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 281 毫秒
1.
Bovine serum albumin (BSA), as well as completely reduced BSA denoted by P (SH)35, are adsorbed on the hanging mercury drop electrode (HMDE) from alkaline buffer solutions. When time is allowed, a monolayer is adsorbed from very dilute (10?9M) BSA solutions in ammoniacal and borate buffers. With a monolayer of adsorbed protein the voltammograms at the HMDE are then identical in a given ammoniacal or borax buffer containing cobalt(III) or (II) and different BSA concentrations. Voltammograms of P (SH)35 are virtually identical with those of native BSA. At the HMDE the second Brdi?ka current is proportional to concentration of cobalt(III) or (II) and the first current nearly so. Incompletely or completely adsorbed BSA or P (SH)35 is not desorbed on keeping the HMDE for one hour in ammonia buffers. An incomplete layer of adsorbed BSA or P (SH)35 is relatively rapidly desorbed at ?1.6 V (vs. SCE) and a complete film at ?1.65 V, some desorption occurring at ?1.6 V. Upon desorption, the second Brdi?ka current decreases faster than the first one; this is particularly striking in 1 M ammonia buffer. The rate of desorption is increased by calcium chloride, but the rate of adsorption is not, or only slightly, increased in the presence of calcium. Incomplete adsorption occurs at ?1.60 V (vs. SCE) and no adsorption at ?1.65 V. Indications are obtained that “presodium currents” yield a slight plateau at ?1.67 to ?1.70 V, the plateau currents being attributed to adsorbed BSA, while unadsorbed BSA yields catalytic currents without a plateau, the currents merging with the residual one of the buffer. Calcium chloride greatly increases the presodium currents. From many kinetic data obtained at the dropping mercury electrode (DME) and from results at the HMDE it is concluded that, depending on the BSA concentration, Brdi?ka currents at the DME are partly of a kinetic and partly of a surface adsorption nature and partly diffusion-controlled. Adsorption equilibrium is not attained at the DME at 25° at concentrations of BSA smaller than 10?6M.  相似文献   

2.
Constant current chronopotentiometric peak H at mercury electrodes was recently shown as a sensitive tool for global and local changes in protein conformation [1]. Large differences between the heights of peak H of native (hBSAnat) and denatured BSA (hBSAden) were observed. The ratio hBSAden/hBSAnat increased with more negative stripping current suggesting that the rate of potential change is important for discrimination between native and denatured BSA. Voltammetric peaks of BSA were less well developed and BSAden/BSAnat was much smaller. It was not possible to discriminate BSAden and BSAnat using carbon electrodes.  相似文献   

3.
When the hanging mercury drop electrode (HMDE) is placed in a solution which is 0.1 M in ammonia and 0.1 M in ammonium chloride and about 5 to 10×10?4M in cobalt(III)-hexamine or cobalt(II) chloride and in very small concentrations of bovine serum albumin (BSA), the protein is slowly adsorbed. When the adsorption is highly incomplete and the HMDE is kept for 30 s at about ?1.05 V vs. SCE, “active cobalt’ is deposited as a complex (Co(0)BSA). This is anodically oxidized at about 0.0 V to unstable Co(I)BSA). When the electrode is then rapidly (500 mV s?1) cathodized, a catalytic hydrogen current (ic) with peak at circa ?1.45 V is observed. In this way it is even possible to detect and estimate BSA in concentrations of the order of 10?12M. A detailed study has been made of the characteristics of ic under several conditions. “Active cobalt” on the HMDE does not affect Brdi?ka currents. Cystine and cysteine also yield the catalytic hydrogen current ic under the same conditions as does BSA.  相似文献   

4.
In this work, open-tubular capillary electrochromatography (OT-CEC) method with bare gold nanoparticles (GNPs)-based stationary phase has been developed and applied for separation of tryptic peptide fragments of native and glycated proteins, bovine serum albumin (BSA), and human transferrin (HTF). The GNPs-based stationary phase was prepared by immobilization of bare GNPs, freshly reduced from tetrachloroaurate(III) ions by citrate reduction, on the sol-gel pretreated inner wall of the fused silica capillary. The separation efficiency, peak capacity, and peptide recovery of this open-tubular capillary column were investigated by varying the experimental parameters such as type and concentration of the buffering constituent and pH of the background electrolyte (BGE), temperature, and separation voltage. The best separations of the above tryptic peptides were achieved in the BGE composed of aqueous 100 mmol/L sodium phosphate buffer, pH 2.5, at separation voltage 10 kV per 47-cm long, 50 μm inside diameter capillary thermostated at 25°C. OT-CEC with bare GNPs stationary phase is shown to be a suitable technique for separation of complex peptide mixtures arising from tryptic digestion of native and glycated BSA and HTF, and for investigation of glycation (nonenzymatic glycosylation) of these proteins.  相似文献   

5.
Sun W  Jiao K 《Talanta》2002,56(6):1073-1080
In this paper, the electrochemical behavior of the interaction of Alizarin Red S (ARS) with bovine serum albumin (BSA) was investigated on the hanging mercury drop electrode (HMDE). In the acidic solution (pH 4.2), ARS can be easily reduced on the HMDE and it has a well-defined polarographic wave at −0.29 V (SCE). On addition of BSA or human serum albumin (HAS) into the ARS solution, the reduction peak current of ARS decreases without the movement of the peak potential and the appearance of new peaks. The study shows that a new electrochemically non-active complex is formed via intercalation of ARS with BSA or HSA, which can not be reduced on the Hg electrode. The decrease of reductive peak current of ARS is proportional to BSA and HSA concentration in the range of 2.0–60 and 2.0–40 mg l−1, respectively. The detection limit of BSA and HSA is 1.0-mg l−1. The analytical results of human serum and urine samples by this method were in good agreement with the Coomassie brilliant blue G-250 assay. The binding number and the binding interaction mechanism are also discussed.  相似文献   

6.
Open‐tubular CEC (OT‐CEC) with a new stationary phase, salophene–lanthanide–Zn2+ complex, has been applied to the separation of tryptic peptides of native BSA and BSA glycated by glucose and ribose. Glycation of proteins (non‐enzymatic modification by sugars) significantly affects their properties and it is of great importance from a physiological point of view. Separation of tryptic peptides of glycated BSA by CZE was poor because of their strong adsorption to the bare fused silica capillary. An improved separation of tryptic peptides of both native and glycated BSA was achieved by OT‐CEC in the fused silica capillary non‐covalently coated with salophene–lanthanide–Zn2+ complex, which suppressed the adsorption of peptides to the capillary and via specific interactions with some (glyco)peptides enhanced selectivity of the separation. Significant differences have been found in OT‐CEC analyses of tryptic hydrolysates of native and glycated BSA. In OT‐CEC‐UV profile of tryptic peptides of native BSA, 44 peaks could be resolved, whereas a reduced number of 38 peaks were observed in the profile of tryptic peptides of glucose‐glycated BSA and only 30 peaks were found in the case of ribose‐glycated BSA. The developed OT‐CEC can be potentially used for monitoring of protein glycation.  相似文献   

7.
The interaction between metal ions and bovine serum albumin (BSA) was studied by using a piezoelectric quartz crystal (PQC) arranged in the electrode-separated configuration. A silanized surface of the PQC was coated with a BSA membrane via a coupling reaction with glutaraldehyde. The frequency shifts obtained from PQC coated with a BSA membrane suggested that various kinds of metal ions could be adsorbed onto the BSA membrane from aqueous solutions containing a low concentration of metal ions (2 or 10 micromol dm(-3)), only when the BSA was denatured with an alkaline solution. Anionic species of Pt(IV) and Au(III) were adsorbed onto the denatured BSA membrane from an acetic acid solution at pH 2.2, and cationic species of Cd(II), Zn(II), Co(II), Ni(II), Cu(II), and Ag(I), and cations, such as Ca2+, Ba2+, and Mg2+, were adsorbed from ammonia buffer at pH 9.5, whereas Al(III), Cr(III), Fe(III), Hg(II), and Pb(II) were hardly adsorbed. The adsorption mechanisms of these metal ions are discussed, based on the electrostatic interaction between the metal ions and the denatured BSA membrane, and complex formation between the metal ions and amino acid residues of the denatured BSA. Further, the PQC coated with a denatured BSA membrane was applied to the determination of Pt and Cd, using large frequency shifts for Pt(IV) and Cd(II).  相似文献   

8.
Described here are the electrochemical parameters for MB on binding to DNA at hanging mercury drop electrode (HMDE), glassy carbon electrode (GCE), and carbon paste electrode (CPE) in the solution and at the electrode surface. MB, which interacts with the immobilized calf thymus DNA, was detected by using single-stranded DNA-modified HMDE or CPE (ssDNA-modified HMDE or CPE), bare HMDE or CPE, and double-stranded DNA-modified HMDE or CPE (dsDNA-modified HMDE or CPE) in combination with adsorptive transfer stripping voltammetry (AdTSV), differential pulse voltammetry (DPV), and alternating current voltammetry (ACV) techniques. The structural conformation of DNA and hybridization between synthetic peptide nucleic acid (PNA) and DNA oligonucleotides were determined by the changes in the voltammetric peak of MB. The PNA and DNA probes were also challenged with excessive and equal amount of noncomplementary DNA and a mixture that contained one-base mismatched and target DNA. The partition coefficient was also obtained from the signal of MB with probe, hybrid, and ssDNA-modified GCEs. The effect of probe, target, and ssDNA concentration upon the MB signal was investigated. These results demonstrated that MB could be used as an effective electroactive hybridization indicator for DNA biosensors. Performance characteristics of the sensor are described, along with future prospects.  相似文献   

9.
It was originally shown [10] that urease retains its enzymatic activity when adsorbed at bare mercury and solid amalgam surfaces. However the opinion later prevailed that, when adsorbed at bare metal electrodes, proteins are irreversibly denatured. Here we confirm that urease is enzymatically active at a bare solid amalgam surface as found by Santhanam et al., and we show that this enzyme is equally active at a thiol-modified amalgam surface. We also show that it is the reduced form of urease, which is enzymatically active at Hg surfaces. Oxidation of the protein, resulting in formation of disulfide bonds, strongly decreases the enzyme activity. Using constant current chronopotentiometric stripping (CPS) we show that the exposure of surface-attached urease to negative potentials results in the protein unfolding. The extent of the unfolding depends upon the amount of time for which the protein is exposed to negative potentials, and at very short times this unfolding can be avoided. At thiol-modified Hg surfaces the protein is less vulnerable to the effects of the electric field. We conclude that the loss of enzymatic activity, resulting from a 10 min exposure of the protein to −0.58 V, is not due to reduction of the disulfide bonds as suggested by Santhanam et al. This loss is probably a result of protein reorientation, due to reduction of the Hg-S bonds (formed by accessible cysteines), followed by prolonged electric field effect on the surface-attached protein.  相似文献   

10.
The self-complementary decamer d(CCAGGCCTGG), native and denatured calf thymus DNAs were studied by means of differential pulse polarography with the DME and by cyclic voltammetry with the HMDE. The decamer (which represents one turn of the DNA double helix in the B form) produced cathodic and anodic signals similar to those yielded by high-molar mass DNAs. By measuring the anodic peak (due to guanine residues), it was possible to detect the decamer at subnanomolar concentrations by adsorptive stripping cyclic voltammetry at relatively short waiting times. The high sensitivity of the electrochemical analysis for small changes in the DNA double helix observed earlier in high-molar mass DNAs, together with the low requirements for the amount of the analyzed decamer sample, suggest that electrochemical techniques may become useful also in oligonucleotide studies.  相似文献   

11.
By method of differential scanning calorimetry the absolute values of heat capacity for the systemwater–biopolymer (globular and fibrillar proteins and DNA) were measured in a wide range of temperatures (from -30 up to 130°C) and concentrations of proteins both in native and denatured states. Thermal properties of humid denatured biopolymers demonstrate a characteristic anomaly in the form of the heat capacity jump at temperature depending on the bound water content. It has been shown that in the systems studied a glass transition, where water serves as a native plasticizer, is observed. It has been established that the S-shaped character of all heat capacity curves obtained on dehydration for native and denatured biopolymers is due to the gradual transition to the glassy state of both native and denatured samples. It was found that thermally denatured humid small globular proteins at subsequent dissolving in water at room temperature are able to restore their native structure. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

12.
The effect of glucose (0–15 mass%) on the kinetics of bovine serum albumin (BSA) denatured aggregation at high concentration in aqueous solution has been studied by differential scanning calorimetry. The observed denatured aggregation process was irreversible and could be characterized by a denaturation temperature (T m), apparent activation energy (E a), the approximate order of reaction, and pre-exponential factor (A). As the glucose concentration increased from 0 to 15 mass%, T m increased, E a also increased from 514.59409±6.61489 to 548.48611±7.81302 kJ mol−1, and A/s−1 increased from 1.24239E79 to 5.59975E83. The stabilization increased with an increasing concentration of glucose, which was attributed to its ability to alter protein denatured aggregation kinetics. The kinetic analysis was carried out using a composite procedure involving the iso-conversional method and the master plots method. The iso-conversional method indicated that denatured aggregation of BSA in the presence and absence of glucose should conform to single reaction model. The master plots method suggested that the simple order reaction model best describe the process. This study shows the combination of iso-conversional method and the master plots method can be used to quantitatively model the denatured aggregation mechanism of the BSA in the presence and absence of glucose.  相似文献   

13.
One major application of surfactants is to prevent aggregation during various processes of protein manipulation. In this work, a bacterial trehalose lipid (TL) with biosurfactant activity, secreted by Rhodococcus sp., has been identified and purified. The interactions of this glycolipid with selected model proteins have been studied by using differential scanning calorimetry (DSC), Fourier-transform infrared (FTIR) spectroscopy, isothermal titration calorimetry (ITC), and fluorescence spectroscopy. Bovine serum albumin (BSA) and cytochrome c (Cyt-c) have been chosen because of their quite different secondary structures: BSA contains essentially no β-sheets and an average 66% α-helix, whereas Cyt-c possesses up to 25% β-sheets and up to 45% α-helical structure. Differential scanning calorimetry shows that addition of TL to BSA at concentrations below the critical micelle concentration (cmc) shifts the thermal unfolding temperature to higher values. FTIR indicates that TL does not alter the secondary structure of native BSA, but the presence of TL protects the protein toward thermal denaturation, mainly by avoiding formation of β-aggregates. Studies on the intrinsic Trp fluorescence of BSA show that addition of TL to the native protein results in conformational changes. BSA unfolding upon thermal denaturation in the absence of TL makes the Trp residues less accessible to the quencher, as shown by a decrease in the value of Stern-Volmer dynamic quenching constant, whereas denaturation in the presence of the biosurfactant prevents unfolding, in agreement with FTIR results. In the case of Cyt-c, interaction with TL gives rise to a new thermal denaturation transition, as observed by DSC, at temperatures below that of the native protein, therefore facilitating thermal unfolding. Binding of TL to native BSA and Cyt-c, as determined by ITC, suggests a rather nonspecific interaction of the biosurfactant with both proteins. FTIR indicates that TL slightly modifies the secondary structure of native Cyt-c, but protein denaturation in the presence of TL results in a higher proportion of β-aggregates than in its absence (20% vs 3.9%). The study of Trp fluorescence upon TL addition to Cyt-c results in a completely opposite scenario to that described above for BSA. In this case, addition of TL considerably increases the value of the dynamic quenching constant, both in native and denatured protein; that is, the interaction with the glycolipid induces conformational changes which facilitate the exposure of Trp residues to the quencher. Considering the structures of both proteins, it could be derived that the characteristics of TL interactions, either promoting or avoiding thermal unfolding, are highly dependent on the protein secondary structure. Our results also suggest the rather unspecific nature of these interactions. These might well involve protein hydrophobic domains which, being buried into the protein native structures, become exposed upon thermal unfolding.  相似文献   

14.
Constant current chronopotentiometric stripping (CPS) peak H due to catalytic hydrogen evolution reaction on Hg‐containing electrodes appeared useful in the analysis of protein complexes with single‐stranded and double‐stranded DNA as well as with peptides. In dependence on stripping current (Istr), structural transition of the protein alone or in complexes can be followed as a result of the protein exposure to electric field effects. For the first time we show here that the CPS analysis can be used for the study of the interaction of BSA with a polysaccharide namely sodium alginate (SA). BSA‐SA complex formation was accompanied by the shift of the structural transition of BSA to lower ‐Istr intensities. Another polysaccharide dextran did not alter Istr‐dependent structural transition of BSA. BSA‐SA complex can be disturbed by an electric field effect or high ionic strength confirming the electrostatic nature of BSA‐SA interaction.  相似文献   

15.
We have studied hydration (and dehydration) of thermally denatured hen egg lysozyme using sorption calorimetry. Two different procedures of thermal denaturation of lysozyme were used. In the first procedure the protein was denatured in an aqueous solution at 90 degrees C, in the other procedure a sample that contained 20% of water was denatured at 150 degrees C. The protein denatured at 90 degrees C showed very similar sorption behavior to that of the native protein. The lysozyme samples denatured at 150 degrees C were studied at several temperatures in the range of 25-60 degrees C. In the beginning of sorption, the sorption isotherms of native and denatured lysozyme are almost identical. At higher water contents, however, the denatured lysozyme can absorb a greater amount of water than the native protein due to the larger number of available sorption sites. Desorption experiments did not reveal a pronounced hysteresis in the sorption isotherm of denatured lysozyme (such hysteresis is typical for native lysozyme). Despite the unfolded structure, the denatured lysozyme binds less water than does the native lysozyme in the desorption experiments at water contents up to 34 wt %. Glass transitions in the denatured lysozyme were observed using both differential scanning calorimetry and sorption calorimetry. Partial molar enthalpy of mixing of water in the glassy state is strongly exothermic, which gives rise to a positive temperature dependence of the water activity. The changes of the free energy of the protein induced by the hydration stabilize the denatured form of lysozyme with respect to the native form.  相似文献   

16.
Many prokaryotic and eukaryotic proteins are modified by post‐translational conjugation to short‐chain poly[(R)‐3‐hydroxybutyrate] (cPHB). The relative lability of ester bonds raises the concern that the cPHB may be substantially degraded by chemical hydrolysis during protein purification, thus increasing the difficulty of its detection and measurement. Here, we compare rates of acid‐ and base‐catalyzed hydrolysis of cPHB conjugated to native and denatured proteins at room temperature. E. coli cytoplasmic proteins, native or denatured by addition of guanidium hydrochloride, were treated with aqueous solutions of H2SO4 or NaOH at concentrations ranging from 0.1–2.0n . The loss of cPHB was measured as a function of time by a chemical assay. We find that cPHB conjugated to native proteins is surprisingly resistant to both acid‐ and base‐catalyzed hydrolysis, whereas cPHB conjugated to denatured proteins is proficiently degraded at rates proportional to acid or base concentration. The results suggest that cPHB occupies a highly protective environment within native proteins.  相似文献   

17.
The reversibility of the adsorption-desorption cycle was established by comparing the thermostability (determined by differential scanning calorimetry) and secondary structure (obtained by circular dichroism spectroscopy) of BSA before adsorption, adsorbed on, and exchanged from silica particles. Circular dichroism was also measured as a function of temperature at a given wavelength. Adsorbed BSA presents a higher thermostability and a lower alpha-helix content than the native protein while it regains its conformation when released from the surface back into the solution; the homomolecular exchange is reversible.The changes in ellipticity (at a given wavelength) as a function of the temperature show that the thermal denaturation of native, adsorbed, and exchanged BSA proceeds in two steps. For the dissolved protein, the first step up to 50 degrees C involves a slight change in the structure while in the 50-90 degrees C temperature range the actual unfolding takes place. For the adsorbed BSA, the first step proceeds up to 60 degrees C and includes some intermolecular association between the adsorbed protein molecules, which may be responsible for the increased thermostability. The unfolding occurs in the 60-90 degrees C range; it is less cooperative and involves a lower enthalpy change than the native protein. Adsorbed BSA presents the same secondary structure as that observed for dissolved BSA that has passed a heating-cooling cycle. Copyright 2001 Academic Press.  相似文献   

18.
Bovine (BSA) and human (HSA) serum albumins are frequently used in biophysical and biochemical studies since they have a similar folding, a well known primary structure, and they have been associated with the binding of many different categories of small molecules. One important difference of BSA and HSA is the fact that bovine albumin has two tryptophan residues while human albumin has a unique tryptophan. In this work results are presented for the interaction of BSA and HSA with several ionic surfactants, namely, anionic sodium dodecyl sulfate (SDS), cationic cethyltrimethylammonium chloride (CTAC) and zwitterionic N-hexadecyl-N,N-dimethyl-3-ammonium-1-propanesulfonate (HPS), as monitored by fluorescence spectroscopy of intrinsic tryptophans and circular dichroism spectroscopy. On the interaction of all three surfactants with BSA, at low concentrations, a quenching of fluorescence takes place and Stern-Volmer analysis allowed to estimate their 'effective' association constants to the protein: for SDS, CTAC and HPS at pH 7.0 these constants are, respectively, (1.4+/-0.1) x 10(5) M(-1), (8.9+/-0.1) x 10(3) M(-1) and (1.4+/-0.1) x 10(4) M(-1). A blue shift of maximum emission is observed from 345 to 330 nm upon surfactant binding. Analysis of fluorescence emission spectra allowed to separate three species in solution which were associated to native protein, a surfactant protein complex and partially denatured protein. The binding at low surfactant concentrations follows a Hill plot model displaying positive cooperativity and a number of surfactant binding sites very close to the number of cationic or anionic residues present in the protein. Circular dichroism data corroborated the partial loss of secondary structure upon surfactant addition showing the high stability of serum albumin. The interaction of the surfactants with HSA showed an enhancement of fluorescence at low concentrations, opposite to the effect on BSA, consistent with the existence of a unique buried tryptophan residue in this protein with considerable static quenching in the native state. The effects of surfactants at low concentrations were very similar to those of myristic acid suggesting a non specific binding through hydrophobic interaction modulated by eletrostatic interactions. The changes in the vicinity of the tryptophan residues are discussed based on the recently published crystallographic structure of HSA myristate complex (S. Curry et al., Nat. Struct. Biol. 5 (1998) 827).  相似文献   

19.
Nanotextured diamond surfaces with geometrical properties close to protein dimensions were used for the realization of direct electron transfer of cytochrome c (cyt c) without any covalent bonding. The peroxidase activity of native and denatured cyt c was also investigated. Cyclic voltammograms of native cyt c show quasi-reversible electron transfer reactions, while no heme redox activity is detected for denatured cyt c. Unfolding (denaturation) of cyt c can be achieved in the presence of hydrogen peroxide. Partially or fully denatured cyt c showed higher peroxidase activity than native cyt c. This is because denatured cyt c loses its tertiary structure and hydrogen peroxide is easier to access the heme redox center. The apparent Michaelis–Menten constant Km for native and denatured cyt c has been determined to be 0.23 mM and 0.08 mM.  相似文献   

20.
The denatured state of a miniprotein BBA1 is studied under the native condition with the AMBER/Poisson-Boltzmann energy model and with the self-guided enhanced sampling technique. Forty independent trajectories are collected to sample the highly diversified denatured structures. Our simulation data show that the denatured BBA1 contains high percentage of native helix and native turn, but low percentage of native hairpin. Conditional population analysis indicates that the native helix formation and the native hairpin formation are not cooperative in the denatured state. Side-chain analysis shows that the native hydrophobic contacts are more preferred than the non-native hydrophobic contacts in the denatured BBA1. In contrast, the salt-bridge contacts are more or less nonspecific even if their populations are higher than those of hydrophobic contacts. Analysis of the trajectories shows that the native helix mostly initiates near the N terminus and propagates to the C terminus, and mostly forms from 3(10)-helix/turn to alpha helix. The same analysis shows that the native turn is important but not necessary in its formation in the denatured BBA1. In addition, the formations of the two strands in the native hairpin are rather asymmetric, demonstrating the likely influence of the protein environment. Energetic analysis shows that the native helix formation is largely driven by electrostatic interactions in denatured BBA1. Further, the native helix formation is associated with the breakup of non-native salt-bridge contacts and the accumulation of native salt-bridge contacts. However, the native hydrophobic contacts only show a small increase upon the native helix formation while the non-native hydrophobic contacts stay essentially the same, different from the evolution of hydrophobic contacts observed in an isolated helix folding.  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