首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A number of glycyl-tRNA synthetase (glyS) mutants have been isolated as glycine auxotrophs in Salmonella typhimurium. One of the mutants, glyS141, has a glycyl-tRNA synthetase with a Km for glycine that is 700 times higher than the wild-typeKm. Prototrophic revertants glyS141 occur at high spontaneous frequencies (>5 × 10?5). The majority of these revertants contain large tandem duplications including the mutant glyS gene. Some of the duplications cover at least 22% of the chromosome. The duplications overlap with a large duplication isolated previously by a different selection procedure (Straus &; Hoffmann, 1975). Evidence has been obtained which suggests that formation of the duplications may occur by recA-dependent recombination. The Gly+ phenotype of revertants carrying the duplications does not appear to be explainable simply by the increased gene dosage of glyS.  相似文献   

2.
残余牛血清白蛋白含量检测试剂盒抗干扰性研究   总被引:2,自引:0,他引:2  
为了对目前使用的残余牛血清蛋白(BSA)含量检测试剂盒的抗干扰性进行评价,选用19个企业的12个品种,共计28份样品进行检测,包括冻干疫苗和液体疫苗两种剂型。分别检测15ng/ml BSA对照样品、二倍稀释的疫苗样品和添加15ng/ml BSA的疫苗样品。将添加BSA的疫苗样品的检测结果减去未添加BSA的疫苗样品的结果,其数值应当位于BSA对照样品均值的95%可信区间内。多数品种的疫苗添加BSA后回收率在85%和115%之间。个别制品的回收率在82%~83%之间。实验研究结果证明目前使用的BSA检测试剂盒具有较好的抗干扰作用。  相似文献   

3.
In this work, we have studied the interaction between the anticancer drug doxorubicin (doxo) and condensed DNA, using optical tweezers. To perform this task, we use the protein bovine serum albumin (BSA) in the working buffer to mimic two key conditions present in the real intracellular environment: the condensed state of the DNA and the abundant presence of charged macromolecules in the surrounding medium. In particular, we have found that, when doxo is previously intercalated in disperse DNA, the drug hinders the DNA condensation process upon the addition of BSA in the buffer. On the other hand, when bare DNA is firstly condensed by BSA, doxo is capable to intercalate and to unfold the DNA condensates at relatively high concentrations. In addition, a specific interaction between BSA and doxo was verified, which significantly changes the chemical equilibrium of the DNA–doxo interaction. Finally, the presence of BSA in the buffer stabilizes the double‐helix structure of the DNA–doxo complexes, preventing partial DNA denaturation induced by the stretching forces.  相似文献   

4.
5.
Shi XY  Cao H  Ren FL  Xu M 《化学与生物多样性》2007,4(12):2780-2790
The interaction between bovine serum albumin (BSA) and tinidazole (Tindamax; 1) in aqueous solution was investigated in detail by means of UV/VIS and fluorescence spectroscopy, as well as through resonance light-scattering (RLS) spectroscopy. The apparent binding constant and number of binding sites were determined at three different temperatures, as well as the average binding distances between 1 and the nearest amino acid residue(s) of BSA, as analyzed by means of F?rster's theory of non-radiation energy transfer. Compound 1 was found to quench the inner fluorescence of BSA by forming a tight 1:1 aggregate, based on both static quenching and non-radiation energy transfer. The entropy change upon complexation was positive, and the enthalpy change was negative, indicating that the observed spontaneous binding is mainly driven by electrostatic interactions.  相似文献   

6.
In response to apoptotic stimuli, the pro-apoptotic protein Bax inserts in the outer mitochondrial membrane, resulting in the formation of pores and the release of several mitochondrial components, and sealing the cell's fate. To study the binding of Bax to membranes, we used an in vitro system consisting of 50nm diameter liposomes prepared with a lipid composition mimicking that of mitochondrial membranes in which recombinant purified full-length Bax was inserted via activation with purified tBid. We detected the association of the protein with the membrane using fluorescence fluctuation methods, and found that it could well be described by an equilibrium between soluble and membrane-bound Bax and that at a high protein-to-liposome ratio the binding seemed to saturate at about 15 Bax proteins per 50nm diameter liposome. We then obtained structural data for samples in this saturated binding regime using small-angle neutron scattering under different contrast matching conditions. Utilizing a simple model to fit the neutron data, we observed that a significant amount of the protein mass protrudes above the membrane, in contrast to the conjecture that all of the membrane-associated Bax states are umbrella-like. Upon protein binding, we also observed a thinning of the lipid bilayer accompanied by an increase in liposome radius, an effect reminiscent of the action of antimicrobial peptides on membranes.  相似文献   

7.
BACKGROUND: The classic model of estrogen action requires that the estrogen receptor (ER) activates gene expression by binding directly or indirectly to DNA. Recent studies, however, strongly suggest that ER can act through nongenomic signal transduction pathways and may be mediated by a membrane bound form of the ER. Estradiol covalently linked to membrane impermeable BSA (E2-BSA) has been widely used as an agent to study these novel membrane-associated ER events. However, a recent report suggests that E2-BSA does not compete for E2 binding to purified ER in vitro. To resolve this apparent discrepancy, we performed competition studies examining the binding of E2 and E2-BSA to both purified ER preparations and ER within intact cells. To eliminate potential artifacts due to contamination of commercially available E2-BSA preparations with unconjugated E2 (usually between 3-5%), the latter was carefully removed by ultrafiltration. RESULTS: As previously reported, a 10-to 1000-fold molar excess of E2-BSA was unable to compete with 3H-E2 binding to ER when added simultaneously. However, when ER was pre-incubated with the same concentrations of E2-BSA, the binding of 3H-E2 was significantly reduced. E2-BSA binding to a putative membrane-associated ER was directly visualized using fluorescein labeled E2-BSA (E2-BSA-FITC). Staining was restricted to the cell membrane when E2-BSA-FITC was incubated with stable transfectants of the murine ERalpha within ER-negative HeLa cells and with MC7 cells that endogenously produce ERalpha. This staining appeared highly specific since it was competed by pre-incubation with E2 in a dose dependent manner and with the competitor ICI-182,780. CONCLUSIONS: These results demonstrate that E2-BSA does bind to purified ER in vitro and to ER in intact cells. It seems likely that the size and structure of E2-BSA requires more energy for it to bind to the ER and consequently binds more slowly than E2. More importantly, these findings demonstrate that in intact cells that express ER, E2-BSA binding is localized to the cell membrane, strongly suggesting a membrane bound form of the ER.  相似文献   

8.
Small-angle neutron scattering and contrast variation were used to study the solution structure of GroEL and GroEL/GroES chaperonins complexed with a nonnative variant of the polypeptide substrate, subtilisin (PJ9). The subtilisin was 86% deuterated (dPJ9) so that it contrasted sufficiently with the chaperonin, allowing the contrast variation technique to be used to separate the scattering from the two components bound in the complex. Both the native double-ring GroEL and a single-ring mutant were used with dPJ9 bound in a 1:1 stoichiometry per GroEL toroid. This allowed both the position and the shape of dPJ9 in the GroEL/dPJ9 complexes to be determined. A single-ring GroEL/GroES variant complexed with one dPJ9 molecule was used to study the structural changes of dPJ9 in GroEL/GroES/dPJ9 complexes formed with ADP and with ATP. It was found that both the shape and the position of the bound dPJ9 in the GroEL/GroES/dPJ9 complex with ADP were the same as those in the GroEL/dPJ9 complex. However, dPJ9 assumed a more symmetric shape when bound in the GroEL/GroES/dPJ9 complex with ATP. This important observation reflects the relative ability of ATP to promote refolding of protein substrates relative to that of ADP.  相似文献   

9.
The fluorescence quenching spectrum of bovine serum albumin (BSA) was investigated in the presence of felodipine (FLD) by spectroscopic methods including fluorescence spectroscopy and UV–Vis absorption spectroscopy. Stern–Volmer quenching was successfully applied and the corresponding thermodynamic parameters, namely enthalpy change (ΔH), free energy change (ΔG) and entropy change (ΔS) at different temperatures (304, 314 and 324 K) were calculated according to the Van't Hoff relation. This revealed that the hydrophobic interaction plays a major role in stabilizing the complex. The fluorescence spectrum of BSA was studied in presence of various concentrations of SDS surfactant. The distance (r) between donor (BSA) and acceptor (FLD) was obtained according to fluorescence resonance energy transfer (FRET). The synchronous fluorescence spectroscopy was used to investigate the effect of FLD on BSA molecule. The result shows that the conformation of BSA was changed in the presence of felodipine. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

10.
A new microsensor application based on piezoresistive microcantilever technology has been used to study the interaction of anti-bovine serum albumin (a-BSA) with bovine serum albumin (BSA). A thin layer of BSA attached to a glass slide was used as the active sensing layer for the detection of a-BSA in solution. This design produced a large, consistent cantilever deflection when exposed to the analyte. In this system, the cantilever deflection is measured as a simple resistance change in the piezoresistive channel within the cantilever. In a second set of experiments, 3:1 BSA:PEO protein/polymer blended substrates were used as the active sensing layer for the detection of a-BSA in an aerosol delivery. A distinct signature for the analyte, separate from the water vapor carrier, is obtained for this system.  相似文献   

11.
The interaction between fasudil hydrochloride (FSD) and bovine serum albumin (BSA) was investigated using fluorescence and ultraviolet spectroscopy under imitated physiological conditions. The Stern–Volmer quenching model has been successfully applied and the results revealed that FSD could quench the intrinsic fluorescence of BSA effectively via static quenching. The binding constants and binding sites for the BSA–FSD system were evaluated. The corresponding thermodynamic parameters obtained at different temperatures indicated that hydrophobic force played a major role in the interaction of FSD and BSA. The distance between the donor (BSA) and the acceptor (FSD) was obtained according to fluorescence resonance energy transfer (FRET). Synchronous fluorescence spectroscopy and FT‐IR spectra showed that the conformation of BSA was changed in the presence of FSD. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

12.
Small-angle neutron scattering of the trans-unsaturated DEPC has been investigated as a function of pressure at 12, 18.6 and 35 degrees C. A pressure-induced structural phase transition from a liquid-crystalline state to a gel state is observed at the temperatures studied. The critical pressure of this transition increases with increasing temperature with a delta P/delta T value of 51 bar/C degrees. The small-angle neutron scattering results indicate that the effect of the trans double bonds in DEPC is to enhance the conformational disorder in the hydrocarbon chains. In DEPC bilayers, a pressure-induced conformational ordering process is observed not only in the liquid-crystalline phase but also in the gel phase, which indicates that conformational disorder exists in the liquid-crystalline phase as well as in the gel phase.  相似文献   

13.
14.
15.
The effect of the co-lyophilization of bovine serum albumin (BSA) with poly(ethylene glycol) (PEG) on the BSA encapsulation efficiency and formation of soluble BSA aggregates upon solid-in-oil-in-oil (s/o/o) encapsulation in poly(lactic-co-glycolic) acid (PLGA) microspheres was investigated. Suspension of the lyophilized BSA-PEG formulations in methylene chloride produced small protein powder particles of less than 1 m diam. and this afforded high encapsulation efficiencies of typically 90% ameliorating one of the problems in s/o/o encapsulation. Formation of soluble BSA aggregates upon s/o/o encapsulation followed by 24 h in vitro release was between 5% and 22%, much lower than values of 59% reported for BSA without stabilizing excipients. Therefore, PEG also afforded BSA stabilization during s/o/o encapsulation. Sustained release occurred over ca. 2 months and was complete.  相似文献   

16.
The binding of several different categories of small molecules to bovine (BSA) and human (HSA) serum albumins has been studied for many years through different spectroscopic techniques to elucidate details of the protein structure and binding mechanism. In this work we present the results of the study of the interactions of BSA and HSA with the anionic sodium dodecyl sulfate (SDS), cationic cethyltrimethylammonium chloride (CTAC) and zwitterionic N-hexadecyl-N,N-dimethyl-3-ammonium-1-propanesulfonate (HPS) monitored by fluorescence spectroscopy of the intrinsic tryptophans at pH 5.0. Similarly to pH 7.0 and 9.0, at low concentrations, the interaction of BSA with these surfactants shows a quenching of fluorescence with Stern-Volmer quenching constants of (1.1+/-0.1)x10(4) M(-1), (3.2+/-0.1)x10(3) M(-1) and (2.1+/-0.1)x10(3) M(-1) for SDS, HPS and CTAC, respectively, which are associated to the 'effective' association constants to the protein. On the interaction of these surfactants with HSA, an opposite effect was observed as compared to BSA, i.e., an enhancement of fluorescence takes place. For both proteins, at low surfactant concentrations, a positive cooperativity was observed and the Hill plot model was used to estimate the number of surfactant binding sites, as well as the association constants of the surfactants to the proteins. It is worthy of notice that the binding constants for the surfactants at pH 5.0 are lower as compared to pH 7.0 and 9.0. This is probably due to fact that the protein at this acid pH is quite compact reducing the accessibility of the surfactants to the hydrophobic cavities in the binding sites. The interaction of myristic acid with both proteins shows a similar fluorescence behaviour, suggesting that the mechanism of the interaction is the same. Recently published crystallographic studies of HSA-myristate complex were used to perform a modelling study with the aim to explain the fluorescence results. The crystallographic structure reveals that a total of five myristic acid molecules are asymmetrically bound in the macromolecule. Three of these sites correspond to higher affinity ones and correlate with high association constants described in the literature. Our models for BSA and HSA with bound SDS suggest that the surfactant could be bound at the same sites as those reported in the crystal structure for the fatty acid. The differences in tryptophan vicinity upon surfactant binding are explored in the models in order to explain the observed spectroscopic changes. For BSA the quenching is due to a direct contact of a surfactant molecule with the indole of W131 residue. It is clear that the binding site in BSA which is very close, in contact with tryptophan W131, corresponds to a lower affinity site, explaining the lower binding constants obtained from fluorescence studies. In the case of HSA the enhancement of fluorescence is due to the removal of static quenching of W214 residue in the intact protein caused by nearby residues in the vicinity of this tryptophan.  相似文献   

17.
The interaction of fractionated poly(acrylic acid)s (PAA) with bovine serum albumin (BSA) has been studied by measuring the hydrolysis rate of p-nitrophenyl acetate catalysed by BSA in the presence of PAA. The binding of PAA with BSA, which prohibits the catalytic action of BSA, increases with increasing molecular weight of PAA. The change in the electronic spectra of BSA-PAA solutions supports this molecular weight dependence. Circular dichroism of BSA shows that the binding of PAA does not induce any conformational change in BSA.  相似文献   

18.
The influence of bovine serum albumin (BSA) on the formation of J-aggregates of meso-tetra(4-sulfonatophenyl)porphine (TPPS4) in aqueous acid solution (pH 1.3) has been investigated by means of absorption and fluorescence spectroscopy. TPPS4 concentration was kept constant at 2 microM while BSA concentration was varied to get TPPS4 : BSA molar ratios from 1 : 0.005 to 1 : 20. In the presence of protein at all used concentrations the intensity of J-aggregates absorption band was higher than that in the pure solution. Spectral changes indicated that the dynamic equilibrium of the aggregated TPPS4 species was highly dependent on the molar ratio between TPPS4 and BSA. Small relative concentrations of BSA (TPPS4 : BSA, 1 : 0.005-1 : 0.1) had a stimulating effect on formation of J-aggregates. Several fractions of J-aggregates located in protein and aqueous moieties were detected in mixed solutions at intermediate BSA concentrations (TPPS4 : BSA, 1 : 0.5-1 : 8), when the absorbance intensity of the J-aggregates was the highest. At the highest used BSA concentrations (TPPS4 : BSA, 1 : 10-1 : 20) the spectral properties of the remaining J-aggregates were similar to those typical for pure porphyrin solution. Additionally, the split of the Soret band into two with peaks at 440 nm and 423 nm was followed by the simultaneous appearance of Q bands and reflected the formation of TPPS4-BSA complexes including both protonated and deprotonated TPPS4 forms.  相似文献   

19.
Although chiral anionic compounds, notably a large number of organic acids, have been found to be readily separated into enantiomers on BSA-based columns, the structural requirements for an efficient enantiomer discrimination by the protein is still not very well known. Since it is often observed that very hydrophobic acids, like many of the antiinflammatory “profens,” can be resolved with large separation factors for the enantiomers, a systematic study of a series of racemic α-substituted alkanoic acids was made. The series of analytes was prepared from α-amino acids, RCH(NH2)CO2H (where R = C1-C6), by reaction with N-(chloroformyl)-carbazole. A rapid increase in the capacity ratios of both enantiomers was found with increasing length of R. The effect, however, was larger for the last eluted enantiomer, leading to a substantial increase in the separation factor; this being 7.3 for R = C6 in 20 mM phosphate buffer (pH 8.0) with 30% of acetonitrile. Further, the separation factor also increased with decreasing organic modifier content. Thus when the R = C6-analyte was run at a mobile phase concentration of 20% acetonitrile and a flow rate of 1.5 ml/min, the time difference between the two eluted enantiomers exceeded 20 hr. A reasonable interpretation of our results seems to be that enantioselectivity is promoted by increased hydrophobic interaction. Since the anionic charge of the analyte is also taking part in the retention mechanism, a tight binding of the analyte will result from simultaneous electrostatic and hydrophobic interaction. When the latter is increased, less conformational freedom will be left for the analyte and the steric configuration at the α-carbon atom will become more and more important. Steric hindrance by the α-substituent in the first eluted enantiomer will counteract the tight binding caused by the combined binding interactions and lead to a smaller increase in the capacity ratio.  相似文献   

20.
Monodisperse solutions of bovine rhodopsin monomers, devoid of lipid, associated with a linear polyoxyethylene alcohol detergent have been prepared. The composition and homogeneity of these complexes have been determined by hydrodynamic characterisation. Each rhodopsin molecule is associated with about 110 monomers of the detergent. These rhodopsin-detergent complexes have been studied by small-angle neutron scattering. Partial or total deuteration of the detergent, as well as variation of the 2H2O/H2O ratio in the solvent, were used to eliminate the detergent—solvent contrast at various protein—solvent contrasts. The size and shape of the detergent micelle and of the rhodopsin-detergent complexes were shown to be independent of solvent or detergent deuteration. Mixture of selectively deuterated detergent molecules allowed us to obtain an homogeneous scattering density for the detergent part of the micelles and therefore to eliminate totally its contribution to the scattering when it is contrast matched. Neutron scattering from rhodopsin alone was then measured even in highly deuterated solvents, with low incoherent background, as for a water-soluble protein. Supplementary neutron scattering measurements on rhodopsin-dodecyl dimethylamine oxide micelles confirmed essentially the results reported by Yeager (1975). Analysis of the neutron scattering data indicates that most of the hydrophobic residues of rhodopsin form a compact region which has zero hydration, this probably being the part which is embedded in the disc membrane, and that the unhydrated rhodopsin molecule is asymmetrically arranged with respect to the membrane. Comparison with the results of a small-angle X-ray scattering study (Sardet et al., 1976) implies that the peripheral regions on both sides of the membrane are highly hydrated. Several schematic models are discussed.  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