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 0 毫秒
1.
Measurements of differential scanning calorimetry (d.s.c.) have been made on the complex bovine serum albumin (BSA)--sodium dodecyl sulphate (SDS) under various conditions. There are two peaks P1 and P2 in the d.s.c. curve for BSA at pH 7 and in the absence of NaCl, indicating the presence of the heat-induced transition of BSA. There are three peaks P1, P2 and P3 in the curve for the system with the molar mixing ratio SDS/BSA = 1. With the increase in the amount of SDS, the peak P3 grows at the expense of P1 and P2. There is only a single peak P3 in the curve for the systems SDS/BSA > 7, and no peak at SDS/BSA = 50 and 100. There is a single peak P12 in the curve for BSA at pH 7 and in the presence of 0.05 M NaCl, indicating that the heat-induced transition is suppressed. There are two peaks P12 and P3 for the systems SDS/BSA = 1-5; the area ratio of the peak P3 to P12 increases with the increase in the amount of SDS. There is only a single peak P3 when SDS/BSA > 7, and no peak at SDS/BSA = 50. It is concluded that the peak P3 is a product of SDS regardless of the presence or absence of NaCl. Values of thermal denaturation temperature (Td) and enthalpy (delta H) of thermal denaturation indicate that the complex AD12 (A = BSA, D = SDS) is in the most thermostabilized state.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Using defatted and SH-blocked bovine serum albumin (BSA), measurements of differential scanning calorimetry (d.s.c.) have been made mainly in NaSCN solution. BSA undergoes a heat-induced conformational transition in a particular range of pH and ionic strength and is separated into two thermally independent units, each of which has different thermostability in acidic and alkaline pH regions. Comparisons were made of the pH dependencies of the enthalpy of thermal denaturation (ΔH) and the temperature of thermal denaturation (Td) in 0.01 NaSCN with those in 0.01 NaCl. It has been found that the stabilizing effect of NaSCN on BSA is larger than that of NaCl at pH 3.5–8, and that the heat-induced transition occurs by the electrostatic repulsive forces among the positively charged amino acid residues in a segment Arg 184–Arg 216 containing Trp 212 and the primary binding sites of anions. At ionic strength 0.01, the relative effectiveness of anions in suppressing the heat-induced transition and increasing the thermostability of BSA follows the order ClO4 − SCN > I > SO42− > Br > Cl. At ionic strength 0.1, the heat-induced transition is suppressed in all the salt solutions, and a Td increase follows the order ClO4 SCN > I > Br > Cl SO42−. Thus, the highly chaotropic ions thermostabilize BSA more markedly than kosmotropic ions in the low and moderate salt concentrations. In contrast, chaotropic ions destabilize BSA and kosmotropic ions stabilize BSA at the higher concentrations. An adequate amount of NaCl or NaSCN prevents the destruction of the environment of the binding site in the segment containing Trp 212 in 4 urea solution at pH 7.0.  相似文献   

3.
Reverse micelle extraction is still in the stage of laboratory. Major limitation associated with use of synthetic surfactants in reverse micelle extraction process is the unfolding or denaturation of proteins. Sugar surfactants are thought non-toxic and environmentally benign, and can exhibit interesting interfacial properties, but the application of sugar-based surfactants in protein extraction is still limited. In the present study, we extracted bovine serum albumin (BSA) by using reverse micelles from glucosylammonium (GA) and lactosylammonium (LA) surfactants (with dicarboxylate as counter ion). It was found that under optimum condition, (1) the maximum forward extraction efficiency was ca. 86% with GA, while only around 50% with LA, and (2) almost all BSA solubilized in reverse micelles prepared from GA could be recovered into aqueous phase, while the recovery of BSA from the reverse micelles of LA was lower. In addition, the optimum extraction parameters were closely related to surfactant structure. Therefore, the electrostatic interaction, H-bonding and sugar head size should be important for BSA transfer.  相似文献   

4.
The mechanism of interaction between bovine serum albumin (BSA) and 2-naphthylamine (2-NA) in aqueous solution was investigated by fluorescence spectroscopy, circular dichroism (CD) spectra, and UV-vis spectroscopy. It was proved from fluorescence spectra that the fluorescence quenching of BSA by 2-NA was a result of the formation of complex between 2-NA and BSA, and the binding constants (K(a) ) as well as the numbers of binding sites for 2-NA in BSA were determined according to the modified Stern-Volmer equation. The results of synchronous fluorescence and CD spectra demonstrated 2-NA could decrease the amount of α-helix of BSA, leading to the loosening of protein skeleton. UV-vis spectroscopy and resonance light scattering spectra (RLS) results also suggested the conformation of BSA were changed and the BSA aggregation occured, which could induce toxic effects on the organism.  相似文献   

5.
Thermal stability of Momordica charantia seed lectin (MCL) was investigated as a function of protein concentration, pH, scan rate, and at different ligand concentrations by using high-sensitivity differential scanning calorimetry (DSC). The DSC endotherm obtained at pH 7.4 consists of two entities with transition temperatures at ca. 333.7 K, and 338 K. The unfolding process is irreversible and could be described by a three-state model. For MCL tetramer ΔHcHv ratio is close to 4 for the first transition and ∼2 for the second transition, suggesting that four and two cooperative units are involved in the first and second transitions, respectively. In the presence of lactose both transitions shifted to higher temperatures, suggesting that ligand binds preferentially to the native conformation of MCL. Endotherms recorded as a function of pH indicate that MCL is more stable at lower pH. Chemical unfolding of MCL, induced by Gdn.HCl, was investigated by monitoring the intrinsic fluorescence properties of the protein. The results obtained indicate that chemical denaturation of MCL can also be described by a three-state process, involving an intermediate populated at ∼3–4 M Gdn.HCl. These observations suggest that the chemical and thermal unfolding processes are similar in that both of them proceed via an intermediate. The far UV and near UV CD spectra of MCL were nearly identical at different pH values and indicate that its secondary and tertiary structure do not change significantly with pH, suggesting that the structure of the protein is stable over a wide pH range.  相似文献   

6.
The nanosecond fluorescence depolarization method was applied to measure the fluorescence lifetime () and the rotational correlation time () of bovine serum albumin (BSA) labeled with 1-dimethylaminonaphthalene-5-sulfonyl chloride (dansyl-Cl). Changes of and of dansyl BSA in the guanidine denaturation and in the thermal denaturation were examined. In parallel, the secondary structural change of dansyl BSA was followed by circular dichroism measurements. The magnitude of was almost unchanged between 1 and 2 M guanidine, where the secondary structure of the protein was predominantly disrupted; whereas that of began to increase before the disruption of secondary structure in the guanidine denaturation. In the thermal denaturation, in contrast, changes of both and occurred in a temperature range where the secondary structure was predominantly disrupted. The volume of equivalent sphere (V e ) and the axial ratio () for the BSA were 3.6–3.8×10–19 cm3 and 3.6 at 2M guanidine as against 2.1×10–19 cm3 and 2.2 in the absence of guanidine (25°C), respectively. The magnitudes ofV e and were 4.9×10–19 cm3 and 4.5 at 65°C, respectively. Although the secondary structural change of dansyl BSA was irreversible in the thermal denaturation,V e and were reversible.  相似文献   

7.
The helicities in various fragments of bovine serum albumin (BSA) were examined in the thermal denaturation and in sodium docecyl sulfate (SDS) denaturation. The thermal denaturation was examined in a temperature range between 2 and 65°C. The helicity decreased with a rise of temperature and it recovered to some degree upon cooling temperature. A rather high reversibility was observed in the BSA fragments, which were located in the N-terminal of the parent protein and then contained the first large loop with no disulfide bridge. The high reversibility was available also for the helicity in the first large loop of the fragment, disulfide bridges of which were reduced. The fragments, which were smaller than one domain, became unstable in the SDS denaturation. The helicities of such fragments decreased in lower SDS concentrations compared with those of the intact BSA and the large fragments, which contained one or more domains. A resistance to the SDS denaturation appeared in the helices of every large loop even after the fragmentation. On the other hand, helicities of the fragments decreased to 20–25% upon the reduction of disulfide bridges. However, the helicities of these fragments increased to 35–40% in the SDS denaturation.  相似文献   

8.
Abstract

The binding of drugs to serum proteins is governed by weak non-covalent forces. In this study, the nature and magnitude of the interactions between piroxicam (PRX) and bovine serum albumin (BSA) was assessed using spectroscopic, calorimetric and computational molecular methods. The fluorescence data revealed an atypical behavior during PRX and BSA interaction. The quenching process of tryptophan (Trp) by PRX is a dual one (approximately equal static and dynamic quenched components). The FRET results indicate that a non-radiative transfer of energy occurred. The association constant and the number of binding sites indicate moderate PRX and BSA binding. The competitive binding study indicates that PRX is bound to site I from the hydrophobic pocket of subdomain IIA of BSA. The synchronous spectra showed that the microenvironment around the BSA fluorophores and protein conformation do not change considerably. The Trp lifetimes revealed that PRX mainly quenches the fluorescence of Trp-213 situated in the hydrophobic domain. The CD and DSC investigation show that addition of PRX stabilizes the protein structure. ITC results revealed that BSA-PRX binding involves a combination of electrostatic, hydrophobic and hydrogen interactions. The analysis of the computational data is consistent with the experimental results. This thorough investigation of the PRX-BSA binding may provide support for other studies concerning moderate affinity drugs with serum protein.

Communicated by Ramaswamy H. Sarma  相似文献   

9.
The mechanism of the interaction between bovine serum albumin (BSA) and desvenlafaxine was studied using fluorescence, ultraviolet absorption, 3‐dimensional fluorescence spectroscopy, circular dichroism, synchronous fluorescence spectroscopy, cyclic voltametry, differential scanning calorimetry, and attenuated total reflection–Fourier transform infrared spectroscopic techniques under physiological condition at pH 7.4. Stern‐Volmer calculations authenticate the fluorescence of BSA that was quenched by desvenlafaxine in a collision quenching mode. The fluorescence quenching method was used to evaluate number of binding sites “n” and binding constant K A that were measured, and various thermodynamic parameters were evaluated at different temperatures by using the van't Hoff equation and differential scanning calorimetry technique, which indicated a spontaneous and hydrophobic interaction between BSA and desvenlafaxine. According to the Förster theory we calculate the distance between the donor, BSA and acceptor, desvenlafaxine molecules. Furthermore, circular dichroism and attenuated total reflection–Fourier transform infrared spectroscopy indicate nominal changes in the secondary structure of the protein.  相似文献   

10.
The mechanism of formaldehyde-protein interactions was investigated by determining the effects of formaldehyde on the common protein bovine serum albumin (BSA). The effects at the molecular level were determined by fluorescence, ultraviolet absorption, and circular dichroism (CD) spectrometry. Formaldehyde could decrease the amount of alpha-helix, leading to loosening of the protein skeleton. In the loose structure, internal amino acids are exposed and the characteristic fluorescence of BSA is obviously quenched. The spectroscopic results reveal that formaldehyde exposure induces changes in the microenvironment and conformation of serum albumin, which could lead to toxic effects on the organism.  相似文献   

11.
《Process Biochemistry》2014,49(4):623-630
Herein, we report the effect of N,N-bis(dodecyloxycarbonylmethyl)-N,N,N,N-tetramethyl-1,2-ethanediammonium dibromide (dodecyl betainate gemini or DBG) on the structure and function of bovine serum albumin (BSA) by using fluorescence, time resolved fluorescence, circular dichroism and dynamic light scattering techniques. The Stern–Volmer quenching constants KSV and the corresponding thermodynamic parameters viz ΔH, ΔG and ΔS have been estimated by the fluorescence quenching method. The results indicated that DBG binds spontaneously with BSA through hydrophobic interaction. Time resolved fluorescence data show that the quenching follows the static mechanism pathway. It can be seen from far-UV CD spectra that the α-helical network of BSA is disrupted and its content increases from 71% to 79% at lower concentrations which again decreases to 38% at higher concentration. DLS measurements suggested that hydrodynamic radius (Rh) decreases in the presence of 30 and 40 μM of DBG while it increases when the concentration of DBG was 70 and 100 μM. The molecular docking study indicated that DBG is embedded into subdomain IIA of BSA and binds with the R-914, R-195 and R-217 residues by hydrogen bonding and by hydrophobic interaction.  相似文献   

12.
Some properties of a fragment of bovine serum albumin containing residues 184–582 of the protein sequence, produced by cyanogen bromide cleavage, have been reported. Urea-induced difference spectra of the fragment showed considerable exposure of aromatic chromophores by 8 M urea. Reversible unfolding of the fragment by urea, as followed by difference spectral measurements at 30°C, pH 7.0, occurred in two distinct steps involving at least 3 major conformational states, namely the native (N), intermediate (X) and completely denatured (D) states. The co-operativity values for the two transitions, N⇌X and X⇌Dwere found to be 4.0 and 16.4, respectively. Analysis of the data on bilirubin binding to bovine serum albumin and its fragment suggested that the fragment retains significant amount of its native structure. However, hydrodynamic parameters such as Stokes radius (3.f14 nm), diffusion coefficient (6.98 × 10−7cm2/s) and frictional ratio (1.32) obtained by analytical gel chromatography as well as intrinsic viscosity (4.31 ml/g) indicates some asymmetry in the fragment molecule.  相似文献   

13.
Serum albumin, a protein naturally abundant in blood plasma, shows remarkable ligand binding properties of numerous endogenous and exogenous compounds. Most of serum albumin binding sites are able to interact with more than one class of ligands. Determining the protein‐ligand interactions among mammalian serum albumins is essential for understanding the complexity of this transporter. We present three crystal structures of serum albumins in complexes with naproxen (NPS): bovine (BSA‐NPS), equine (ESA‐NPS), and leporine (LSA‐NPS) determined to 2.58 Å (C2), 2.42 Å (P61), and 2.73 Å (P212121) resolutions, respectively. A comparison of the structurally investigated complexes with the analogous complex of human serum albumin (HSA‐NPS) revealed surprising differences in the number and distribution of naproxen binding sites. Bovine and leporine serum albumins possess three NPS binding sites, but ESA has only two. All three complexes of albumins studied here have two common naproxen locations, but BSA and LSA differ in the third NPS binding site. None of these binding sites coincides with the naproxen location in the HSA‐NPS complex, which was obtained in the presence of other ligands besides naproxen. Even small differences in sequences of serum albumins from various species, especially in the area of the binding pockets, influence the affinity and the binding mode of naproxen to this transport protein. Proteins 2014; 82:2199–2208. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
Using defatted and SH-blocked bovine serum albumin (BSA), the measurement of differential scanning calorimetry (d.s.c.) was performed in the range pH 3-11 and ionic strength 0.001-1 M. The shape of the d.s.c. curve was classified into four regimes: (i) the curve with no peak, (ii) that with a peak, (iii) that with a peak having a shoulder, and (iv) that with two peaks. The presence of two peaks was interpreted by the concept of 'heat-induced transition'. The BSA molecule is composed of two domains, thermodynamically independent owing to the formation of a crevice in BSA in a particular range of pH and ionic strength; this gives two peaks in the d.s.c. curve. The enthalpy (delta H) from the d.s.c. curve was plotted against pH and against the NaCl concentration. The value of delta H increased with the increase in the ionic strength in the pH range 5.6-9.0. The temperature of thermal denaturation (the temperature of the peak maximum, Td) was raised with the increase in the ionic strength in the pH range 4.5-9.0, but was lowered in the pH range 3.5-4.0. BSA was stabilized in the neutral-alkaline pH range by the presence of NaCl, but was destabilized in the acidic pH range.  相似文献   

15.
Quinine (QN) and quinidine (QD), the chief quinoline alkaloids of various species of cinchona bark, are stereoisomers to each other. In this study, a series of appropriate and efficient methods have been applied to compare the binding modes of QN and QD with bovine serum albumin (BSA). The isothermal titration calorimetry and room temperature phosphorescence results show that both QN and QD can interact with BSA at one binding site to form drug–protein complexes, mainly through enthalpic driving force with the binding affinity order: QN > QD. The fluorescence resonance energy transfer and time‐resolved fluorescence spectroscopy exhibits that QN has a larger energy transfer and more intensified binding capacity for BSA than QD. Data of dynamic light scattering reveal that the aggregate state of BSA is changed during this binding process, and the particle size distribution of QN‐BSA bioconjugate is larger than that of QD. Nuclear magnetic resonance analysis indicates that aromatic protons make more contribution during ligand‐protein complexation than that of aliphatic protons. The circular dichroism spectra exhibit different degrees of changes in BSA secondary structures in the presence of QN and QD, respectively. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

16.
Thoppil AA  Sharma R  Kishore N 《Biopolymers》2008,89(10):831-840
Binding of the antibiotic drug carbenicillin to bovine serum albumin (BSA) has been studied using isothermal titration calorimetry (ITC) in combination with fluorescence and circular dichroism (CD) spectroscopies. The thermodynamic parameters of binding have been evaluated as a function of temperature, ionic strength, and in the presence of anionic, cationic and nonionic surfactants, tetrabutylammonium bromide, and sucrose. The values of van't Hoff enthalpy do not agree with the calorimetric enthalpy indicating conformational changes in the protein upon drug binding. These observations are supported by the intrinsic fluorescence and CD spectroscopic measurements. A reduction in the binding affinity of carbenicillin to BSA is observed with increase in ionic strength of the solution, thereby suggesting, prevailing of electrostatic interactions in the binding process. The involvement of hydrophobic interactions in the binding of the drug to the protein is also indicated by a slight reduction in binding constant in the presence of tetrabutylammonium bromide. The experiments in the presence of sucrose suggest that hydrogen bonding is perhaps not dominant in the binding. The anionic surfactant sodium dodecyl sulphate (SDS) is observed to completely interfere in the ionic interactions in addition to its partial denaturing capacity. However, the presence of cationic surfactant hexadecyl trimethylammonium bromide (HTAB) and nonionic surfactant Triton-X 100 induce a slight reduction in the values of binding affinity. These calorimetric and spectroscopic results, provide quantitative information on the binding of carbenicillin to BSA and suggests that the binding is dominated by electrostatic interactions with contribution from hydrophobic interactions. (c) 2008 Wiley Periodicals, Inc. Biopolymers 89: 831-840, 2008.This article was originally published online as an accepted preprint. The "Published Online" date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com.  相似文献   

17.
Lipid thermal transition patterns of the very low density lipoproteins in native and variously treated egg yolk plasma and extracted total very low density lipoproteins lipids have been recorded by differential scanning calorimetry in the temperature range 220–300 K, after lowering the freeze endotherm of free water in the sample with ethylene glycol. Three distinguishable patterns of lipid endotherms, designated types 1, 2 and 3 were obtained, respectively, from (i) native very low density lipoproteins in egg yolk plasma, (ii) freeze damaged very low density lipoproteins in gelled egg yolk plasma and (iii) extracted total lipids of very low density lipoproteins dispersed in water. Protein-depleted ‘lipid core’ particles of very low density lipoproteins obtained by exhaustive proteolysis of egg yolk plasma gave type 2 lipid transition pattern suggesting similarities in its lipid association with that of the freeze damaged very low density lipoproteins. Freezing the ‘lipid cores’ of very low density lipoproteins led to phase separation and gave type 3 lipid transition pattern of water-dispersed, phase-separated total very low density lipoprotein lipids. Relative heat uptake of native very low density lipoproteins in egg yolk plasma was about 15% lower than the freeze damaged sample or of the extracted total lipids. Treatments which prevented aggregation and gelation of very low density lipoproteins in egg yolk plasma during frozen storage, namely with additives such as glycerol or NaCl, gave subsequent lipid transition pattern intermediate between type 1 and 2, indicating that while very low density lipoprotein aggregation is prevented, additives do not altogether prevent changes in lipid association in these particles.  相似文献   

18.
The interaction of the cationic Gemini surfactant hexamethylene‐1,3‐bis (tetradecyldimethylammonium bromide) (14‐6‐14) with bovine serum albumin (BSA) has been investigated by fluorescence quenching spectra and three‐dimensional (3D) fluorescence spectra. The Stern–Volmer quenching constants KSV and the corresponding thermodynamic parameters ΔH, ΔG and ΔS have been estimated by the fluorescence quenching method. The results indicated that hydrophobic forces were the predominant intermolecular forces between BSA and the surfactant. Competitive experiments and the number of binding sites calculation show that 14‐6‐14 can be inserted in site‐II (in subdomain IIIA) of BSA. The effect of 14‐6‐14 on the conformation of BSA was evaluated by synchronous fluorescence spectroscopy and 3D fluorescence spectral methods. The results show that the conformation of BSA was changed dramatically in the presence of 14‐6‐14, by binding to the Trp and Try residues of BSA. The investigation provides interaction between BSA and 14‐6‐14 as a model for molecular design and industrial research. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

19.
Serum albumins being the most abundant proteins in the blood and cerebrospinal fluid are significant carriers of essential transition metal ions in the human body. Studies of copper (II) complexes have gained attention because of their potential applications in synthetic, biological, and industrial processes. Study of binding interactions of such bioinorganic complexes with serum albumins improves our understanding of biomolecular recognition process essential for rational drug design. In the present investigation, we have applied quantitative approach to explore interactions of novel synthesized copper (II) complexes viz. [Cu(L1)(L2)ClO4] (complex I), [Cu(L2)(L3)]ClO4] (complex II) and [Cu(L4)2(H2O)2] (complex III) with bovine serum albumin (BSA) to evaluate their binding characteristics, site and mode of interaction. The fluorescence quenching of BSA initiated by complexation has been observed to be static in nature. The binding interactions are endothermic driven by entropic factors as confirmed by high sensitivity isothermal titration calorimetry. Changes in secondary and tertiary structure of protein have been studied by circular dichroism and significant reduction in α-helical content of BSA was observed upon binding. Site marking experiments with warfarin and ibuprofen indicated that copper complexes bind at site II of the protein.  相似文献   

20.
Qasem RJ 《AAPS PharmSciTech》2006,7(1):E104-E110
The purpose of this study was to compare the effects of denaturation by microwave irradiation on release properties of 2 physically different proteins. Matrices were prepared from water-soluble bovine serum albumin loaded with metoclopramide and sorbed with adequate amount of moisture were thermally denatured in a microwave oven. The release profile of the rather insoluble denatured albumin matrices followed the classical Fickian diffusion profile. The release rate was dependent on the degree of denaturation, which was highly dependent on the level of moisture originally absorbed by the albuminoidal matrices and the period of exposure to microwave energy. Consersely, attempts to reduce the rate of drug release through microwave irradiation of metoclopramide-loaded matrices prepared from water-insoluble gluten were futile. The denaturation process was shown to be limited to the relatively water-soluble protein core fraction, while aggregation between neighboring gluten proteins in the matrix was not achieved even in the presence of considerable amounts of sorbed water. Published: February 10, 2006  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