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 140 毫秒
1.
Papain, a cysteine protease isolated from the latex of Carica papaya, is known to undergo irreversible thermal unfolding. In this study, we found that thermal unfolding of papain is accompanied by a simultaneous self-assembly process where this protein is observed to aggregate above 50°C. The extent of aggregation increased with increasing protein concentration from 3–40 μM. The aggregation was confirmed by enhanced turbidity, light scattering intensity, 1-anilino-8-naphthalene sulfonate (ANS) fluorescence intensity and by transmission electron microscopy. Furthermore, we noted that post-micellar concentration of sodium dodecyl sulfate (SDS) remarkably suppresses the thermal aggregation of papain. Far-UV circular dichroism studies revealed that SDS significantly enhances α-helical content of the protein and also tends to prevent its unfolding, and thus inhibits aggregation. Additionally, papain showed maximal activity at 65°C in neutral buffer. However, in the presence of 6 mM SDS (above its critical micellar concentration), the enzyme lost activity by about 10-fold. Thus, promoting the helical propensity of the protein does not appear to be a suitable strategy to overcome the aggregation related problems of industrially important proteins such as papain, which are not only required to be protected against aggregation but also need to remain functionally active in the presence of aggregation inhibitors.  相似文献   

2.
Pulsed-field-gradient nuclear magnetic resonance (PFG-NMR) is used to obtain the true hydrodynamic size of complexes of peptides with sodium dodecyl sulfate SDS micelles. The peptide used in this study is a 19-residue antimicrobial peptide, GAD-2. Two smaller dipeptides, alanine–glycine (Ala–Gly) and tyrosine–leucine (Tyr–Leu), are used for comparison. We use PFG-NMR to simultaneously measure diffusion coefficients of both peptide and surfactant. These two inputs, as a function of SDS concentration, are then fit to a simple two species model that neglects hydrodynamic interactions between complexes. From this we obtain the fraction of free SDS, and the hydrodynamic size of complexes in a GAD-2–SDS system as a function of SDS concentration. These results are compared to those for smaller dipeptides and for peptide-free solutions. At low SDS concentrations ([SDS] ≤ 25 mM), the results self-consistently point to a GAD-2–SDS complex of fixed hydrodynamic size R = (5.5 ± 0.3) nm. At intermediate SDS concentrations (25 mM < [SDS] < 60 mM), the apparent size of a GAD-2–SDS complex shows almost a factor of two increase without a significant change in surfactant-to-peptide ratio within a complex, most likely implying an increase in the number of peptides in a complex. For peptide-free solutions, the self-diffusion coefficients of SDS with and without buffer are significantly different at low SDS concentrations but merge above [SDS] = 60 mM. We find that in order to obtain unambiguous information about the hydrodynamic size of a peptide-surfactant complex from diffusion measurements, experiments must be carried out at or below [SDS] = 25 mM.  相似文献   

3.
Murraya koenigii miraculin‐like protein (MKMLP) gradually precipitates below pH 7.5. Here, we explore the basis for this aggregation by identifying the aggregation‐prone regions via comparative analysis of crystal structures acquired at several pH values. The prediction of aggregation‐prone regions showed the presence of four short peptides either in beta sheets or loops on surface of the protein. These peptides were distributed in two patches far apart on the surface. Comparison of crystal structures of MKMLP, determined at 2.2 Å resolution in pH 7.0 and 4.6 in the present study and determined at 2.9 Å in pH 8.0 in an earlier reported study, reveal subtle conformational differences resulting in gradual exposure of aggregation‐prone regions. As the pH is lowered, there are alterations in ionic interactions within the protein interactions of the chain with water molecules and exposure of hydrophobic residues. The analysis of symmetry‐related molecular interfaces involving one patch revealed shortening of nonpolar intermolecular contacts as the pH decreased. In particular, a decrease in the intermolecular distance between Trp103 of the aggregation‐prone peptide WFITTG (103–108) unique to MLPs was observed. These results demonstrated that aggregation occurs due to the cumulative effect of the changes in interactions in two aggregation‐prone defined regions. Proteins 2014; 82:830–840. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
Protein aggregation is commonly observed during protein refolding. To better understand this phenomenon, the intermolecular interactions experienced by a protein during unfolding and refolding are inferred from second virial coefficient (SVC) measurements. It is accepted that a negative SVC is indicative of protein-protein interactions that are attractive, whereas a positive SVC indicates net repulsive interactions. Lysozyme denatured and reduced in guanidinium hydrochloride exhibited a decreasing SVC as the denaturant was diluted, and the SVC approached zero at approximately 3 M GdnHCl. Further dilution of denaturant to renaturation conditions (1.25 M GdnHCl) led to a negative SVC, and significant protein aggregation was observed. The inclusion of 500 mM L-arginine in the renaturation buffer shifted the SVC to positive and suppressed aggregation, thereby increasing refolding yield. The formation of mixed disulfides in the denatured state prior to refolding also increased protein solubility and suppressed aggregation, even without the use of L-arginine. Again, the suppression of aggregation was shown to be caused by a shift from attractive to repulsive intermolecular interactions as reflected in a shift from a negative to a positive SVC value. To the best of our knowledge, this is the first time that SVC data have been reported for renaturation studies. We believe this technique will aid in our understanding of how certain conditions promote renaturation and increase protein solubility, thereby suppressing aggregation. SVC measurements provide a useful link, for protein folding and aggregation, between empirical observation and thermodynamics.  相似文献   

5.
Oxidation extent of myofibrillar protein (MP) from silver carp (Hypophthalmichthys molitrix) was affected by the content and type of lipid peroxidation (LPO) products. Oxidized linoleic acid (OLA) was selected as a main representative of lipid peroxidation to investigate the effects of oxidative modification of LPO products on MP structure. Structural changes of the oxidized myofibrillar protein were evaluated by the contents of carbonyl and total sulfhydryls, surface hydrophobicity, SDS-PAGE and Fourier transform infrared spectroscopy. Heating procedure was also applied for further evaluation of gelling properties. The results from SDS-PAGE indicated that aggregation and denaturation of myosin occurred in the oxidized system. The presence of OLA intensified oxidation-initiated loss of a-helix conformation as well as tertiary structure of MP. With the addition of OLA concentration less than 3 mM, a remarkably enhanced gelling capacity of MP was observed. While the excessive covalent bond (OLA > 5 mM) could lead to the breakage of protein-protein bonds, causing the collapse of the gel structure. The gelation procedure induced by OLA involved simultaneous protein oxidation and internal cross-linking.  相似文献   

6.
We investigate the interaction of Ca2+ (0–500 µM) and a membrane mimic (0.60 mM SDS) with both the R- and mixed spin states hemoglobin S (HbS) as a function of time. These interactions were carried out at pH 5.0. We aim at ascertaining if there is or are differences in the UV–Visible spectra of such interactions to account for the dynamics of calcium ion concentrations [Ca2+] in initiating structures which may ultimately suggest HbS polymerization and or resistance to Plasmodium attack. From our results, we conclude that (a) simultaneous interaction of 40 µM Ca2+ and 0.60 mM SDS with the R state protein would promote structural formations that can “lock up” the protein for nucleation on the membranes and or become cytotoxic to the parasite; (b) simultaneous R state HbS-SDS or R state HbS-Ca2+ would lead to enhanced hemin formation and less deoxyHb species. This condition is unlikely to precipitate polymerization in the HbS but the resulting hemin would poison the parasite; (c) the mixed spin state HbS-SDS and 40 µM Ca2+ interaction yields more toxic products to that of the interaction of the mixed spin HbS-SDS with 500 µM Ca2+ thus suggesting why the 40 µM Ca2+ is important in parasite Hb proteolysis; and (d) pronounced structural changes on interaction with SDS and Ca2+ are more in the R state to the mixed spin state.  相似文献   

7.
Prevention of undesirable protein aggregation is an extremely important strategy in protein science, medicine, and biotechnology. Arginine is one of the most widely used low molecular weight solution additives effective in suppressing aggregation, assisting refolding of aggregated proteins, and enhancing the solubility of aggregation-prone unfolded molecules in vitro. However, the mechanism of suppression of protein aggregation by arginine is not well understood. To address the mechanism, two model systems have been investigated: protection of alcohol dehydrogenase (ADH) and insulin from heat- and dithiothreitol-induced aggregation, respectively, in the presence of arginine. Using dynamic light scattering (DLS) technique, we have demonstrated the concentration-dependent suppression of light scattering intensity of both ADH and insulin aggregates upon addition of arginine to the incubation medium, a significant effect being revealed in the physiological concentration range of arginine (1-10 mM). DLS studies showed that arginine shifted the populations of nanoparticles with higher hydrodynamic radii to the lower ones, suggesting that the preventive effect of arginine on the protein aggregation process arises because it suppresses intermolecular interactions among aggregation-prone molecules. The results of turbidity measurements were also shown to be consistent with these findings.  相似文献   

8.
Sodium dodecyl sulfate (SDS), as an anionic surfactant, can induce protein conformational changes. Recent investigations demonstrated different effects of SDS on protein amyloid aggregation. In the present study, the effect of SDS on amyloid aggregation of bovine serum albumin (BSA) was evaluated. BSA transformed to β-sheet-rich amyloid aggregates upon incubation at pH 7.4 and 65°C, as demonstrated by thioflavin T fluorescence, circular dichroism, and transmission electron microscopy. SDS at submicellar concentrations inhibited BSA amyloid aggregation with IC50 of 47.5 μM. The inhibitory effects of structural analogs of SDS on amyloid aggregation of BSA were determined to explore the structure–activity relationship, with results suggesting that both anionic and alkyl moieties of SDS were critical, and that an alkyl moiety with chain length ≥10 carbon atoms was essential to amyloid inhibition. We attributed the inhibitory effect of SDS on BSA amyloid aggregation to interactions between the detergent molecule and the fatty acid binding sites on BSA. The bound SDS stabilized BSA, thereby inhibiting protein transformation to amyloid aggregates. This study reports for the first time that the inhibitory effect of SDS on albumin fibrillation is closely related to its alkyl structure. Moreover, the specific binding of SDS to albumin is the main driving force in amyloid inhibition. This study not only provides fresh insight into the role of SDS in amyloid aggregation of serum albumin, but also suggests rational design of novel antiamyloidogenic reagents based on specific-binding ligands.  相似文献   

9.
In the present paper, the interactions of the origin binding protein (OBP) of herpes simplex virus type 1 (HSV1) with synthetic four-way Holliday junctions (HJs) were studied using electrophoresis mobility shift assay and the FRET method and compared with the interactions of the protein with duplex and single-stranded DNAs. It has been found that OBP exhibits a strong preference for binding to four-way and three-way DNA junctions and possesses much lower affinities to duplex and single-stranded DNAs. The protein forms three types of complexes with HJs. It forms complexes I and II which are reminiscent of the tetramer and octamer complexes with four-way junction of HJ-specific protein RuvA of Escherichia coli. The binding approaches saturation level when two OBP dimers are bound per junction. In the presence of Mg2+ ions (≥2 mM) OBP also interacts with HJ in the stacked arm form (complex III). In the presence of 5 mM ATP and 10 mM Mg2+ ions OBP catalyzes processing of the HJ in which one of the annealed oligonucleotides has a 3′-terminal tail containing 20 unpaired thymine residues. The observed preference of OBP for binding to the four-way DNA junctions provides a basis for suggestion that OBP induces large DNA structural changes upon binding to Box I and Box II sites in OriS. These changes involve the bending and partial melting of the DNA at A+T-rich spacer and also include the formation of HJ containing Box I and Box II inverted repeats and flanking DNA sequences.  相似文献   

10.
In the present work, we study the effect of odorant binding on the thermal stability of honey bee (Apis mellifera L.) odorant-binding protein 14. Thermal denaturation of the protein in the absence and presence of different odorant molecules was monitored by Fourier transform infrared spectroscopy (FT-IR) and circular dichroism (CD). FT-IR spectra show characteristic bands for intermolecular aggregation through the formation of intermolecular β-sheets during the heating process. Transition temperatures in the FT-IR spectra were evaluated using moving-window 2D correlation maps and confirmed by CD measurements. The obtained results reveal an increase of the denaturation temperature of the protein when bound to an odorant molecule. We could also discriminate between high- and low-affinity odorants by determining transition temperatures, as demonstrated independently by the two applied methodologies. The increased thermal stability in the presence of ligands is attributed to a stabilizing effect of non-covalent interactions between odorant-binding protein 14 and the odorant molecule.  相似文献   

11.
Spinocerebellar ataxia type 3 (SCA3) is one of nine polyglutamine (polyQ) diseases all characterized by the presence of intraneuronal inclusions that contain aggregated protein. Aggregation of ataxin-3, the causative protein of SCA3, has been well characterized in vitro, with both pathogenic and non-pathogenic length ataxin-3 undergoing fibrillogenesis. However, only ataxin-3 containing an expanded polyQ tract leads to SCA3. Therefore other cellular factors, not present in previous in vitro studies, may modulate aggregation during disease. The interactions between fibrillar species and cell membranes have been characterized in a number of amyloid diseases, including Huntington’s Disease, and these interactions affect aggregation and toxicity. We have characterized the effects of the membrane mimetic sodium dodecyl sulfate (SDS) on ataxin-3 structure and aggregation, to show that both micellar and non-micellar SDS have differing effects on the two stages of ataxin-3 aggregation. We also demonstrate that fibrillar ataxin-3 binds phospholipids, in particular phosphorylated phosphotidylinositols. These results highlight the effect of intracellular factors on the ataxin-3 misfolding landscape and their implications in SCA3 and polyQ diseases in general are discussed.  相似文献   

12.
Protein aggregation generally results from association between hydrophobic regions of individual monomers. However, additional mechanisms arising from specific interactions, such as intermolecular disulfide bond formation, may also contribute to the process. The latter is proposed to be the initiating pathway for aggregation of immunoglobulin (IgG), which is essential for triggering its immune response. To test the veracity of this hypothesis, we have employed fluorescence correlation spectroscopy to measure the kinetics of aggregation of IgG in separate experiments either allowing or inhibiting disulfide formation. Fluorescence correlation spectroscopy measurements yielded a diffusion time (τD) of ~200 µsec for Rhodamine‐labeled IgG, corresponding to a hydrodynamic radius (RH) of 56 Å for the IgG monomer. The aggregation kinetics of the protein was followed by monitoring the time evolution of τD under conditions in which its cysteine residues were either free or blocked. In both cases, the progress curves confirmed that aggregation proceeded via the nucleation‐dependent polymerization pathway. However, for aggregation in the presence of free cysteines, the lag times were shorter, and the aggregate sizes bigger, than their respective counterparts for aggregation in the presence of blocked cysteines. This result clearly demonstrates that formation of intermolecular disulfide bonds represents a preferred pathway in the aggregation process of IgG. Fluorescence spectroscopy showed that aggregates formed in experiments where disulfide formation was prevented denatured at lower concentration of guanidine hydrochloride than those obtained in experiments where the disulfides were free to form, indicating that intermolecular disulfide bridging is a valid pathway for IgG aggregation. Proteins 2015; 83:169–177. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
Highly concentrated human recombinant interleukin-1 receptor antagonist (IL-1ra) aggregates at elevated temperature without perturbation in its secondary structure. The protein aggregation can be suppressed depending on the buffer ionic strength and the type of anion present in the sample solution. Phosphate is an approximately 4-fold weaker suppressant than either citrate or pyrophosphate on the basis of the measured protein aggregation rates. This is in agreement with the strength of protein-anion interactions at the IL-1ra single anion-binding site as judged by the estimated dissociation constant values of 2.9 mM, 3.8 mM, and 13.7 mM for pyrophosphate, citrate, and phosphate, respectively. The strength of binding also correlates with the anion size and with the number of ionized groups available per molecule at a given pH. Affinity probing of IL-1ra with methyl acetyl phosphate (MAP) in combination with proteolytic digestion and mass spectral analysis show that an anion-binding site location on the IL-1ra surface is contributed by lysine-93 and lysine-96 of the loop 84-98 as well as by lysine-6 of the unstructured N-terminal region 1-7. The replacement of lysine-93 with alanine by site-directed mutagenesis results in dramatically suppressed IL-1ra aggregation. Furthermore, when the unstructured N-terminal region of IL-1ra is removed by limited proteolysis, a 2-fold increase in the time course of the aggregation lag phase is observed for the truncated protein. An anion-controlled mechanism of IL-1ra aggregation is proposed by which the anion competition for the protein cationic site prevents formation of intermolecular cation-pi interactions and, thus, interferes with the protein asymmetric self-association pathway.  相似文献   

14.
The early intermediates in the protein aggregation pathway, the elusive soluble aggregates, play a pivotal role in growth and maturation of ordered aggregates such as amyloid fibrils. Blocking the growth of soluble oligomers is an effective strategy to inhibit aggregation. To decipher the molecular mechanisms and develop better strategies to arrest aggregation, it is imperative to understand how the size, molecular dynamics, activity and growth kinetics of soluble aggregates are affected when aggregation is inhibited. With this objective, in the present study we have investigated the influence of additives such as SDS, CTAB (cetyltrimethylammonium bromide) and DTT (dithiothreitol) on the slow aggregation of HEWL (hen eggwhite lysozyme) at pH 12.2. For this purpose, techniques such as steady-state and time-resolved fluorescence anisotropy of covalently labelled dansyl probe, gel-filtration chromatography, estimation of free thiol groups, thioflavin T and ANS (8-anilinonaphthalene-1-sulfonic acid) fluorescence, CD and atomic-force microscopy were employed to monitor the soluble oligomers over a period spanning 30 days. The results of the present study reveal that: (i) the spontaneous formation of soluble aggregates is irreversible and abolishes activity; (ii) the initial growth of aggregates (0-24 h) is promoted by a gradual increase in the exposure of hydrophobic surfaces; (iii) subsequently intermolecular disulfide bonds are critical for the assembly and stability of aggregates; (iv) the tight molecular packing inside large aggregates which contributed to slow (approximately 5 ns) and restricted segmental motion of dansyl probe was clearly loosened up in the presence of additives, enabling fast (1-2 ns) and free motion (unlike DTT, the size of lysozyme complexes with surfactants, was large, due to a conglomeration of proteins and surfactants); (v) the aggregates show reduced helical content compared with native lysozyme, except in the presence of SDS; and (vi) DTT was more potent than SDS/CTAB in arresting the growth of aggregates.  相似文献   

15.
A comparative study of amyloid properties of the aggregates of smooth muscle titin (SMT) from chicken gizzard was carried out. These aggregates were formed in two solutions: 0.15 M glycine-KOH, pH 7.2–7.4 (SMT(Gly)) and 0.2 M KCl, 10 mM imidazole, pH 7.0 (SMT(KCl)). Electron microscopy data showed that SMT aggregates has an amorphous structure in both cases. The results of atomic-force microscopy demonstrated slight differences in morphology in two types of aggregates. The SMT(Gly) aggregates were represented as branching chains, composed of spherical aggregates approximately 300–500 nm in diameter and up to 35 nm in height. The SMT(KCl) aggregates formed sponge-like structures with strands of 8–10 nm in height. Structural analysis of SMT aggregates by X-ray diffraction revealed the presence of cross-β-sheet structure in the samples under study. In the presence of SMT(Gly) aggregates, thioflavine T fluorescence intensity was higher (~3-fold times) compared with that in the presence of SMT(KCl) aggregates. Congo red-stained SMT(Gly) aggregates had yellow to apple-green birefringence under polarized light, which was not observed for SMT(KCl) aggregates. Dynamic light scattering data showed the similar rate of aggregation for both types of aggregates, though SMT(KCl) aggregates were able to partially disaggregate under increased ionic strength of the solution. The ability of SMT to aggregation followed by disaggregation may be functionally significant in the cell.  相似文献   

16.
As the most abundant glycoprotein component of pulmonary surfactant, SP-A (Mr = 30,000-36,000) plays a central role in the organization of phospholipid bilayers in the alveolar air space. SP-A, isolated from lung lavage, exists in oligomeric forms (N = 6, 12, 18, ...), mediated by collagen-like triple helices and intermolecular disulfide bonds. These protein-protein interactions, involving the amino-terminal domain of SP-A, are hypothesized to facilitate the alignment of surfactant lipid bilayers into unique tubular myelin structures. SP-A reorganization of surfactant lipid was assessed in vitro by quantitating the calcium-dependent light scattering properties of lipid vesicle suspensions induced by SP-A. Accelerated aggregation of unilamellar vesicles required SP-A and at least 3 mM free calcium. The initial rate of aggregation was proportional to the concentration of canine SP-A over lipid:protein molar ratios ranging from 200:1 to 5000:1. Digestion with bacterial collagenase or incubation with dithiothreitol (DTT) completely blocked lipid aggregation activity. Both treatments decreased the binding of SP-A to phospholipids. The conditions used in the DTT experiments (10 mM DTT, nondenaturing Tris buffer, 37 degrees C) resulted in the selective reduction and 14C-alkylation of the intermolecular disulfide bond involving residue 9Cys, whereas the four cysteines found in the noncollagenous domain of SP-A were inefficiently alkylated with [14C]-iodoacetate. HPLC analysis of tryptic SP-A peptides revealed that these four cysteine residues participate in intramolecular disulfide bond formation (138Cys-229Cys and 207Cys-221Cys). Our data demonstrate the importance of the quaternary structure (triple helix and intermolecular disulfide bond) of SP-A for the aggregation of unilamellar phospholipid vesicles.  相似文献   

17.
Polygalacturonase-3 was isolated and purified to homogeneity from palmyrah palm (Borassus flabellifer L.) fruit using Con A-Sepharose affinity column. The purified enzyme migrated as a single band on native and SDS–polyacrylamide gel electrophoresis. The molecular mass of the purified enzyme was estimated to be 66 kDa by size elution chromatography. Optimum polygalacturonase activity as a function of pH and temperature was determined using polygalacturonic acid as substrate. Optimum pH and temperature values ranged between the pH?4.0–5.0 and temperature 30–40 °C. At the optimum pH and temperature, the Km and Vmax values were determined by Lineweaver–Burk method. The value Km (0.33 mM) reveals that polygalacturonase has significant reactivity towards polygalacturonic acid. The enzyme showed varied responses towards divalent and monovalent metal ions. Ca2+ activated the polygalacturonase-3 enzyme protein. Both teepol and cetyltrimethylammonium bromide inhibited polygalacturonase-3 activity by 44 %, while 2-mercaptoethanol stimulated the enzyme marginally.  相似文献   

18.
We have studied interactions of cutinase (HiC) from Humicula insolens and sodium dodecyl sulphate (SDS) by parallel calorimetric and fluorescence investigations of systems in which the concentration of both components was changed systematically. Results from the two methods exhibit a number of synchronous characteristics, when plotted against the total SDS concentration, [SDS]tot. The molecular origin of several of these anomalies was assigned, and five intervals of [SDS]tot in which different modes of interactions dominated were identified. Going from low to high [SDS]tot, these modes were: binding of (a few) SDS to native HiC, formation of oligomeric protein aggregates, denaturation of HiC and adsorption of SDS on denatured protein. For [SDS]tot>3-6 mM (depending on the protein concentration), the adsorption saturated, and no further protein-detergent interaction could be detected. Two particularly conspicuous anomalies in the calorimetric data were ascribed to respectively denaturation and saturation. It was found that [SDS]tot at these points depended linearly on the (total) protein concentration, [HiC]. We suggest that this reflects the balance between bound and free SDS [SDS]tot=[SDS]aq+[HiC] Nb where [SDS]aq and Nb are, respectively, the aqueous ("free") concentration of SDS and the average number of SDS bound per protein. Interpretation of the results along these lines showed that at 22 degrees C and pH 7.0, HiC denatures with approximately 14 bound surfactant molecules at [SDS]aq=1.0 mM. Saturation is characterized by Nb approximately 39 and [SDS]aq=2.2 mM. The latter value is equal to CMC in the (protein free) buffer. These results are discussed with respect to the SDS-binding capacity of HiC and the origin and location of the saturation point.  相似文献   

19.
The thermal properties of cowpea protein isolates (CPI) were studied by differential scanning calorimetry under the influence of various conditions. An increase in the pH of protein extraction, from 8.0 to 10.0, during CPI preparation promoted a partial denaturation of cowpea proteins. Increases in enthalpy change of denaturation (ΔH) and temperature of denaturation (Td) were detected with increasing protein concentration from 7.5 to 10.5% (w/w). This behavior suggests that denaturation involves a first step of dissociation of protein aggregates. Calcium induced thermal stabilization in cowpea proteins, the increase in Td was ca. 0.3 °C/mM for protein dispersions of 7.5% (w/w) for 0 to 40 mM CaCl2. High hydrostatic pressure (HHP) induced denaturation in CPI in a pressure level dependent manner. The presence of calcium protected cowpea proteins towards HHP-induced denaturation when pressure level was 400 MPa, but not when it was 600 MPa. Thermal properties of cowpea protein isolates were very sensitive to processing conditions, these behaviors would have implications in processing of CPI-containing foodstuff.  相似文献   

20.
The effect of sodium dodecyl sulfate (SDS) on human, bovine, porcine, rabbit and sheep serum albumins were investigated at pH 3.5 by using various spectroscopic techniques like circular dichroism (CD), intrinsic fluorescence and dynamic light scattering (DLS). In the presence of 4.0 mM SDS the secondary structure of all the albumins were not affected as measured by CD but fluorescence spectra revealed 8.0 nm blue shift in emission maxima. We further checked the stability of albumins in the absence and presence of 4.0 mM SDS by urea and temperature at pH 3.5. In the absence of SDS, urea starts unfolding both secondary as well as tertiary structural elements of the all the albumins at ∼2.0 M urea but in the presence of 4.0 mM SDS, urea was unable to unfold even up to 9.0 M. The albumins were thermally less stable at pH 3.5 with decrease in Tm but in the presence of 4.0 mM SDS, the Tm was increased. From this study, it was concluded that SDS is showing a protective effect against urea as well as thermal denaturation of albumins. This behavior may be due to electrostatic as well as the hydrophobic interaction of SDS with albumins. Further, we have proposed the mechanism of action of urea. It was found that urea interacted with proteins directly when proteins are in charged form. Indirect interaction may be taking place when the environment is more hydrophobic.  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