. In addition, sod1Δ and sod1Δsod2Δ had higher iron levels than the wild-type, independently of H2O2 stress. Fe levels were increased in sod2Δ following H2O2 In addition, the sod2Δ mutant was more sensitive to H2O2 treatment than the wild-type. These results suggest that sod2Δ sensibility may be associated with •OH production by the Fenton reaction. This increased iron demand in the sod2Δ mutant may be a reflection of the cells’ efforts to reconstitute proteins that are inactivated in conditions of excess superoxide. MDA levels were assayed by HPLC in these mutants. The highest MDA levels could be observed after 10mM H2O2 treatment in the sod1Δsod2Δ double mutant. After treatment with a GSH inhibitor, the MDA level was still higher in the same strain. Thus, both direct and indirect GSH pathways are involved in the protection of lipid membranes and proteins in these mutants and may constitute an adaptative response to enhanced basal oxidative damage produced by superoxide.  相似文献   

13.
Rate constants for C(α)-proton transfer from racemic 2-(1-hydroxyethyl)-3,4-dimethylthi-oazolium ion catalyzed by lyoxide ion and various oxygen-containing and amine buffers were determined by iodination at 25°C and ionic strength 1.0 in H2O. Thermodynamically unfavorable C(α)-proton transfer to oxygen-containing and amine bases shows general base catalysis with a Brønsted β value of ≥0.92 for bases of pKa ≤ 15; this indicates that the thermodynamically favorable protonation reaction in the reverse direction has a Brønsted α value ≤0.08, which is consistent with diffusion-controlled reprotonation of the C(α)-enamine by most acids. General base catalysis is detectable because there is an 85-fold negative deviation from the Brønsted correlation by hydroxide ion. Primary kinetic isotope effects of (kH/kD)obsd = 1.0 for thermodynamically unfavorable proton transfer to buffer bases and hydroxide ion (ΔpKa ≤ −6) and a secondary solvent isotope effect of kDO/kHO = 2.3 for C(α)-proton transfer are consistent with a very late, enamine-like transition state and rate-limiting diffusional separation of buffer acids from the C(α)-enamine in the rate-limiting step, as expected for a “normal” acid. The second-order rate constants for catalysis by buffer bases were used to calculate a pKa of 21.8 for the C(α)-proton assuming a rate constant of 3 × 109 −1 s−1 for the diffusion-controlled reprotonation of the C(α)-enamine by buffer acids in the reverse direction. It is concluded (i) that C(α)-proton removal occurs at the maximum possible rate for a given equilibrium constant, and (ii) that C(α)-enamines can have a significant lifetime in aqueous solution and on thiamin diphosphate-dependent enzymes.  相似文献   

14.
A rapid and sensitive kinetic assay of lanosterol 14α-demethylation has been developed and analyzed. Three substrates, [32-3H]-24,25-dihydrolanosterol, [32-3H]lanost-8-en-3β,32-diol, and [32-3H]lanost-7-en-3β-32-diol, were studied. In all cases, the rate of tritium released into aqueous solution provided a simple and direct assay of 14α-demethylase activity. The kinetic parameters of Km and Vmax for each substrate have been determined in a reconstituted system from rat liver. The percentage of turnover monitored by the novel tritium release assay was comparable to that observed by conventional GC methods. Separation of unreacted sterol from tritiated formate and water via reverse-phase chromatography permitted several samples to be analyzed at once.  相似文献   

15.
Photosynthetic responses to irradiance and temperature of “leaves” and receptacles were compared in February (vegetative stage) and May (reproductive stage) in the seaweed, Hizikia fusiforme (Harvey) Okamura (Sargassaceae, Phaeophyta) from Nanao Island, Shantou, China. Irradiance-saturated photosynthesis (Pmax) was significantly higher in receptacles than in “leaves” on a fresh weight basis, and that of “leaves” was greater in May than in February at ambient seawater temperatures. The optimum temperature for Pmax was 30C for both “leaves” and receptacles, being 5–10C higher than the ambient seawater temperature. The apparent photosynthetic efficiencies were greater in receptacles than in “leaves” within the tested temperature range of 10–40C. The irradiance for saturating photosynthesis for both “leaves” and receptacles was temperature-dependent, with the highest values (about 200μmolphotonsm−2s−1) at 30C.  相似文献   

16.
Oxidative deamination of putrescine, the precursor of polyamines, gives rise to γ-aminobutyraldehyde (ABAL). In this study an aldehyde dehydrogenase, active on ABAL, has been purified to electrophoretic homogeneity from rat liver cytoplasm and its kinetic behaviour investigated. The enzyme is a dimer with a subunit molecular weight of 51,000. It is NAD+-dependent, active only in the presence of sulphhydryl compounds and has a pH optimum in the range 7.3–8.4. Temperatures higher than 28°C promote slow activation and the process is favoured by the presence of at least one substrate. Km for aliphatic aldehydes decreases from 110 μM for ABAL and acetaldehyde to 2–3 μM for capronaldehyde. The highest relative V-values have been observed with ABAL (100) and isobutyraldehyde (64), and the lowest with acetaldehyde (14). Affinity for NAD+ is affected by the aldehyde present at the active site: Km for NAD+ is 70 μM with ABAL, 200 μM with isobutyraldehyde and capronaldehyde, and>800 μM with acetaldehyde. The kinetic behaviour at 37°C is quite complex; according to enzymatic models, NAD+ activates the enzyme (Kact 500 μM) while NADH competes for the regulatory site (Kin 70 μM). In the presence of high NAD+ concentrations (4 mM), ABAL promotes further activation by binding to a low-affinity regulatory site (Kact 10 mM). The data show that the enzyme is probably an E3 aldehyde dehydrogenase, and suggest that it can effectively metabolize aldehydes arising from biogenic amines.  相似文献   

17.
A method based on gas chromatography–mass spectrometry–selected-ion monitoring was developed to measure the main metabolites of 17α-methyltestosterone, 17α-methyl-5α-androstan-3α,17β-diol and 17α-methyl-5β-androstan-3α,17β-diol, in human urine. 17α-Methyl-[2H3]-5α-androstan-3α,17β-diol and 17α-methyl-[2H3]-5β-androstan-3α,17β-diol were used as internal standards. The methods involved purification using a Sep-Pak C18 cartridge, hydrolysis by β-glucuronidase from Ampullaria and derivatization with N-methyl-N-trimethylsilyl-trifluoroacetamide/dithioerythriol/ammonium iodide. Quantitation was achieved by selected-ion monitoring of the characteristic fragment ions ([(M+H)−2×TMSOH]+) of the di-TMS derivatives on the chemical ionization mode. The method provides a specific, sensitive and reliable technique to determine the urine levels of 17α-methyl-5α-androstan-3α,17β-diol and 17α-methyl-5β-androstan-3α,17β-diol, and can be applied to pharmacokinetic studies of 17α-methyltestosterone.  相似文献   

18.
Three series of new cannabinoids were prepared and their affinities for the CB1 and CB2 cannabinoid recptors were determined. These are the 1-methoxy-3-(1′,1′-dimethylalkyl)-, 1-deoxy-11-hydroxy-3-(1′,1′-dimethylalkyl)- and 11-hydroxy-1-methoxy-3-(1′,1′-dimethylalkyl)-Δ8-tetrahydrocannabinols, which contain alkyl chains from dimethylethyl to dimethylheptyl appended to C-3 of the cannabinoid. All of these compounds have greater affinity for the CB2 receptor than for the CB1 receptor, however only 1-methoxy-3-(1′,1′-dimethylhexyl)-Δ8-THC (JWH-229, 6e) has effectively no affinity for the CB1 receptor (Ki=3134±110 nM) and high affinity for CB2 (Ki=18±2 nM).  相似文献   

19.
The kinetics of the decomposition reactions of the CO(py)3(CO3)(H2O)+ ion have been investigated in aqueous perchloric acid solutions over a range of hydrogen ion concentrations (0.10 to 5.0 M) and at two ionic strengths (I = 1.0 and 5.0 M). At the lower ionic strength, plots of ln (AtA versus time show a nonlinearity that is consistent with that expected for consecutive first-order reactions. The rates of the faster reaction are similar to those reported for the spontaneous reduction of aquopyridine-cobalt(III) cations. At the higher ionic strength, the above noted curvature is not apparent and the decarboxylation kinetics of the title complex may be described by a pseudo-first-order rate constant: kobs = k[H3O+]. At 20°C, k = (1.75−+0.09) s−1 M−1 with activation parameters ofΔH = (97 −+ 4) kJ mol−1 and ΔS = −(54 −+ 32) J deg−1 mol−1. These kinetic parameters are compared with those previously reported for the similar complexes, Co(py)4CO3+ and Co(py)2(CO3)(H2O)2+.  相似文献   

20.
We investigate the effect of the prostaglandin D2 metabolite Δ12−PGD2 (9−Deoxy−Δ9, Δ12−13,14-dihydroprostaglandin D2) on collagen synthesis in human osteoblast. Δ12-PGJ2 at 10−5M enhanced collagen synthesis in the presence of 2 mM α-glycerophosphate-2Na. The stimulative effect appeared as early as 3 days after addition and continued until 22 days. The enhancement of type I collagen synthesis was confirmed by polyacrylamide gel electrophoresis. The potency was the same as 101t-8M 1 α, 25 dihydroxy vitamine D3 (1,25(OH)2D3). Northern blot analysis showed that 10−5M Δ 12-PGD2 and 10−8M 1,25(OH)2D3 enhanced the transcribtion of type 1 procollagen (α1) mRNA levels in osteoblasts.  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