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 interactions of dodecyltrimethylammonium bromides (DTABs) with hen egg lysozyme have been investigated at pH = 7.0 and 27 degrees C in phosphate buffer by isothermal titration calorimetry. DTAB interacts endothermically and activate lysozyme. The endothermicity of the lysozyme-DTAB interaction is in marked contrast to the exothermic interactions between sodium dodecyl sulphate (SDS) and lysozyme which have been attributed to specific binding between the anionic sulphate head groups and cationic amino acid residues. The enthalpies of interaction between the cationic surfactant (DTAB) and lysozyme are dominated by the endothermic unfolding of the native structure followed by an exothermic solvation of the lysozyme-DTAB complex by the addition of extra DTAB. A new direct calorimetric method to follow protein denaturation, and the effect of surfactants on the stability of proteins was introduced. The extended solvation model was used to reproduce the enthalpies of lysozyme-DTAB interaction over the whole range of DTAB concentrations. The solvation parameters recovered from the new equation, attributed to the structural change of lysozyme and its biological activity. At low concentrations of DTAB, the binding is mainly electrostatic, with some simultaneous interaction of the hydrophobic tail with nearby hydrophobic patches on the lysozyme. These initial interactions presumably cause some protein unfolding and expose additional hydrophobic sites. The DTAB-induced denaturation enthalpy of lysozyme is 86.46 +/- 0.02 kJ mol(-1).  相似文献   

2.
Yang J  Zhao J  Fang Y 《Carbohydrate research》2008,343(4):719-725
Interactions between the polyelectrolyte sodium alginate (NaAlg) and the anionic surfactant sodium dodecyl sulfate (SDS) have been investigated by microcalorimetric techniques. The polymer-surfactant interactions were observed between NaAlg and SDS at different pH values in dilute solution. The thermodynamic parameters for their interaction process are evaluated from the results of the observed dilution enthalpy curves. As the pH value of the solution decreases from 7 to 6, NaAlg polymers have an obvious effect on the cmc of SDS as a simple salt, which indicates no association between SDS and NaAlg owing to electrostatic repulsion. With the progressive decrease of pH value from 5 to 3, the hydrophobic segments in the alginate chains are increasing and the hydrophilic segments decreasing, and the aggregation between SDS and alginate due to hydrophobic interactions is observed.  相似文献   

3.
The extraction and purification of phycocyanin from Calothrix sp., cyanobacteria isolated from rice fields in Cuernavaca, Morelos, Mexico is described. Phycocyanin was extracted with 2 mg of lysozyme/g wet biomass, and purified by anion chromatography using Q-Sepharose fast-flow (Pharmacia®, 1.5 cm×10 cm) column and hydrophobic interaction chromatography with methyl macro-prep (Bio-Rad®, 1.5 cm×20 cm) column. The purified protein showed a pI of 5.2 and has two subunits with apparent molecular mass of 21–17 kDa each. The estimated molecular mass of native purified phycocyanin was 114 kDa, suggesting a stereochemistry of (β)3.  相似文献   

4.
The interaction of dodecyl trimethylammonium bromide (DTAB), a cationic surfactant, with calf thymus DNA has been studied by various methods, including potentiometric technique using DTAB-selective plastic membrane electrode at 27 and 37 degreesC, isothermal titration microcalorimetry and UV spectrophotometry at 27 degreesC using 0.05 M Tris buffer and 0.01 M NaCl at pH 7.4. The free energy is calculated from binding isotherms on the basis of Wyman binding potential theory and the enthalpy of binding according to van't Hoff relation. The enthalpy of unfolding has been determined by subtraction of the enthalpy of binding from the microcalorimetric enthalpy. The results show that, after the interaction of first DTAB molecule to DNA (base molarity) through the electrostatic interaction, the second DTAB molecule also binds to DNA through electrostatic interaction. At this stage, the predom-inant DNA conformational change occurs. Afterwards up to 20 DTAB molecules, below the critical micelle concentration of DTAB, bind through hydrophobic interactions.  相似文献   

5.
Several blue copper proteins are known to change the active site structure at alkaline pH (alkaline transition). Spectroscopic studies of Met16Phe, Met16Tyr, Met16Trp, and Met16Val pseudoazurin variants were performed to investigate the second sphere role through alkaline transition. The visible electronic absorption and resonance Raman spectra of Met16Phe, Met16Tyr, and Met16Trp variants showed the increasing of axial component at pH 11 like wild-type PAz. The visible electronic absorption and far-UV CD spectra of Met16Val demonstrated that the destabilization of the protein structure was triggered at pH > 11. Resonance Raman (RR) spectra of PAz showed that the intensity-weighted averaged Cu–S(Cys) stretching frequency was shifted to higher frequency region at pH 11. The higher frequency shift of Cu–S(Cys) bond is implied the stronger Cu–S(Cys) bond at alkaline transition pH 11. The visible electronic absorption and far-UV CD spectra of Met16X PAz revealed that the Met16Val variant is denatured at pH > 11, but Met16Phe, Met16Tyr, and Met16Trp mutant proteins are not denatured even at pH > 11. These observations suggest that Met16 is important to maintain the protein structure through the possible weak interaction between methionine –SCH3 part and coordinated histidine imidazole moiety. The introduction of π–π interaction in the second coordination sphere may be contributed to the enhancement of protein structure stability.  相似文献   

6.
The enthalpies of interaction of glucose oxidase at 25°C with a homologous series of n-alkyltrimethylammonium bromides (TABs) at pH 10 and a homologous series of n-alkylsulfates at pH 3.2 have been measured by microcalorimetry. For the n-dodecyl member of each series, DTAB and sodium n-dodecylsulfate (SDS), the binding of the surfactants to glucose oxidase as measured by equilibrium dialysis has been used in combination with the enthalpy data to obtain the Gibbs energy ( ), enthalpy ( ) and entropy ( ) of binding per surfactant molecule as a function of the number of surfactant molecules bound ( ). The thermodynamic parameters for the glucose oxidase interaction with DTAB at pH 10 and SDS at pH 3.2 are very similar and show that the interactions are entropically driven. The observed enthalpies of interaction of glucose oxidase with the homologous n-alkylsulfates have been analysed in terms of the interactions between the anionic surfactant head group and cationic sites on the protein, hydrophobic binding and the thermal contributions arising from protein unfolding. At surfactant concentrations of 0.5 c.m.c., the enthalpy of unfolding of glucose oxidase is estimated to be 3610 ± 560 kJ mol−1.  相似文献   

7.
Aspergillus niger produces multiple forms of polygalacturonases with molecular masses ranging from 30 to 60 kDa. The high molecular weight polygalacturonase (61 ± 2 kDa) from A. niger possesses a pH optimum of 4.3 and a pI of 3.9. The enzyme exhibited high sensitivity, both in terms of activity and structure, in the pH range of 4.3–7.0. The enzyme was irreversibly inactivated at pH 7.0. The enzyme is predominantly rich in parallel β structure. There is unfolding of the enzyme molecule between 4.3 and 7.0 resulting in irreversible loss of secondary and tertiary structure with the exposure of hydrophobic surfaces. ANS binding measurements, intrinsic fluorescence and acrylamide quenching measurements have confirmed the unfolding and exposure of hydrophobic surfaces. The midpoint of pH transition for both activity and secondary structure is 6.2 ± 0.1. The pH-induced changes of polygalacturonase confirm the role of histidine residues in structure and activity of the enzyme. The irreversible nature of inactivation is due to the unfolding induced exposure of hydrophobic surfaces leading to association/aggregation of the molecule. Size exclusion chromatography measurements have established the association of enzyme at higher pH. Urea induced unfolding measurements at pH 4.3 and 7.0 have confirmed the loss in stability as we approach neutral pH.  相似文献   

8.
Metalloprotein analysis by capillary isoelectric focusing   总被引:1,自引:0,他引:1  
Capillary isoelectric focusing (cIEF) was used to analyze three metalloproteins: conalbumin, transferrin and metallothionein (MT). Two different ampholyte mixtures were employed that generated linear pH gradients of 3–10 and 5–8. Several different proteins and one peptide known isoelectric points (pIs) were used to establish linear relationships between peak migration time and pI. These standards were also used as internal markers to estimate peak pI values of the metalloproteins subjected to cIEF. Conalbumin (iron-free) subjected to cIEF with a pH gradient of 3–10 yielded a single major component (pI 7.17). When the protein was saturated with iron (2 Fe3+/mol protein), a shift to lower pI was observed with a major peak (pI 6.24) and a lesser peak (pI 6.09). Mixing iron-free with iron-saturated conalbumin or adding iron to iron-free conalbumin prior to cIEF produced an additional peak (pI 6.68) that was presumed to be conalbumin containing a single iron atom (monoferric form). Human transferrin subjected to cIEF with a pH range of 3–10 gave a similar separation pattern to conalbumin with four major peaks at pI values of 6.25 (apotransferrin), 5.96 (monoferric form), 5.48 and 5.34 (differic forms). Additional resolution of the molecular forms of both conalbumin and transferrin was achieved using a narrower pH gradient (5–8). Rabbit liver MT subjected to cIEF with a pH gradient of 3–10 gave a complex separation pattern with two prominent peaks (pI values of 3.73 and 3.56) that were presumed to be the fully metal-saturated MT-1 and MT-2 isoforms. When individual MT isoforms (MT-1 and MT-2) were separately subjected to cIEF with a pH gradient of 3–10, heterogeneous peaks with higher pI values (4.12–4.74) were observed. In contrast, horse kidney MT gave a single predominant peak with a pI of 4.09. MT samples could be separated using pH gradient of 5–8 despite the fact that their apparent pI values were below the limits of the pH gradient established. In general, the heterogeneity observed for conalbumin, transferrin and MT proteins subjected to cIEF reflects the presence or absence of bound metal. Thus, cIEF represents a potentially useful analytical method which can provide information concerning the metal-binding characteristics of these and perhaps other metalloproteins.  相似文献   

9.
Annelid erythrocruorins are respiratory proteins with high cooperativity and low autoxidation rates. The giant extracellular hemoglobin of the earthworm, Glossoscolex paulistus (HbGp), has a molecular mass of 3.6 MDa. In this work, isothermal titration calorimetry (ITC), together with DLS and fluorescence emission have been used to investigate the interaction of SDS with the HbGp in the oxy‐form, at pH 7.0. Our ITC and DLS results show that addition of SDS induces oxy‐HbGp oligomeric dissociation, while a small amount of protein aggregation is observed only by DLS. Moreover, the oligomeric dissociation process is favored at lower protein concentrations. The temperature effect does not influence significantly the interaction of SDS with the hemoglobin, due to the similarities presented by the critical aggregation concentration (cac) and critical micelle concentration (cmc′) for the mixtures. The increase of oxy‐HbGp concentration leads to a slight variation of the cac values for the SDS‐oxy‐HbGp mixture, attributed mainly to the noncooperative electrostatic binding of surfactant to protein. However, the cmc′ values increase considerably, associated to a more cooperative hydrophobic binding. Complementary pyrene fluorescence emission studies show formation of pre‐micellar structures of the mixture already at lower SDS concentrations. This study opens the possibility of the evaluation of the surfactant effect on the hemoglobin stability by ITC, which is made for the first time with this extracellular hemoglobin. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1065–1076, 2014.  相似文献   

10.
The interaction of alpha-amylase with n-alkylammonium bromides above and below their critical micellar concentrations (cmc) has been studied in buffer at pH 7 and 10 by UV spectrophotometry, photon correlation spectroscopy and Doppler microelectrophoresis. This interaction produces a complex that is dependent on pH of the medium. This complex appears at surfactant concentrations below the cmc, which means that individual surfactant molecules can bind tightly to native alpha-amylase. The complex maintains its aggregation state when the concentration of surfactants with a hydrocarbon chain of 16 carbons increases, but not for surfactants of 12 and 14 carbons. Measurements of zeta-potential indicate the influence of electrostatic and hydrophobic forces. When the size of the aggregate is maximal, proteins are at their point of zero charge. In such conditions, Van der Waals forces and contacts between the alkyl chain and the hydrophobic core of the protein favour the formation of a larger aggregate.  相似文献   

11.
The amyloid β‐peptide fragment comprising residues 25–35 (Aβ25‐35) is known to be the most toxic fragment of the full length Aβ peptide which undergoes fibrillation very rapidly. In the present work, we have investigated the effects of the micellar environment (cationic, anionic, and nonionic) on preformed Aβ25‐35 fibrils. The amyloid fibrils have been prepared and characterized by several biophysical and microscopic techniques. Effects of cationic dodecyl trimethyl ammonium bromide (DTAB), cetyl trimethylammonium bromide (CTAB), anionic sodium dodecyl sulfate (SDS), and nonionic polyoxyethyleneoctyl phenyl ether (Triton X‐100 or TX) on fibrils have been studied by Thioflavin T fluorescence, UV–vis spectroscopy based turbidity assay and microscopic analyses. Interestingly, DTAB and SDS micelles were observed to disintegrate prepared fibrils to some extent irrespective of their charges. CTAB micelles were found to break down the fibrillar assembly to a greater extent. On the other hand, the nonionic surfactant TX was found to trigger the fibrillation process. The presence of a longer hydrophobic tail in case of CTAB is assumed to be a reason for its higher fibril disaggregating efficacy, the premise of their formation being largely attributed to hydrophobic interactions. Proteins 2016; 84:1213–1223. © 2016 Wiley Periodicals, Inc.  相似文献   

12.
Antioxidants solubilized in micellar solutions can change micellar properties like the size and shape of micelles, critical micellar concentration (cmc) and viscosity. Interactions arising between antioxidants and the surfactant determine the locations of antioxidants and vice versa. The location and interaction are dependent on the type of both the antioxidant and surfactant. Influences of various antioxidants on the physical and structural properties were tested in micellar systems of cationic CTAB, non-ionic Brij 58 and anionic SDS. The antioxidants used to investigate the effects of gradually increasing lipophilicity were gallic acid (GA) and the gallate esters from methyl to octyl gallate (MG-OG). Hydroxy cinnamic acids (HCAs) like -coumaric acid (pC), caffeic acid (CA), ferulic acid (FA) and sinapic acid (SA) were employed to observe effects of functional groups like hydroxyl and methoxy groups. Micellar size and shape determined by small angle neutron scattering (SANS), viscosity and cmc were chosen to characterize the antioxidant influence. In Brij 58 systems propyl gallate (PG) did not affect the cmc or aggregation number but decreased micellar size slightly due to an intercalation of PG into the region of the polyoxyethylene chain and the first adjacent alkyl chain methylene groups. In SDS systems the micellar size and cmc decreased in the presence of PG. This was attributed to PG residing in the Stern layer. However, in CTAB systems micelles swelled at low PG concentration and in the presence of GA, while higher PG concentrations and more lipophilic antioxidants led to a sphere-to-rod transition with a simultaneous increase in viscosity and decrease in cmc. This revealed the intercalation of antioxidants in the palisade layer of CTAB micelles entering into strong interactions of electrostatic and hydrophobic origin. It could be demonstrated that the interactions became stronger the more lipophil an antioxidant is and the more hydroxyl groups are attached to the aromatic ring. Differences in the location and interaction of antioxidant and micelles are proposed as being responsible for the effectiveness of antioxidants.  相似文献   

13.
The microviscosity in mixed micelles formed in dilute aqueous solutions of sodium dodecyl sulphate (SDS) and a set of non-ionic cellulose ethers of different hydrophobicity has been determined by means of steady-state fluorescence probe techniques. Two hydrophobic probes have been applied in this investigation: 1,3-di(1-pyrenyl)propane (P3P) and perylene. Reference measurements of microviscosity have also been performed on SDS solutions including the uncharged polymers poly(ethyleneoxide) (PEO) or poly(vinylpyrrolidone) (PVP). All compositions investigated showed qualitatively the same general behaviour with an abrupt increase in microviscosity at the critical surfactant concentration where the polymer-surfactant interaction starts (c1) followed by a maximum and an asymptotically declining region as the surfactant concentration was increased further. Comparison with a recent investigation of a specific ethyl(hydroxyethyl)cellulose (EHEC fraction CST-103)/ SDS/water system (Evertsson & Nilsson (1997) Macromolecules, 30, 2377) revealed that the maximum in microviscosity generally corresponds to a low degree of SDS adsorption (≈ 0.5 mmol of SDS per gram of polymer) and consequently to a high polymer content of the mixed micelles formed in the type of systems studied herein. The hydrophobicity of the cellulose derivatives was found to correlate to the amplitude of the overall microviscosity pattern for the mixed micelles, i.e. an increased polymer hydrophobicity gave an increased rigidity of the polymersurfactant aggregates. An approximately exponential relation was demonstrated between the maxima in microviscosity of the different mixed micelles and the surface activities of the corresponding cellulose derivatives. All polymer/surfactant combinations investigated gave aggregates with a higher rigidity than ordinary SDS micelles. The microviscosity of the mixed micelles of the cellulose derivatives and SDS formed close to c1 increased as the temperature rose from 20 to 50 °C. This effect was attributed to an increased hydrophobicity of the cellulose ethers upon temperature elevation, hence giving rise to further close-packing of the aggregate structures.  相似文献   

14.
Useful materials can be made from cycloamylose (CA) and the functional properties of CA could be improved by complexation with surfactants. Isothermal titration calorimetry (ITC) was used to investigate interactions between CA and surfactants in buffered solutions. Three surfactants with C12 non-polar tail groups and charged [anionic: sodium dodecyl sulfate (SDS); cationic: dodecyl trimethylammonium bromide (DTAB)] or non-charged headgroups [non-ionic: polyoxyethylene 23 lauryl ether (Brij35)] were used in this study. The effects of temperature, pH, and salt concentration were also studied. All three surfactants bound to CA; however, Brij35 binding to CA was negligible. Enthalpy changes associated with binding of surfactants to CA were exothermic except for interactions measured at 50 °C. There was no effect of pH on surfactant demicellization or CA binding. Salt concentration affected surfactant demicellization, but the amount of SDS bound to CA at saturation was unaffected by salt. When the titration curves obtained for CA with SDS and DTAB were fitted, it could be analyzed using a model based on a single set of identical sites.  相似文献   

15.
The solution properties of hydroxyethyl cellulose (HEC) and hydrophobically modified hydroxyethyl cellulose (HM-HEC) have been investigated by means of viscometric and spectroscopic techniques involving free radical and fluorescent probes. The greater viscosity of HM-HEC solutions above a critical polymer concentration (Cp) of approximately 0·2% has been interpreted in terms of the formation of a three-dimensional network structure in which the polymer chains are effectively crosslinked by the intermolecular association of neighbouring hydrophobic side chains. Cp is considerably less than the predicted polymer coil overlap concentration (C*) of approximately 1%.

The interaction of the polymers with an anionic surfactant, sodium dodecyl sulphate (SDS) has also been investigated. A mechanism involving the interaction of free surfactant with the regions of intermolecular hydrophobic association is suggested to account for the considerable differences in the rheological behaviour of the polymers in the presence of SDS.  相似文献   


16.
Peroxidase oxidation of o-dianisidine, 3,3′,5,5′-tetramethylbenzidine, and o-phenylenediamine in the presence of sodium dodecyl sulfate (SDS), an anionic surfactant, was spectrophotometrically studied. It was found that 0.1–100 mM SDS concentrations stabilize intermediates formed in the peroxidase oxidation of these substrates. The cause of the stabilization is an electrostatic interaction between positively charged intermediates and negatively charged surfactant.  相似文献   

17.
The study includes partitioning of proteins in aqueous two-phase systems consisting of the polymer dextran and the non-ionic surfactant C12E5 (pentaethylene glycol mono-n-dodecyl ether). In this system a micelle-enriched phase is in equilibrium with a polymer-enriched phase. Charges can be introduced into the micelles by the addition of charged surfactants. The charge of the mixed micelles is easily varied in sign and magnitude independently of pH, by the addition of different amounts of negatively charged surfactant, sodium dodecyl sulphate (SDS), or positively charged surfactant dodecyl trimethyl ammonium chloride (DoTAC). A series of water-soluble model proteins (BSA, β-lactoglobulin, myoglobin, cytochrome c and lysozyme), with different net charges at pH 7.1, have been partitioned in non-charged systems and in systems with charged mixed micelles or charged polymer (dextran sulphate). It is shown that partition coefficients for charged proteins in dextran-C12E5 systems can be strongly affected by addition of charged surfactants (SDS, DoTAC) or polymer (dextran sulphate) and that the effects are directly correlated to protein net charge.  相似文献   

18.
Peroxidase oxidation of o-dianisidine, 3,3',5,5'-tetramethylbenzidine, and o-phenylenediamine in the presence of sodium dodecyl sulfate (SDS), an anionic surfactant, was spectrophotometrically studied. It was found that 0.1-100 mM SDS concentrations stabilize intermediates formed in the peroxidase oxidation of these substrates. The cause of the stabilization is an electrostatic interaction between positively charged intermediates and negatively charged surfactant.  相似文献   

19.
The interaction of bradykinin (BK) with lipids has been followed by steady-state fluorescence measurements. Addition of either cerebroside sulfate (CS) or phosphatidylinositol (PI), solubilized with the nonionic surfactant C12E8, to BK or its analogue [Gly6]-BK enhances the relative fluorescence intensity of peptide emission at 288 nm. Fluorometric titration of the peptide with lipid has been used to quantitate the interactions in terms of stoichiometry and equilibrium constant. Job's method of continuous variation for the BK-CS interaction gave a stoichiometry of 1:2 for the complex. The value of the equilibrium constant, K, for the interaction of either BK or [Gly6]-BK with CS is 1.5.10(4) M-1. The BK-PI interaction is weaker; K = 5.0.10(3) M-1. Although electrostatic forces no doubt play a major role in these interactions, measurements on the model peptide Gly-Phe-Gly indicate that the phenylalanine residues of BK are disposed in the hydrophobic environment provided by the lipid-C12E8 mixed micelle. 13C-NMR measurements on [99% 13C alpha-Gly6]-BK show that there is no change in its cis/trans ratio upon interaction with CS. The increase in the relative fluorescence intensity of BK accompanying its cooperative interaction with sodium dodecyl sulfate (SDS) implicates the role of hydrophobic forces in this interaction as well. These results bear on the interpretation of the changes in circular dichroism (CD) of BK caused by SDS.  相似文献   

20.
Molecules, whose pK(a) values can be easily fine-tuned by their microenvironment, are expected to be profoundly affected by the heterogeneous environments of membranes. Membrane parameters can have a strong effect in choosing a particular structural form of a molecule for incorporation/interaction. A case study has been presented for piroxicam, a non-steroidal anti-inflammatory drug of oxicam group, whose targets are cyclooxygenases, which are membrane active proteins. The structural dynamism of piroxicam is reflected in the ease with which it can switchover or convert from one prototropic form to the other guided by its environment. In this work we have studied the effect of varying hydrophobic chain length and surface charges in fine-tuning the interaction of piroxicam with micelles. Interaction of piroxicam with three types of micelles with identical negatively charged head groups and varying tail lengths viz., sodium dodecyl sulfate (S12S), sodium decyl sulfate (S10S) and sodium octyl sulfate (S8S) shows that there is a shift in the apparent pK(a) in the direction that favors the switchover or conversion from the anionic form to the global neutral form. The binding constants of piroxicam with three micelles show a linear dependence on chain length. Interaction was also studied with micelles having oppositely charged head groups and different chain lengths viz., dodecyl N,N,N-trimethyl ammonium bromide (DTAB) and cetyl N,N,N-trimethyl ammonium bromide (CTAB). For micelles having identical chain lengths but oppositely charged head groups viz., S12S and DTAB, pK(a) shifts in two opposite directions compared to that in the absence of any surfactant. This is expected when electrostatic force is the only driving force. This case study demonstrates the effect of hydrophobic chain length and surface charges in fine-tuning the equilibrium between different structural forms of piroxicam. Our results also imply that for structurally dynamic drugs like piroxicam the nature of the biomembranes, characterized by different membrane parameters, should play a crucial role in choosing a particular structural form of the drug that will be finally presented to their targets.  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