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 31 毫秒
1.
The three-dimensional solution structure of the 40 residue amyloid beta-peptide, Abeta(1-40), has been determined using NMR spectroscopy at pH 5.1, in aqueous sodium dodecyl sulfate (SDS) micelles. In this environment, which simulates to some extent a water-membrane medium, the peptide is unstructured between residues 1 and 14 which are mainly polar and likely solvated by water. However, the rest of the protein adopts an alpha-helical conformation between residues 15 and 36 with a kink or hinge at 25-27. This largely hydrophobic region is likely solvated by SDS. Based on the derived structures, evidence is provided in support of a possible new location for the transmembrane domain of Abeta within the amyloid precursor protein (APP). Studies between pH 4.2 and 7.9 reveal a pH-dependent helix-coil conformational switch. At the lower pH values, where the carboxylate residues are protonated, the helix is uncharged, intact, and lipid-soluble. As the pH increases above 6. 0, part of the helical region (15-24) becomes less structured, particularly near residues E22 and D23 where deprotonation appears to facilitate unwinding of the helix. This pH-dependent unfolding to a random coil conformation precedes any tendency of this peptide to aggregate to a beta-sheet as the pH increases. The structural biology described herein for Abeta(1-40) suggests that (i) the C-terminal two-thirds of the peptide is an alpha-helix in membrane-like environments, (ii) deprotonation of two acidic amino acids in the helix promotes a helix-coil conformational transition that precedes aggregation, (iii) a mobile hinge exists in the helical region of Abeta(1-40) and this may be relevant to its membrane-inserting properties and conformational rearrangements, and (iv) the location of the transmembrane domain of amyloid precursor proteins may be different from that accepted in the literature. These results may provide new insight to the structural properties of amyloid beta-peptides of relevance to Alzheimer's disease.  相似文献   

2.
针铁矿吸附水中硫酸根离子的试验研究   总被引:7,自引:0,他引:7  
孙进  王志国  李龙海 《黄金》2005,26(3):43-45
利用合成的针铁矿对水中硫酸根离子进行吸附试验。试验表明,吸附量主要受pH值和硫酸根离子浓度影响。通过配位体交换理论,很好地解释了吸附量随pH下降而增加,随硫酸根离子浓度的下降而减少。  相似文献   

3.
Aggregation and deposition of the 40-42-residue amyloid beta-protein (Abeta) are early and necessary neuropathological events in Alzheimer's disease. An understanding of the molecular interactions that trigger these events is important for therapeutic strategies aimed at blocking Abeta plaque formation at the earliest stages. Heparan sulfate proteoglycans may play a fundamental role since they are invariably associated with Abeta and other amyloid deposits at all stages. However, the nature of the Abeta-heparan sulfate proteoglycan binding has been difficult to elucidate because of the strong tendency of Abeta to self-aggregate. Affinity co-electrophoresis can measure the binding of proteoglycans or glycosaminoglycans to proteins without altering the physical state of the protein during the assay. We used affinity co-electrophoresis to study the interaction between Abeta and the glycosaminoglycan heparin and found that the aggregation state of Abeta governs its heparin-binding properties: heparin binds to fibrillar but not nonfibrillar Abeta. The amyloid binding dye, Congo red, inhibited the interaction in a specific and dose-dependent manner. The "Dutch" mutant AbetaE22Q peptide formed fibrils more readily than wild type Abeta and it also attained a heparin-binding state more readily, but, once formed, mutant and wild type fibrils bound heparin with similar affinities. The heparin-binding ability of aggregated AbetaE22Q was reversible with incubation in a solvent that promotes alpha-helical conformation, further suggesting that conformation of the peptide is important. Studies with another human amyloidogenic protein, amylin, suggested that its heparin-binding properties were also dependent on aggregation state. These results demonstrate the dependence of the Abeta-heparin interaction on the conformation and aggregation state of Abeta rather than primary sequence alone, and suggest methods of interfering with this association.  相似文献   

4.
Conformational studies were performed on the synthetic tricosapeptide N-acetyl-SKKALKKLQKEQEKQRKKEERAL-amide, representing the highly basic segment (residues 30-52) of the N-terminal extension of yeast cytoplasmic aspartyl-tRNA synthetase. Circular dichroism experiments show that, in aqueous solution at neutral pH, the peptide adopts a random conformation. The effects of pH, temperature, addition of trifluoroethanol (TFE), and titration with polyanions on the conformation of the peptide were studied. In TFE or in the presence of an equimolar concentration of (phosphate)18, the peptide adopts a 100% alpha-helical conformation. A partially alpha-helical conformation is induced by (phosphate)4 or d(pT)8 (respectively 40% and 35% helical content). Raising the pH in aqueous solution promotes 75% alpha-helicity, with a transition pK of 9.9 reflecting deprotonation of lysine residues. On the basis of these results, nuclear magnetic resonance studies were carried out in TFE as well as in aqueous solution in the presence of (phosphate)18, to determine the structure of the molecule. Complete 1H resonance assignments were obtained by conventional two-dimensional NMR techniques. A total of 138 interproton constraints derived from NOESY experiments were used to calculate the three-dimensional structure by a two-stage distance geometry/simulated annealing procedure. The two deduced structures were highly similar and show that nine cationic residues are segregated on one face of a helical structure, providing an ideal polycationic interface for binding to polyanionic surfaces.  相似文献   

5.
Human erythrocyte band 3 was purified essentially free of peripheral proteins, in particular band 4.2, using affinity chromatography. Band 3 protein was then reconstituted into liposomes of lipid type and ratio approximating that of erythrocyte membranes. Stilbenedisulfonate inhibition of band 3 mediated efflux of radiolabeled sulfate from preloaded liposomes was used to test the functionality and correct orientation of the protein. When sulfate efflux, mediated by purified band 3, was compared with partially purified band 3, which contained detectable amounts of bands 4.1 and 4.2, a clear difference in efflux was measured. Sulfate efflux was approximately 30% faster from liposomes containing purified band 3 compared with those containing partially purified protein. In order to investigate further any specific effect of band 4.2 protein on band 3 mediated anion transport, band 4.2 was purified. Increasing amounts of band 4.2 were complexed with purified band 3 and then reconstituted into liposomes. Increasing amounts of band 4.2 complexed with band 3 caused a decrease in band 3 mediated anion transport. The effect of band 4.2 on band 3 mediated anion transport appears to be specific since increasing concentrations of band 4.2 added exogenously to band 3 in reconstituted vesicles (rather than complexed with band 3 before reconstitution) produced no significant changes in sulfate efflux. Further, when increasing amounts of band 4.2 were added to the functionally active transmembrane domain of band 3 and then reconstituted into vesicles, there was also no significant change in sulfate efflux.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Sulfate conjugates are a heterogeneous class of polar, anionic metabolites that result from the conjugation of endogenous and exogenous compounds. Sulfate conjugates exhibit a high degree of binding to albumin, the extent of which usually exceeds those of their parent compounds. Preponderant direct and indirect evidence suggests that sulfation activity is slightly higher in the periportal than in the perivenous (centrilobular) region of the liver, but recent immunohistochemical studies imply that specific isoforms of the sulfotransferases may also be preferentially localized in the perivenous region. Entry of sulfate conjugates into the liver cell is poor unless discrete carriers are present. Although known transport carriers exist for the sulfated bile acids, the specificity of the carriers for drug sulfate conjugates is presently unknown. The removal of sulfates is usually by way of biliary excretion while, on occasion, sulfates can be desulfated and participate in futile cycling with their parent compounds. The binding, transport, and hepatic elimination of various drug sulfate conjugates are examined. Non-recirculating studies carried out in the perfused rat liver with the multiple indicator dilution technique under varying input sulfate conjugate concentrations have provided essential information on the effects of vascular (red blood cells and plasma protein) binding on transport and removal of the conjugates. These studies clearly demonstrate the need to study protein binding, transmembrane transfer characteristics across the liver basolateral (sinusoidal) and canalicular membranes, and enzyme zonation in a distributed-in-space fashion in order to properly define the handling of sulfate conjugates in the liver.  相似文献   

7.
The effect of genital prolapse on voiding   总被引:1,自引:0,他引:1  
Jovin and co-workers have demonstrated that DNA strands containing guanine-adenine repeats generate a parallel-stranded homoduplex. Here we propose that the homoduplex is a dimer of the ordered single strand discovered by Fresco and co-workers at acid pH. The Fresco single strand is shown here to be stabilized in aqueous ethanol where adenine is not protonated. Furthermore, we demonstrate that the strands dimerize at higher salt concentrations without significantly changing their conformation, so that the dimerization is non-cooperative. Hence, the Jovin homoduplex can form through a non-cooperative dimerization of two cooperatively melting single strands. The available data indicate that the guanines stabilize the Fresco single strand whereas the adenines cause dimerization owing to their known intercalation or clustering tendency. The guanine-adenine repeat dimer seems to be a DNA analog of the leucine zipper causing dimerization of proteins.  相似文献   

8.
The dissolution of chalcopyrite in ferric sulfate and ferric chloride media   总被引:1,自引:0,他引:1  
The literature on the ferric ion leaching of chalcopyrite has been surveyed to identify those leaching parameters which are well established and to outline areas requiring additional study. New experimental work was undertaken to resolve points still in dispute. It seems well established that chalcopyrite dissolution in either ferric chloride or ferric sulfate media is independent of stirring speeds above those necessary to suspend the particles and of acid concentrations above those required to keep iron in solution. The rates are faster in the chloride system and the activation energy in that medium is about 42 kJ/mol; the activation energy is about 75 kJ/mol in ferric sulfate solutions. It has been confirmed that the rate is directly proportional to the surface area of the chalcopyrite in both chloride and sulfate media. Sulfate concentrations, especially FeSO4 concentrations, decrease the leaching rate substantially; furthermore, CuSO4 does not promote leaching in the sulfate system. Chloride additions to sulfate solutions accelerate slightly the dissolution rates at elevated temperatures. It has been confirmed that leaching in the ferric sulfate system is nearly independent of the concentration of Fe3+, ka[Fe3+]0.12. In ferric chloride solutions, the ferric concentration dependence is greater and appears to be independent of temperature over the interval 45 to 100 °C.  相似文献   

9.
The mmr gene, cloned from Mycobacterium tuberculosis, was shown to confer to Mycobacterium smegmatis resistance to tetraphenylphosphonium (TPP), erythromycin, ethidium bromide, acriflavine, safranin O, and pyronin Y. The gene appears to code for a protein containing four transmembrane domains. Studies of [3H]TPP intracellular accumulation strongly suggest that the resistance mediated by the Mmr protein involves active extrusion of TPP.  相似文献   

10.
We have studied the effects of extrinsic environmental conditions on the conformation of surfactin, a heptapeptide biosurfactant from Bacillus subtilis, in aqueous solutions. It has been made clear that temperature, pH, Ca2+ ions and the synthetic nonionic surfactant hepta-ethylene glycol (C12E7) affect the conformation of surfactin in aqueous solutions. The beta-sheet formation reached a maximum at 40 degrees C both in presence and absence of (C12E7) and the nonionic surfactant enhances the beta-sheet formation even at 25 degrees C. Ca2 + induced the formation of alpha-helices and caused this transition at 0.3 mM with surfactin monomers or at 0.5 mM with surfactin micelles, but above these transition concentrations of Ca2+ beta-sheets were observed. In micellar solution the beta-sheet structure was stabilized at pH values below 7 or upon addition of Ca2+ in concentrations above 0.5 mM. Our results indicated that the bioactive conformation of surfactin is most likely the beta-sheets when the molecules are assembled in micelles. The beta-sheet structure in micelles could be retained by tuning the micelles. Surfactin micelles could be tuned in the bioactive conformation by manipulating pH, temperature, Ca2+ or (C12E7) concentrations in surfactin solutions. Our results strongly indicated that Ca2+ and other molecules (such as C12E7) may function as directing templates in the assembly and conformation of surfactin in micelles. Thus, we suggest environmental manipulation and template-aided micellation (TAM) as a new approach for preparing predesigned micelles, microemulsions or micro-spheres for specific application purposes.  相似文献   

11.
The folding of Ser-Tyr-Pro-Phe-Asp-Val (SYPFDV), and sequence variants of this peptide (SYPYD and SYPFD) are studied computationally in an explicit water environment. An atomically detailed model of the peptide is embedded in a sphere of TIP3P water molecules and its optimal structure is computed by simulated annealing. At distances from the peptide that are beyond a few solvation shells, a continuum solvent model is employed. The simulations are performed using a mean field approach that enhances the efficiency of sampling peptide conformations. The computations predict a small number of conformations as plausible folded structures. All have a type VI turn conformation for the peptide backbone, similar to that found using NMR. However, some of the structures differ from the experimentally proposed ones in the packing of the proline ring with the aromatic residues. The second most populated structure has, in addition to a correctly folded backbone, the same hydrophobic packing as the conformation measured by NMR. Our simulations suggest a kinetic mechanism that consists of three separate stages. The time-scales associated with these stages are distinct and depend differently on temperature. Electrostatic interactions play an initial role in guiding the peptide chain to a roughly correct structure as measured by the end-to-end distance. At the same time or later the backbone torsions rearrange due to local tendency of the proline ring to form a turn: this step depends on solvation forces and is helped by loose hydrophobic interactions. In the final step, hydrophobic residues pack against each other. We also show the existence of an off the pathway intermediate, suggesting that even in the folding of a small peptide "misfolded" structures can form. The simulations clearly show that parallel folding paths are involved. Our findings suggest that the process of peptide folding shares many of the features expected for the significantly larger protein molecules.  相似文献   

12.
13.
The solution behaviour of selected proteins has been studied under conditions promoting precipitation, binding to mildly hydrophobic adsorbents or partition. Solvophobic theory may be used to describe these forms of protein partition. The tendency of a protein to partition therein is dependent upon surface properties of the protein solute mediated by the concentration and nature of added salts. As applied to partitioning in poly(ethylene glycol) (PEG)-salt systems this implies that linear (Br?nsted) relationships apply only to proteins partitioned close to the critical point. At longer tie-line lengths protein partitioning is increasingly influenced by salting-out forces. This is confirmed by the observed behaviour of the proteins. The point at which this behaviour changes has been unambiguously defined enabling the direct comparison of phase transition of proteins during partition in all systems. The results obtained show that phase transition during adsorption and partition occur at similar concentrations of salt. This is less than that required to promote precipitation. It appears, from these limited studies, that top-phase preferring proteins are partitioned at salt concentrations above those required to cause adsorption. Proteins preferring the lower phase are partitioned at salt concentrations close to or below those required for adsorption. This raises questions regarding the solvated molecular form of the partitioned proteins and the definition of the partition coefficient.  相似文献   

14.
Binding of calcium to calmodulin (CAM) induces specific structural rearrangements in the whole protein molecule. Ca2+ organizes and stabilizes the four-domains structure of calmodulin in a helical, active conformation that can bind to its target proteins; the central helix remaining flexible is an essential condition for their bio-recognition. The conformation of calmodulin, and its efficacy to interact with target proteins, is profoundly altered when bound to metal ions other than calcium. As recently reported, the local structural changes of CaM, which occur upon aluminium binding, lead to the impairment of protein flexibility and to the loss of its ability to interact with several other proteins, which may decrease or inhibit the regulatory character of calmodulin. In this study we followed conformational changes occurring in the calmodulin molecule after aluminium binding using highly specific monoclonal antibodies (mAbs) able to differentiate between the conformational states of calmodulin, as well as mAbs which recognize aluminium free or bound to proteins. Under the same experimental conditions, mAb CAM-1, a Ca2+ conformation sensitive antibody raised against calmodulin, fails to recognize the calmodulin-aluminium complex, despite the presence of Ca2+, while the anti-Al antibodies show a maximal binding pattern towards their antigen. These data suggest that Al3+ ions bind to calmodulin in the presence of Ca2+ ions, leading to an inactive, reversible conformation, instead of its physiological active form. Alteration of the conformation of calmodulin imposed by Al binding may have possible implications in the neurotoxicity mechanism related to Alzheimer's disease.  相似文献   

15.
The enthalpies of interaction of urea with five globular proteins, ribonuclease A, trypsin, beta-lacto-globulin, ovalbumin and bovine serum albumin have been measured in aqueous solution at pH 7.0, I=0.005 M and 25 degrees C over a range of urea molality m from 0-15 mmol g-1 (where a 1 molal solution contains 1 mmol g-1). For all the proteins the interaction is exothermic, and there is an appreciable heat evolution at low urea concentrations, m less than 5 mmol g-1, which increases sharply at higher urea concentrations when the proteins undergo unfolding. If account is taken of the endothermic enthalpies of unfolding of the native proteins, the enthalpies of interactions of urea per unit mass denatured protein lie in the range -45 to -75 J g-1, corresponding to an average binding enthalpy of -23 kJ mol-1 bound urea.  相似文献   

16.
Characterization and affinity applications of cellulose-binding domains   总被引:1,自引:0,他引:1  
Cellulose-binding domains (CBDs) are discrete protein modules found in a large number of carbohydrolases and a few nonhydrolytic proteins. To date, almost 200 sequences can be classified in 13 different families with distinctly different properties. CBDs vary in size from 4 to 20 kDa and occur at different positions within the polypeptides; N-terminal, C-terminal and internal. They have a moderately high and specific affinity for insoluble or soluble cellulosics with dissociation constants in the low micromolar range. Some CBDs bind irreversibly to cellulose and can be used for applications involving immobilization, others bind reversibly and are more useful for separations and purifications. Dependent on the CBD used, desorption from the matrix can be promoted under various different conditions including denaturants (urea, high pH), water, or specific competitive ligands (e.g. cellobiose). Family I and IV CBDs bind reversibly to cellulose in contrast to family II and III CBDs which are in general, irreversibly bound. The binding of family II CBDs (CBD(Cex)) to crystalline cellulose is characterized by a large favourable increase in entropy indicating that dehydration of the sorbent and the protein are the major driving forces for binding. In contrast, binding of family IV CBDs (CBD(N1)) to amorphous or soluble cellulosics is driven by a favourable change in enthalpy which is partially offset by an unfavourable entropy change. Hydrogen bond formation and van der Waals interactions are the main driving forces for binding. CBDs with affinity for crystalline cellulose are useful tags for classical column affinity chromatography. The affinity of CBD(N1) for soluble cellulosics makes it suitable for use in large-scale aqueous two-phase affinity partitioning systems.  相似文献   

17.
The mode of reconstitution of chicken erythrocyte and reticulocyte chromatin has been investigated. Chromatin was dissociated in 2 M NaCl, 5 M urea, and 0.01 M potassium phosphate (pH 7.2) and was dialyzed against various NaCl concentrations in 5 M urea and 0.01 M potassium phosphate (pH 7.2). Histone reassociation to DNA occurs with the binding of histone H5 at 0.5 M NaCl in 5 M urea, followed by histone H1 at 0.4 M NaCl in 5 M urea. All the classes of histones are reassociated with DNA at 0.2 M NaCl in 5 M urea and binding of all classes of histones is complete in 0.1 M NaCl and 5 M urea. Nonhistone proteins reassociate with DNA before and at the same time that histones reassociate with DNA. Binding of nonhistone proteins to DNA appears to be complete in 5 M urea and 0.01 M potassium phosphate (pH 7.2). There is also found in both erythrocyte and reticulocyte chromatin a nonhistone protein present in relatively high concentrations, which remains associated with DNA in 2 M NaCl and 5 M urea. This tightly bound protein appears as one major band when chromatographed on sodium dodecyl sulfate-polyacrylamide gels, with a molecular weight of 95 000. This protein is soluble in phenol and sodium dodecyl sulfate but is insoluble in 5 M urea or 4 M guanidine hydrochloride. A fraction of reticulocyte nonhistone proteins was found to bind to DNA-cellulose in 5 M urea. The majority of these proteins elute at 0.15 M NaCl in 5 M urea but a significant fraction elutes at NaCl concentrations at which the bulk of the histones do not bind to DNA. The proteins that bind to free DNA have low molecular weights and do not show species speciificity. Approximatley 50% of the reticulocyte nonhistone protein does not bind to a DNA-cellulose column in 5 M urea and may require histones for complete reassociation.  相似文献   

18.
Sodium dodecyl sulfate binds to S-carboxyamidomethyl-k-casein in a highly cooperative manner at a concentration near the critical micelle concentration, showing a strong dependence on ionic strength. The maximum number of sodium dodecyl sulfate molecules bound is attained above the critical micelle concentration, and is very close to the micelle aggregation number in the absence of protein. The binding sites on the protein for sodium dodecyl sulfate are localized mainly on para-k-casein part, which is a hydrophobic fragment of k-casein produced by rennin attack. The mode of the action of sodium dodecyl sulfate on S-carboxyamidomethyl-k-casein resembles that of several integral membrane proteins, rather than of water soluble proteins. On considering possible situations, it is suggested that the unusual interaction of S-carboxyamidomethyl-k-casein with sodium dodecyl sulfate is responsible for an anomalous migration of reduced k-casein observed in sodium dodecyl sulfate polyacrylamide gel electrophoresis. Further, the suggestion was made by the binding studies of sodium dodecyl sulfate and non-ionic detergents that the sites which were involved in self-association of S-carboxyamidomethyl-k-casein participated in the binding sites of detergents.  相似文献   

19.
Conformations of the globular protein staphylococcal enterotoxin B have been examined experimentally by ultraviolet circular dichroism (CD) and visible optical rotatory dispersion (ORD). Chen-Yang-Chau analysis (Chen, Y.-H., Yang, J.T., and Chau, K. H. (1974), Biochemistry 13, 3350) of the far-ultraviolet CD spectrum of native enterotoxin B revealed (assuming an average helix length of 11 residues) 9% alpha helix, 38% beta structure, and 53% random coil. A fourfold increase in alpha-helix was observed for enterotoxin exposed to 0.2% sodium dodecyl sulfate, behavior typical for globular proteins of low helical content. Values of -40 to -50 for the Moffitt-Yang parameter b0 calculated from visible ORD suggested 6-13% alpha helix in native enterotoxin. Application of a new predictive model (Chou, P. Y., and Fasman, G. D. (1974), Biochemistry 13,222) to the amino acid sequence of enterotoxin B indicated 11% alpha helix, 34% beta structure, and 55% coil in native enterotoxin. The excellent agreement for the amount of alpha and beta conformation utilizing different optical and predictive methods indicates beta structure as the dominant secondary structure in native enterotoxin B. Most of the beta structure is predicted by Chou-Fasman analysis to reside in two large regions of antiparallel beta sheet involving residues 81-148 and residues 184-217. Such highly cooperative regions of anti-parallel beta sheet account for the slow unfolding of enterotoxin B in concentrated guanidine hydrochloride and rapid folding of guanidine hydrochloride denatured enterotoxin B to native conformation(s) (Warren, J.R., Spero, L., and Metzger, J. F. (1974), Biochemistry 13, 1678). A more than twofold increase in alpha-helix content with a small diminution in beta structure was detected by CD and ORD upon acidification of aqueous enterotoxin to pH 2.5. Thus, the beta structure of enterotoxin B appears to resist isothermal denaturation and constitutes a stable interior core of structure in the enterotoxin molecule.  相似文献   

20.
采用t-BAMBP-磺化煤油系配置的有机相对含铷、铯的高盐废液进行除钾试验,然后进行萃取分离试验,分别探索了硫酸铝用量与pH对除钾的影响,萃取过程中料液pH、萃取剂浓度、萃取时间、萃取相比、萃取级数对铷、铯萃取率,以及洗涤过程中洗涤相比、洗涤级数对铷铯洗脱率的影响。研究表明,料液pH=12.5、萃取相比O/A=1、选用1 mol/L萃取剂在室温下萃取5 min,铯萃取率可达99.95%。对高盐废液进行11级萃取(6级萃取、3级洗涤、2级反萃)连续试验,有机相中铯萃取率达99%以上,65%的铷留在水相中,可以较好地将铷与铯分离。最终制备出纯度98.3%的氯化铯产品。  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