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 46 毫秒
1.
The fungus Humicola lutea 120-5 cultivated in casein-containing media, in the presence or absence of inorganic phosphate (Pi), excretes three different molecular forms of acid phosphatase (with Mr values of approximately 140, 70 and 35 kDa). The enzyme forms were isolated and purified 30–100-fold by a procedure involving two steps of ion-exchange chromatography and Sephadex G-200 gel chromatography. It was found that the fungus excretes only one of the phosphatases with the highest Mr (140 kDa) during growth on medium with inorganic nitrogen source (NaNO3). This form (designed AcPh I) was assumed to be a constitutive, since it showed resistance to high Pi-concentrations (10 mM) and its biosynthesis was not affected by the type of nitrogen source (casein or NaNO3). The other two forms (AcPh II-70 and AcPh III-35 kDa) were competitively inhibited by Pi (K i = 0.5 and 0.2 mM, respectively) and were induced by casein. The K m values of AcPh I and AcPh II were estimated as 1.3 mM, while AcPh III showed a higher affinity for p-nitrophenylphosphate (pNPP) with K m of 0.5 mM. The AcPh I–III fractions demonstrated a pH optimum in the range of 4.5–4.8 and an optimal temperature of 55 °C using pNPP as a substrate. This revised version was published online in November 2006 with corrections to the Cover Date.  相似文献   

2.
A series of systematically modified vanadyl-β-diketone complexes, VO(β-diketone)2, bearing substituent groups with different electron inductive properties were synthesized and evaluated as inhibitors against calf-intestine alkaline phosphatase (APase). A combination of biochemical and quantum mechanical techniques were employed to identify structure-activity relationships relevant for rational design of phosphatase inhibitors. Kinetic parameters and activation free energy, enthalpy, and entropy for calf-intestine APase-catalyzed dephosphorylation of para-nitrophenylphosphate were also determined along with the inhibition constants (Ki) for the VO(β-diketone)2 complexes. Increased positive charge on the vanadyl group increases the inhibition potency of the complex while the absence of an available coordination site on the complex decreases its inhibition potency. These findings correlate well with the results of ab initio electron density calculations for the complexes.  相似文献   

3.
We describe the cloning, expression and purification of the bovine XM866409 form of pyroglutamyl peptidase type-1 (PAP1). The cloned nucleotide sequence has an ORF coding for a primary sequence of 209 amino acid residues, which displays 98% identity with the human AJ278828 form of the enzyme. Three amino acid residues at positions 81, 205 and 208 were found to vary between the two sequences. The recombinant bovine PAP1 with a C-terminal His6 tag (rBtaPAP16H) was expressed in Escherichia coli XL10-Gold cells and purified by immobilised nickel ion affinity chromatography resulting in a yield of 2.6 mg of PAP1 per litre of culture. Purified rBtaPAP16H had a specific activity of 3633 units mg−1. SDS-PAGE revealed a band for bovine PAP1 with a molecular weight of ∼24 kDa, which is in good agreement with previously reported data on PAP1. The K m and k cat values obtained for rBtaPAP16H were 59 μM and 3.5 s−1, respectively. The optimum pH for activity was 9.0–9.5 and the optimum temperature was 37 °C. rBtaPAP16H was found to have an absolute requirement for the thiol-reducing agent DTT, consistent with the expected property of a cysteine protease. Kinetic studies using the peptides pGlu-His-Pro-NH2 (TRH), pGlu-Ala and pGlu-Val revealed K i values of 44.1, 141 and 652.17 μM, respectively. The lowest K i, observed for Thyrotropin-releasing Hormone (TRH), indicates that rBtaPAP16H has a higher affinity for tripeptides over dipeptides.  相似文献   

4.
为探究岩溶植物的光合生理适应机制,采用Li-6400XT便携式光合作用测量系统,对广西平果市岩溶区8种适生植物的叶片净光合速率(Pn)、气孔导度(Gs)、胞间CO2浓度(Ci)、蒸腾速率(Tr)、水分利用效率(WUE)和气孔限制值(Ls)等光合特征参数进行了测定分析。结果表明:(1)6个光合特征参数在种内和种间均存在不同程度的变异,并且种内变异均大于种间变异。(2)Gs和Tr的变化主要来源于种间变异(46.72%~49.76%),而Pn、Ci、WUE和Ls变化主要来源于种内变异(48.66%~64.50%)。在生活型水平上,Pn、Gs和Tr的种内变异表现为常绿植物小于落叶植物,而Ci、WUE和Ls则相反。(3)各参数的种间变异均表现为落叶植...  相似文献   

5.

Background

Cationic lipid DNA complexes based on DOTAP (1,2-dioleoyl-3-(trimethyammonium) propane) and mixtures of DOTAP and cholesterol (DC) have been previously optimized for transfection efficiency in the absence of serum and used as a non-viral gene delivery system. To determine whether DOTAP and DC lipid DNA complexes could be obtained with increased transfection effciency in the presence of high serum concentrations, the composition of the complexes was varied systematically and a total of 162 different complexes were analyzed for transfection efficiency in the presence and absence of high serum concentrations.

Results

Increasing the ratio of DOTAP or DC to DNA led to a dose dependent enhancement of transfection efficiency in the presence of high serum concentrations up to a ratio of approximately 128 nmol lipid/μg DNA. Transfection efficiency could be further increased for all ratios of DOTAP and DC to DNA by addition of the DNA condensing agent protamine sulfate (PS). For DOTAP DNA complexes with ratios of ≤ 32 nmol/μg DNA, peak transfection efficiencies were obtained with 4 μg PS/μg DNA. In contrast, increasing the amount of PS of DC complexes above 0.5 μg PS /μg DNA did not lead to significant further increases in transfection efficiency in the presence of high serum concentrations. Four complexes, which had a similar high transfection efficiency in cell culture in the presence of low serum concentrations but which differed largely in the lipid to DNA ratio and the amount of PS were selected for further analysis. Intravenous injection of the selected complexes led to 22-fold differences in transduction efficiency, which correlated with transfection efficiency in the presence of high serum concentrations. The complex with the highest transfection efficiency in vivo consisted of 64 nmol DC/ 16 μg PS/ μg DNA. Physical analysis revealed a predicted size of 440 nm and the highest zeta potential of the complexes analyzed.

Conclusions

Optimization of cationic lipid DNA complexes for transfection efficiency in the presence of high concentrations of serum led to the identification of a DC complex with high transduction efficiency in mice. This complex differs from previously described ones by higher lipid to DNA and PS to DNA ratios. The stability of this complex in the presence of high concentrations of serum and its high transduction efficiency in mice suggests that it is a promising candidate vehicle for in vivo gene delivery.  相似文献   

6.
A low-Mr tight binding proteinase inhibitor was purified from bovine muscle by alkaline denaturation of cysteine proteinases, gel filtration on Sexphadex G-75 and affinity chromatography on carboxymethyl-papain-Sepharose. Chromatofocusing separated three isoforms which are similar in their Mr of about 14 000, their stability with heating at 80°C and their inhibitory activity towards cathepsin H, cathepsin B and papain. The equilibrium constants (Ki) were determined for these three cysteine proteinases but for cathepsin H, association (kass) and dissociation (kdiss) rate constants were also evaluated. Ki values of 56 nM and 8.4 nM were found for cathepsin B and cathepsin H, respectively. For papain, Ki was in the range of 0.1–1 nM. The kinetic features of enzyme-inhibitor binding suggest a possible role for this low-Mr protein inhibitor in controlling ‘in vivo’ cathepsin H proteolytic activity. With regard to cathepsin B, such a physiological role was less evident.  相似文献   

7.
Antimycin A is the most frequently used specific and powerful inhibitor of the mitochondrial respiratory chain. We used all-atom molecular dynamics (MD) simulations to study the dynamic aspects of the interaction of antimycin A with the Qi site of the bacterial and bovine bc1 complexes embedded in a membrane. The MD simulations revealed considerable conformational flexibility of antimycin and significant mobility of antimycin, as a whole, inside the Qi pocket. We conclude that many of the differences in antimycin binding observed in high-resolution x-ray structures may have a dynamic origin and result from fluctuations of protein and antimycin between multiple conformational states of similar energy separated by low activation barriers, as well as from the mobility of antimycin within the Qi pocket. The MD simulations also revealed a significant difference in interaction between antimycin and conserved amino acid residues in bovine and bacterial bc1 complexes. The strong hydrogen bond between antimycin and conserved Asp-228 (bovine numeration) was observed to be frequently broken in the bacterial bc1 complex and only rarely in the bovine bc1 complex. In addition, the distances between antimycin and conserved His-201 and Lys-227 were consistently larger in the bacterial bc1 complex. The observed differences could be responsible for a weaker interaction of antimycin with the bacterial bc1 complex.  相似文献   

8.
Objective: Our goal was to test any association between human plasma circulating levels of monocyte chemoattractant protein‐1 (cMCP‐1) and insulin resistance and to compare monocyte chemoattractant protein‐1 (MCP‐1) adipose tissue gene expression and cMCP‐1 in relation with inflammatory markers. Research Methods and Procedures: cMCP‐1 was measured in n = 116 consecutive control male subjects to whom an insulin sensitivity (Si) test was performed. Circulating levels of soluble CD14, soluble tumor necrosis factor receptor type 2 (sTNFR2), soluble interleukin‐6 (sIL‐6), and adiponectin also were measured. Subcutaneous adipose tissue samples were obtained from n = 107 non‐diabetic and type 2 diabetic subjects with different degrees of obesity. Real‐time polymerase chain reaction was used to measure gene expression of MCP‐1, CD68, tumor necrosis factor‐α (TNF‐α), and its receptor TNFR2. Results: In the Si study, no independent effect of cMCP‐1 levels on insulin sensitivity was observed. In the expression study, in non‐diabetic subjects, MCP‐1 mRNA had a positive correlation with BMI (r = 0.407, p = 0.003), TNF‐α mRNA (r = 0.419, p = 0.002), and TNFR2 mRNA (r = 0.410, p = 0.003). In these subjects, cMCP‐1 was found to correlate with waist‐to‐hip ratio (r = 0.322, p = 0.048). In patients with type 2 diabetes, MCP‐1 mRNA was up‐regulated compared with non‐diabetic subjects. TNF‐α mRNA was found to independently contribute to MCP‐1 mRNA expression. In this group, CD68 mRNA was found to correlate with BMI (r = 0.455, p = 0.001). Discussion: cMCP‐1 is not associated with insulin sensitivity in apparently healthy men. TNF‐α is the inflammatory cytokine associated with MCP‐1 expression in subcutaneous adipose tissue.  相似文献   

9.
[3H]-thymidine is commonly used to analyze the accumulation of [3H]-labeled chromatin fragments in cells undergoing apoptosis. This study shows that [3H]-thymidine incorporation within DNA is sufficient per se to inhibit growth and to induce apoptosis in canine kidney epithelial cells and porcine aorta endothelial cells. Despite high-level [3H]-thymidine-DNA labeling, rat vascular smooth muscle cells (VSMC) showed only modest inhibition of growth and induction of apoptosis compared to other cell types. Similarly to serum deprivation, apoptosis triggered by [3H]-thymidine labeling was sharply potentiated by VSMC transfection with a functional analogue of c-myc, E1A-adenoviral protein (VSMC-E1A), and was suppressed by stimulation of cAMP signaling with forskolin as well as by and Na/K pump inhibition with ouabain. Both apoptosis induction and growth suppression seen in [3H]-thymidine-treated VSMC-E1A were reduced by the pan-caspase inhibitor z-VAD.fmk. Thus, our results show that the differential efficiency of the apoptotic machinery determines cell type-specific attenuation of growth in cells with [3H]-thymidine-labeled DNA. They also demonstrate that [3H]-thymidine-treated and serum-deprived VSMC employ common intermediates of the apoptotic machinery, including steps that are potentiated by E1A-adenoviral protein and inhibited by activation of cAMP signaling as well as by inversion of the intracellular [Na+] i /[K+] i ratio.  相似文献   

10.
Chen  C. R.  Condron  L. M.  Sinaj  S.  Davis  M. R.  Sherlock  R. R.  Frossard  E. 《Plant and Soil》2003,256(1):115-130
Vegetative conversion from grass to forest may influence soil nutrient dynamics and availability. A short-term (40 weeks) glasshouse experiment was carried out to investigate the impacts of ryegrass (Lolium perenne) and radiata pine (Pinus radiata) on soil phosphorus (P) availability in 15 grassland soils collected across New Zealand using 33P isotopic exchange kinetics (IEK) and chemical extraction methods. Results from this study showed that radiata pine took up more P (4.5–33.5 mg P pot–1) than ryegrass (1.1–15.6 mg pot–1) from the soil except in the Temuka soil in which the level of available P (e.g., E 1min Pi, bicarbonate extractable Pi) was very high. Radiata pine tended to be better able to access different forms of soil P, compared with ryegrass. There were no significant differences in the level of water soluble P (Cp, intensity factor) between soils under ryegrass and radiata pine, but the levels of Cp were generally lower compared with original soils due to plant uptake. The growth of both ryegrass and radiata pine resulted in the redistribution of soil P from the slowly exchangeable Pi pool (E > 10m Pi, reduced by 31.8% on the average) to the rapidly exchangeable Pi (E 1min-1d Pi, E 1d-10m Pi) pools in most soils. The values of R/r 1 (the capacity factor) were also generally greater in most soils under radiata pine compared with ryegrass. Specific P mineralisation rates were significantly greater for soils under radiata pine (8.4–21.9%) compared with ryegrass (0.5–10.8%), indicating that the growth of radiata pine enhanced mineralisation of soil organic P. This may partly be ascribed to greater root phosphatase activity for radiata pine than for ryegrass. Plant species × soil type interactions for most soil variables measured indicate that the impacts of plant species on soil P dynamics was strongly influenced by soil properties.  相似文献   

11.
A new series of pyrazolo[3,4-d]pyrimidine-6-one derivatives (2a–2j) were prepared by using the Biginelli multicomponent cyclocondensation of 3-methyl-1-phenyl-1H-pyrazol-5(4H)-one (1a), different aromatic aldehydes, and urea with a catalytic amount of HCl at reflux temperature. These compounds were characterized by IR, 1H NMR, 13C NMR, and Mass spectral data. In vitro antiamoebic activity was performed against HM1:IMSS strain of Entamoeba histolytica. The results showed that the compounds 2b, 2i, and 2j with IC50 values of 0.37 µM, 0.04 µM, and 0.06 µM, respectively, exhibited better antiamoebic activity than the standard drug metronidazole (IC50?=?1.33 µM). The toxicological studies of these compounds on human breast cancer MCF-7 cell line showed that the compounds 2b, 2i, and 2j exhibited >80% viability at the concentration range of 1.56–50 µM.  相似文献   

12.
Starch dextrins of different molecular sizes (DPn 311, 142 and 39) were prepared by hydrolyzing a high amylose maize starch in acidic alcohol solutions. The dextrins were dissolved in an aqueous dimethyl sulfoxide solution (90% DMSO), and then the solution was allowed to migrate down into n-butanol separated by a membrane filter. The complex was gradually formed between the dextrin and butanol, and precipitated in the butanol layer. The dextrin–butanol complex yielded V6-I type crystals with broad reflections (d-spacings 1.123, 0.657 and 0.429 nm) under X-ray diffractograms. Platelets of average length less than 100 nm, interspersed in amorphous matrices, were observed in complexes of DPn 311 and 142, but that of DPn 39 showed different morphology, and the formation of complexes was limited. By hydrolyzing the complex of DPn 311 with α-amylase, amorphous matrices were selectively removed, and crystallites of 23–72 nm showing a V6-I X-ray diffraction pattern were obtained. However, crystallites in complexes of DPn 142 and 39 were eroded by amylolysis, forming large aggregates.  相似文献   

13.
Summary The investigation on hydrodynamic parameters of molybdate-stabilized glucocorticoid-receptor complexes from HeLa cell cytosol permitted resolution of four distinct forms. The first one could be detected in concentrated cytosols at low salt concentrations, and had the following properties: sedimentation coefficient = 9 S; R s = 9.3 nm; M r = 357,800; f/f o = 1.83; axial ratio (prolate ellipsoid) = 16. When these cytosol extracts were diluted, a second form could be detected with sedimentation coefficient = 8.3 S; R s = 9.05 nm; M r = 320,700;f/f o = 1.84; axial ratio = 16. Under high salt conditions, glucocorticoid-receptor complexes in concentrated cytosol had the following properties: sedimentation coefficient = 6.4 S; R s, = 6.7 nm; M r = 183,100;f/f o = 1.64; axial ratio = 12. When either these cytosol extracts were diluted, or glucocorticoid-receptor complexes were subjected to repeated analysis, a fourth form was detected with sedimentation coefficient = 3.76 S; R s = 5.67; M r = 91,000; f/f o = 1.75; axial ratio = 14. Besides salt concentration and dilution, the time elapsed between sample dilution and analysis appeared to affect the hydrodynamic properties of glucocorticoid-receptor complexes. On the basis of our findings, it has been concluded that the most likely structure of molybdate-stabilized glucocorticoid-receptor complexes of HeLa cell cytosol can be represented by association of monomers in homodimers, and homotetramers. A homotrimer form could not be deduced from our findings, and the 320,700 glucocorticoid-receptor complex we observed has been suggested to represent an unresolved mixture of trimers and tetramers.  相似文献   

14.
Abstact 3D-QSAR studies using the Comparative Molecular Field Analysis (CoMFA) methodology were conducted to predict the inhibition constants, Ki, and the inhibitor concentrations, IC90 of 127 symmetrical and unsymmetrical cyclic urea and cyclic cyanoguanidine derivatives containing different substituent groups such as: benzyl, isopropyl, 4-hydroxybenzyl, ketone, oxime, pyrazole, imidazole, triazole and having anti-HIV-1 protease activities. A significant cross-validated correlation coefficient (q2) of 0.63 and a fitted correlation coefficient r2 of 0.70 were obtained, indicating that the models can predict the anti-protease activity from poorly to highly active compounds reliably. The best predictions were obtained for: XV643 (predicted log 1/Ki=9.86), a 3,5-dimethoxy-benzyl cyclic urea derivate (molec60, predicted log 1/Ki=8.57) and a benzyl cyclic urea derivate (molec 61, predicted log 1/IC90=6.87). Using the CoMFA method, we also predicted the biological activity of 14 cyclic urea derivatives that inhibit the HIV-1 protease mutants V82A, V82I and V82F. The predicted biological activities of the: (i) XNO63 (inhibitory activity on the mutant HIV-1 PR V82A), (ii) SB570 (inhibiting the mutant HIV-1 PR V82I) and also (iii) XV652 (during the interaction with the mutant HIV-1 PR V82F) were in good agreement with the experimental values.Figure Stereoview of the contour plots of the CoMFA steric and electrostatic fields. The favorable (indicated by blue polyhedra) and unfavorable (represented by red polyhedra) electrostatic areas and also the favorable (shown by green polyhedra) and unfavorable (shown by yellow polyhedra) steric areas formed around the most active molecule, 6a.  相似文献   

15.
In an age-structured population that grows exponentially, each age groupP i(t) at periodt is asymptotically equivalent tox 0 t for some positive number x0. In this paper we show that the speed at which the ith age group reaches its exponential state of equilibrium can be measured by the rate at which the ratio vi(t)=Pi(t)/pi(t–1) converges tox 0. The age specific rate of convergence is determined by considering a quantityr satisfyingv i(t)-x 0 ¦ r t whent is large;R i=Infr (over all initial populations,r satisfying the above inequality) is the R-factor used in numerical analysis to measure the rate at which the sequencev i (t) converges tox 0;S i =- In Ri is then defined as the rate of convergence to stability of the ith age group. The case of constant net maternity rates is studied in detail; in this contextS 0 is compared to the population entropyH, which was proposed by Tuljapurkar (1982) as a measure of the rate of convergence to stability.  相似文献   

16.
The homeostasis of intracellular pH (pHi) affects many cellular functions. Our previous study has established a functional and molecular model of the active pHi regulators in human induced pluripotent stem cells (hiPSCs). The aims of the present study were to further quantify passive pHi buffering power (β) and to investigate the effects of extracellular pH and Na+–H+ exchanger 1 (NHE1) activity on pluripotency in hiPSCs. pHi was detected by microspectrofluorimetry with pH‐sensitive dye‐BCECF. Western blot, immunofluorescence staining, and flow cytometry were used to detect protein expression and pluripotency. Our study in hiPSCs showed that (a) the value of total (βtot), intrinsic (βi), and CO2‐dependent () buffering power all increased while pHi increased; (b) during the spontaneous differentiation for 4 days, the β values of βtot and changed in a tendency of decrease, despite the absence of statistical significance; (c) an acidic cultured environment retained pluripotency and further upregulated expression and activity of NHE1 during spontaneous differentiation; (d) inhibition on NHE1 activity promoted the loss of pluripotency. In conclusion, we, for the first time, established a quantitative model of passive β during differentiation and demonstrated that maintenance of NHE1 at a higher level was of critical importance for pluripotency retention in hiPSCs.  相似文献   

17.
Kinetics of the reactions of purine nucleoside phosphorylases (PNP) from E. coli (PNP-I, the product of the deoD gene) and human erythrocytes with their natural substrates guanosine (Guo), inosine (Ino), a substrate analogue N(7)-methylguanosine (m7Guo), and orthophosphate (Pi, natural cosubstrate) and its thiophosphate analogue (SPi), found to be a weak cosubstrate, have been studied in the pH range 5–8. In this pH range Guo and Ino exist predominantly in the neutral forms (pKa 9.2 and 8.8); m7Guo consists of an equilibrium mixture of the cationic and zwitterionic forms (pKa 7.0); and Pi and SPi exhibit equilibria between monoanionic and dianionic forms (pKa 6.7 and 5.4, respectively). The phosphorolysis of m7Guo (at saturated concentration) with both enzymes exhibits Michaelis kinetics with SPi, independently of pH. With Pi, the human enzyme shows Michaelis kinetics only at pH ∼5. However, in the pH range 5–8 for the bacterial enzyme, and 6–8 for the human enzyme, enzyme kinetics with Pi are best described by a model with high- and low-affinity states of the enzymes, denoted as enzyme-substrate complexes with one or two active sites occupied by Pi, characterized by two sets of enzyme-substrate dissociation constants (apparent Michaelis constants, K m1 and K m2) and apparent maximal velocities (V max1 and V max2). Their values, obtained from non-linear least-squares fittings of the Adair equation, were typical for negative cooperativity of both substrate binding (K m1 < K m2) and enzyme kinetics (V max1/K m1 > V max2/K m2). Comparison of the pH-dependence of the substrate properties of Pi versus SPi points to both monoanionic and dianionic forms of Pi as substrates, with a marked preference for the dianionic species in the pH range 5–8, where the population of the Pi dianion varies from 2 to 95%, reflected by enzyme efficiency three orders of magnitude higher at pH 8 than that at pH 5. This is accompanied by an increase in negative cooperativity, characterized by a decrease in the Hill coefficient from n H ∼1 to n H ∼0.7 for Guo with the human enzyme, and to n H ∼0.7 and 0.5 for m7Guo with the E. coli and human enzymes, respectively. Possible mechanisms of cooperativity are proposed. Attention is drawn to the substrate properties of SPi in relation to its structure.  相似文献   

18.
Mcl-1 has emerged as a potential therapeutic target in the treatment of several malignancies. Peptides representing BH3 region of pro-apoptotic proteins have been shown to bind the hydrophobic cleft of anti-apoptotic Mcl-1 and this segment is responsible for modulating the apoptotic pathways in living cells. Understanding the molecular basis of protein–peptide interaction is required to develop potent inhibitors specific for Mcl-1. Molecular dynamics simulations were performed for Mcl-1 in complex with three different BH3 peptides derived from Mcl-1, Bax, and Bim. Accordingly, the calculated binding free energies using MM-PBSA method are obtained and comparison with the experimentally determined binding free energies is made. The interactions involving two conserved charged residues (Aspi, and Arg/Lysi-4) and three upstream conserved hydrophobic residues (Leui-5, Ile/Vali-2, and Glyi-1, respectively) of BH3 peptides play the important roles in the structural stability of the complexes. The calculated results exhibit that the interactions of Bim BH3 peptides to Mcl-1 is stronger than the complex with Bax 19BH3 peptides. The hydrophobic residues (position i???9, i???8 and i?+?2) of BH3 peptides can be involved in their inhibitory specificity. The calculated results can be used for designing more effective MCL-1 inhibitors.  相似文献   

19.
20.
Intracellular free calcium ([Ca2+]i) is essential for oocyte maturation and early embryonic development. Here, we investigated the role of [Ca2+]i in oocytes from cumulus‐oocyte complexes (COCs) with respect to maturation and early embryonic development, using the calcium‐buffering agent BAPTA‐AM (1,2‐bis[2‐aminophenoxy]ethane‐N,N,N′,N′‐tetraacetic acid tetrakis [acetoxymethyl ester]). COCs were graded based on compactness of the cumulus mass and appearance of the cytoplasm, with Grade 1 indicating higher quality and developmental potential than Grade 3. Results showed that: (i) [Ca2+]i in metaphase‐II (MII) oocytes from Grade‐3 COCs was significantly higher than those from Grade‐1 COCs, and was significantly reduced by BAPTA‐AM; (ii) nuclear maturation of oocytes from Grade‐3 COCs treated with BAPTA‐AM was enhanced compared to untreated COCs; (iii) protein abundance of Cyclin B and oocyte‐specific Histone 1 (H1FOO) was improved in MII oocytes from Grade‐3 COCs treated with BAPTA‐AM; (iv) Ca2+ transients were triggered in each group upon fertilization, and the amplitude of [Ca2+]i oscillations increased in the Grade‐3 group upon treatment with BAPTA‐AM, with the magnitude approaching that of the Grade‐1 group; and (v) cleavage rates and blastocyst‐formation rates were improved in the Grade‐3 group treated with BAPTA‐AM compared to untreated controls following in vitro fertilization and parthenogenetic activation. Therefore, BAPTA‐AM dramatically improved oocyte maturation, oocyte quality, and embryonic development of oocytes from Grade‐3 COCs.  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