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.
Protein A and C, which are major components of the acidic proline-rich proteins in human saliva, were digested, before or after adsorption to hydroxyapatite, with alkaline phosphatase, trypsin, thermolysin and a proteinase preparation from salivary sediment. The results demonstrate that the binding site is located in the proline-poor N-terminal part of the protein, possibly between residues 3 and 25. Phosphoserine is necessary for maximal adsorption of the proteins to hydroxyapatite. When proteins A and C are adsorbed to hydroxyapatite before proteolytic digestion there is a protection of some of the susceptible bonds in the N-terminal part of the proteins and a gradual removal of the proline-rich C-terminal part. Thermolysin can cleave susceptible bonds in the part of the protein that remains bound to hydroxyapatite, but at least some of the resulting peptides are retained on the mineral. Since the ability of the proteins to inhibit hydroxyapatite formation and to bind calcium is located in the N-terminal proline-poor part, it is possible that these activities are retained after proteolytic digestion of the adsorbed proteins.  相似文献   

2.
Statherin is a salivary protein that inhibits the nucleation and growth of hydroxyapatite crystals in the supersaturated environment of the oral cavity. The thermodynamics of adsorption of statherin onto hydroxyapatite crystals have been characterized here by isothermal titration calorimetry and equilibrium adsorption isotherm analysis. At 25 degrees C, statherin adsorption is characterized by an exothermic enthalpy of approximately 3 kcal/mol that diminishes to zero at approximately 25% surface coverage. The initial heat of statherin adsorption increases with temperature, displaying a positive heat capacity change of 194 +/- 7 cal K(-)(1) mol(-)(1) at 25 degrees C. The heat of adsorption during this initial phase is strongly dependent on the buffer species, and from the differential heats of buffer ionization, it can be calculated that approximately one proton is taken up by the crystal or protein upon adsorption. The free energy of adsorption is dominated at all coverages by a large positive entropy (>or=23 cal K(-)(1) mol(-)(1)), which may be partially due to the loss of organized water that hydrates the protein and the mineral surface prior to adsorption. These results are interpreted using a two-site model for adsorption of statherin onto the hydroxyapatite crystals.  相似文献   

3.
Osteonectin, an acidic noncollagenous protein of bone and dentin, has affinity to hydroxyapatite crystals. Binding sites to hydroxyapatite of this protein were determined by a proteolytic experiment and an in vitro binding experiment using synthetic peptide analogues. Osteonectin was adsorbed on hydroxyapatite crystals and digested with trypsin. A peptide was left adsorbed on the crystal even after the digestion. The peptide was identified as an amino terminal peptide containing glutamic acid-rich sequences, which have been assumed to be possible hydroxyapatite-binding sites. Poly glutamic acid sequences were synthesized as models of the binding sites. Glu6 peptide was bound to the hydroxyapatite with a dissociation constant of 2.4 μM. Peptides containing fewer glutamic acids had lower affinity to the crystal. Effects of these peptides on in vitro mineralization were examined by a gel system in microtiter plates. The Glu6 peptide had a positive effect on the mineralization in this system, whereas Asp6 peptide had a negative effect. These effects indicate the presence of an interaction between these peptides and mineral crystals.  相似文献   

4.
Microcalorimetric titrations of bullfrog (Rana catesbeiana) skeletal troponin C with Ca2+ were carried out in the absence of Mg2+ at 25 degrees C and at pH 7.0. The observed enthalpy titration curve was divided into three stages. The first stage of the titration (up to 2 mol of Ca2+/mol of protein) was characterized as an extremely exothermic process (delta H = -52 kJ/mol of site), the second one (titration from 2 to 3 mol of Ca2+/mol of protein) as a weakly endothermic process (delta H = +26 kJ/mol of site), and the final one (over 3 mol of Ca2+/mol of protein) as a moderately exothermic process (delta H = -35 kJ/mol of site). The endothermic process of Ca2+ binding to the third site (the second stage) has the same property as that of the Ca2+ binding to every site of calmodulin but is distinctly different from those of the calmodulin-trifluoperazine complex and parvalbumins. This may suggest that an endothermic nature of Ca2+ binding, the reaction being driven solely by entropy change, is characteristic of the regulatory reactions of Ca2+ binding proteins accompanying the interaction with other proteins. The third Ca2+ binding site of bullfrog troponin C is, therefore, possibly involved in the regulation of muscle contraction.  相似文献   

5.
Human saliva contains a large number of phosphopeptides derived by cleavage of acidic proline-rich proteins (APRPs). These peptides were purified by column chromatography and they constituted 0.5% of APRPs in parotid saliva, but 11% of APRPs in saliva expectorated from the mouth (whole saliva), indicating that there is considerable cleavage of APRPs after secretion from the gland. Similarly to APRP, the phosphopeptides bind Ca2+, but they accounted for only 4% of protein-bound Ca2+ in whole saliva. APRPs as well as the phosphopeptides inhibited formation of hydroxyapatite, but, whereas 19-20 micrograms of APRP was needed for 50% inhibition, only 0.7-3.3 micrograms of purified peptides was needed for the same degree of activity, and the phosphopeptides accounted for 18% of total inhibitory activity in whole saliva. All phosphopeptides adsorbed on hydroxyapatite in vitro, and adsorption of phosphopeptides on tooth surfaces in vivo could also be demonstrated, indicating that they would be able to inhibit unwanted mineral formation on the tooth surface in vivo. Degradation of APRPs after secretion therefore does not lead to a loss of their biological activities.  相似文献   

6.
To identify protein targets for calmodulin (CaM) in the cilia of Paramecium tetraurelia, we employed a 125I-CaM blot assay after resolution of ciliary proteins on SDS/polyacrylamide gels. Two distinct types of CaM-binding proteins were detected. One group bound 125I-CaM at free Ca2+ concentrations above 0.5-1 microM and included a major binding activity of 63 kDa (C63) and activities of 126 kDa (C126), 96 kDa (C96), and 36 kDa (C36). CaM bound these proteins with high (nanomolar) affinity and specificity relative to related Ca2+ receptors. The second type of protein bound 125I-CaM only when the free Ca2+ concentration was below 1-2 microM and included polypeptides of 95 kDa (E95) and 105 kDa (E105). E105 may also contain Ca2+-dependent binding sites for CaM. Both E95 and E105 exhibited strong specificity for Paramecium CaM over bovine CaM. Ciliary subfractionation experiments suggested that C63, C126, C96, E95, and E105 are bound to the axoneme, whereas C36 is a soluble and/or membrane-associated protein. Additional Ca2+-dependent CaM-binding proteins of 63, 70, and 120 kDa were found associated with ciliary membrane vesicles. In support of these results, filtration binding assays also indicated high-affinity binding sites for CaM on isolated intact axonemes and suggested the presence of both Ca2+-dependent and Ca2+-inhibitable targets. Like E95 and E105, the Ca2+-inhibitable CaM-binding sites showed strong preference for Paramecium CaM over vertebrate CaM and troponin C. Together, these results suggest that CaM has multiple targets in the cilium and hence may regulate ciliary motility in a complex and pleiotropic fashion.  相似文献   

7.
A double isotope technique and EGTA buffers were used to measure the binding of Ca2+ to rabbit psoas muscle fibers extracted with detergent and glycerol. These experiments were designed to test the effect of rigor complex formation, determined by the degree of filament overlap, on the properties of the Ca2+-binding sites in the intact filament lattice. In the presence of 5 mM MgCl2 (no ATP), reduction of filament overlap was associated with a reduced binding of Ca2+ over the entire range of free Ca2+ concentrations (5.10(-8)-2.10(-5) M). With maximum filament overlap (sarcomere length 2.1-2.2 micrometer) the maximum bound Ca2+ was equivalent to 4 mol Ca2+/mol troponin and there was significant positive interaction between binding sites, as shown by Scatchard and Hill plots. With no filament overlap (sarcomere length 3.8-4.4 micrometer) the maximum bound Ca2+ was equivalent to 3 mumole Ca2+/mol troponin and graphical analysis indicated a single class of non-interacting sites. The data provide evidence that when cross-bridge attachments between actin and myosin filaments are formed not only does an additional Ca2+ binding site appear, but cooperative properties are imposed upon the binding sites.  相似文献   

8.
Microcalorimetric titrations of calmodulin with Ca2+ and trifluoperazine (TFP) at various molar ratios have been carried out at 25 degrees C and at pH 7.0. Ca2+ binding to calmodulin produces heat (-delta H) in the presence of TFP, while heat is absorbed in the absence of TFP. The total heat produced by Ca2+ binding to all four sites is increased at increasing TFP-to-calmodulin ratios, attaining a plateau at about 7. These results indicate that at the higher ratios, the enthalpy changes (delta H) associated with Ca2+ binding are affected by TFP molecules bound at both high- and low-affinity sites. In addition, the Ca2+ binding reaction of the calmodulin-TFP complex is driven solely by a favorable enthalpy change of -27 kJ/mol of site; the entropy change (delta S) is -35 J/mol/K. These thermodynamic changes are opposite to those for TFP-free calmodulin and distinctly different from other Ca2+ binding proteins such as skeletal and cardiac troponin C and parvalbumin, where the reaction is driven by favorable changes of entropy as well as enthalpy.  相似文献   

9.
The criteria for elution of proteins from hydroxyapatite columns were examined as a function of (1) protein isoelectric point (22 proteins with isoelectric points between 3.5 and 11.0); (2) ionic nature of eluant (Na salts of PO4, F-, Cl-, SCN-, ClO-4, and CaCl2); and (3) structural differences between related proteins. It was found that proteins can be classified into three groups: (1) basic proteins, which elute at similar, moderate molarities of PO4, F-, Cl-, SCN-, and ClO-4, and low (less than 0.003 M) Ca2+; (2) acidic proteins which elute at about equal moderate molarities of PO4 and F-, but do not elute with Ca2+ and usually not with Cl-; (3) neutral proteins, which elute with PO4, F-, and Cl-, but show a strong anion specificity, and do not elute with Ca2+ or SCN-. Furthermore, individual specific polar groups are not in general crucial to binding or desorption, and variations in structure, other than major loosening, do not influence strongly the pattern of protein-hydroxyapatite interaction.  相似文献   

10.
The present study showed that S. mutans and S. sanguis behaved like negatively-charged particles in their interaction with hydroxyapatite in vitro. Phosphate in the system inhibited bacterial uptake by apatite, whereas calcium increased the uptake. A layer of acidic protein inhibited the uptake of bacteria by hydroxyapatite. The opposite was true when a basic protein was first adsorbed to the apatite. A saliva film on the apatite decreased the uptake of bacteria, supporting the view that acidic proteins are selectively adsorbed by hydroxyapatite from saliva. The results indicate clearly that electrostatic forces may be involved in bacterial interaction with tooth surface.  相似文献   

11.
The NH2-terminal alpha fragments of human complement proteases C1-r and C1-s were obtained by limited proteolysis of the native proteins with trypsin, and isolated. C1-r alpha extended from residues 1 to 208 of C1-r A chain, with at least two cleavage sites within disulfide loops, after lysine 134 and arginine 202. C1-s alpha comprised residues 1-192 of the C1-s A chain, with one cleavage site within a disulfide loop, after arginine 186. C1-r alpha was monomeric either in the presence or absence of Ca2+ but formed Ca2(+)-dependent dimers with native C1-s. C1-s alpha dimerized in the presence of Ca2+ and formed Ca2(+)-dependent tetramers (C1-s alpha-C1-r-C1-r-C1-s alpha) with native C1-r. C1-r alpha and C1-s alpha associated in the presence of Ca2+ to form C1-r alpha-C1-s alpha heterodimers. Equilibrium dialysis studies indicated that each alpha region binds Ca2+ with a dissociation constant ranging from 19 microM (native proteins) to 38 microM (fragments). C1-r alpha, C1-r alpha-C1-s alpha, and the native C1-s-C1-r-C1-r-C1-s tetramer bound 0.9, 1.9, and 4.0 Ca2+ atoms/mol, respectively, whereas dimers C1-s alpha-C1-s alpha and C1-s-C1-s incorporated 2.9 and 3.0 Ca2+ atoms/mol. It is concluded that each alpha region contains one high affinity Ca2+ binding site. This 1:1 stoichiometry is maintained upon heterologous (C1-r-C1-s) interaction, whereas the homologous (C1-s-C1-s) interaction provides one additional binding site.  相似文献   

12.
A high affinity calcium binding site that is independent of the gamma-carboxyglutamic acid-rich amino-terminal region, has been demonstrated in bovine protein C, as well as in the other vitamin K-dependent proteins (except prothrombin) involved in blood coagulation. gamma-Carboxyglutamic acid-independent calcium binding in protein C is required for its rapid activation by the thrombin-thrombomodulin complex. We have now isolated a Ca2+-binding fragment from a tryptic digest of bovine protein C. The isolated fragment contains the two domains that are homologous to the epidermal growth factor precursor from the light chain of protein C, and a small disulfide bound peptide derived from the heavy chain. The isolated fragment bound 1 mol of Ca2+/mol of protein with a dissociation constant (Kd) of approximately 1 x 10(-4) M. This is similar to the Kd previously determined for binding of a single Ca2+ ion to protein C lacking the gamma-carboxyglutamic acid region. Immunochemical evidence indicated that Ca2+ binding induced a conformational change both in protein C lacking the gamma-carboxyglutamic acid region and in the isolated fragment.  相似文献   

13.
Brush border myosin I from chicken intestine is phosphorylated in vitro by chicken intestinal epithelial cell protein kinase C. Phosphorylation on serine and threonine to a maximum of 0.93 mol of P/mol of myosin I occurs within an approximately 20 kDa region at the end of the COOH-terminal tail of the 119-kDa heavy chain. The effects of Ca2+ on myosin I phosphorylation by protein kinase C are complex, with up to 4-fold stimulation occurring at 0.5-3 microM Ca2+, and up to 80% inhibition occurring at 3-320 microM Ca2+. Phosphorylation required that brush border myosin I be in its phosphatidylserine vesicle-bound state. Previously unknown Ca2+ stimulation of brush border myosin I binding to phosphatidylserine vesicles was found to coincide with Ca2+ stimulation of phosphorylation. A myosin I proteolytic fragment lacking approximately 20 kDa of its tail retained Ca(2+)-stimulated binding, but showed reduced Ca(2+)-independent binding. Ca(2+)-dependent phosphatidylserine binding is apparently due to the concomitant phosphatidylserine-promoted, Ca(2+)-induced dissociation of up to three of the four calmodulin light chains from myosin I. Four highly basic putative calmodulin-binding sites in the Ca(2+)-dependent phosphatidylserine binding region of the heavy chain were identified based on the similarity in their sequence to the calmodulin- and phosphatidylserine-binding site of neuromodulin. Calmodulin dissociation is now shown to occur in the low micromolar Ca2+ concentration range and may regulate the association of brush border myosin I with membranes and its phosphorylation by protein kinase C.  相似文献   

14.
A new low molecular weight calcium binding protein, designated 12-kDa CaBP, has been isolated from chicken gizzard using a phenyl-Sepharose affinity column followed by ion-exchange and gel filtration chromatographies. The isolated protein was homogeneous and has a molecular weight of 12,000 based on sodium dodecyl sulfate-gel electrophoresis. The amino acid composition of this protein is similar to but distinct from other known low molecular weight Ca2+ binding proteins. Ca2+ binding assays using Arsenazo III (Sigma) indicated that the protein binds 1 mol of Ca2+/mol of protein. The 12-kDa CaBP underwent a conformational change upon binding Ca2+, as revealed by uv difference spectroscopy and circular dichroism studies in the aromatic and far-ultraviolet range. Addition of Ca2+ to the 12-kDa CaBP labeled with 2-p-toluidinylnaphthalene-6-sulfonate (TNS) resulted in a sevenfold increase in fluorescence intensity, accompanied by a blue shift of the emission maximum from 463 to 445 nm. Hence, the probe in the presence of Ca2+ moves to a more nonpolar microenvironment. Like calmodulin and other related Ca2+ binding proteins, this protein also exposes a hydrophobic site upon binding calcium. Fluorescence titration with Ca2+ using TNS-labeled protein revealed the presence of a single high affinity calcium binding site (kd approximately 1 x 10(-6) M).  相似文献   

15.
Surveys of X-ray structures of Ca2+-containing and lanthanide ion-containing proteins and coordination complexes have been performed and structural features of the metal binding sites compared. A total of 515 structures of Ca2+-containing proteins were considered, although the final data set contained only 44 structures and 60 Ca2+ binding sites with a total of 323 ligands. Eighteen protein structures containing lanthanide ions were considered with a final data set containing eight structures and 11 metal binding sites. Structural features analysed include coordination numbers of the metal ions, the identity of their ligands, the denticity of carboxylate ligands, and the type of secondary structure from which the ligands are derived. Three general types of calcium binding site were identified in the final data set: class I sites supply the Ca2+ ligands from a continuous short sequence of amino acids; class II sites have one ligand supplied by a part of the amino acid sequence far removed from the main binding sequence; and class III sites are created by amino acids remote from one another in the sequence. The abundant EF-hand type of Ca2+ binding site was under-represented in the data set of structures analysed as far as its biological distribution is concerned, but was adequately represented for the chemical survey undertaken. A turn or loop structure was found to provide the bulk of the ligands to Ca2+, but helix and sheet secondary structures are slightly better providers of bidentate carboxylate ligation than turn or loop structures. The average coordination number for Ca2+ was 6.0, though for EF-hand sites it is 7. The average coordination number of a lanthanide ion in an intrinsic protein Ca2+ site was 7.2, but for the adventitious sites was only 4.4. A survey of the Cambridge Structural Database showed there are small-molecule lanthanide complexes with low coordination numbers but it is likely that water molecules, which do not appear in the electron density maps, are present for some lanthanide sites in proteins. A detailed comparison of the well-defined Ca2+ and lanthanide ion binding sites suggests that a reduction of hydrogen bonding associated with the ligating residues of the binding sites containing lanthanide ions may be a response to the additional positive charge of the lanthanide ion. Major structural differences between Ca2+ binding sites with weak and strong binding affinities were not obvious, a consequence of long-range electrostatic interactions and metal ion-induced protein conformational changes modulating affinities.  相似文献   

16.
The thermodynamic change in the binding of Ca2+ to a mutant human lysozyme having an engineered Ca2+ binding site (Kuroki, R., Taniyama, Y., Seko, C., Nakamura, H., Kikuchi, M., and Ikehara, M. (1989) Proc. Natl. Acad. Sci. U. S. A. 86, 6903-6907) was analyzed by calorimetry and interpreted in terms of structural information obtained from x-ray crystallography. It was found that the enthalpic contribution for the Ca2+ binding reaction was small, driven primarily by entropy release (10 kcal/mol). This release of entropy was also observed in some organic chelators. Moreover, through the information of the tertiary structures of the apo- and holomutant lysozyme, it was confirmed that the entropy release (10 kcal/mol) upon the binding of Ca2+ arises primarily from the release of bound water molecules hydrating the free Ca2+. Previous studies of Ca2+ binding to proteins have involved significant changes in protein conformation. They can now be reevaluated to determine the contribution of conformational changes to Ca2+ binding. After removing the thermodynamic contribution of Ca2+ binding itself, it is found that upon the binding of Ca2+ the enthalpy change is negative but is almost compensated by the negative entropy change. The negative change in both enthalpy and entropy is characteristic of values seen in the thermodynamic change upon the folding of proteins.  相似文献   

17.
Oriented immobilization of proteins is an important step in creating protein-based functional materials. In this study, a method was developed to orient proteins on hydroxyapatite (HA) surfaces, a widely used bone implant material, to improve protein bioactivity by employing enhanced green fluorescent protein (EGFP) and β-lactamase as model proteins. These proteins have a serine or threonine at their N-terminus that was oxidized with periodate to obtain a single aldehyde group at the same location, which can be used for the site-specific immobilization of the protein. The HA surface was modified with bifunctional hydrazine bisphosphonates (HBPs) of various length and lipophilicity. The number of functional groups on the HBP-modified HA surface, determined by a 2,4,6-trinitrobenzenesulfonic acid (TNBS) assay, was found to be 2.8 × 10(-5) mol/mg of HA and unaffected by the length of HBPs. The oxidized proteins were immobilized on the HBP-modified HA surface in an oriented manner through formation of a hydrazone bond. The relative protein immobilization amounts through various HBPs were determined by fluorescence and bicinchoninic acid (BCA) assay and showed no significant effect by length and lipophilicity of HBPs. The relative amount of HBP-immobilized EGFP was found to be 10-15 fold that of adsorbed EGFP, whereas the relative amount of β-lactamase immobilized through HBPs (2, 3, 4, 6, and 7) was not significantly different than adsorbed β-lactamase. The enzymatic activity of HBP-immobilized β-lactamase was measured with cefazolin as substrate, and it was found that the catalytic efficiency of HBP-immobilized β-lactamase improved 2-5 fold over adsorbed β-lactamase. The results obtained demonstrate the feasibility of our oriented immobilization approach and showed an increased activity of the oriented proteins in comparison with adsorbed proteins on the same hydroxyapatite surface matrix.  相似文献   

18.
Flow dialysis measurements of calcium binding to bovine brain S100 alpha alpha, S100a (alpha beta), and S100b (beta beta) proteins in 20 mM Tris-HCl buffer at pH 7.5 and 8.3 revealed that S100 proteins bind specifically 4 Ca2+ eq/mol of protein dimer. The specific calcium-binding sites had, therefore, been assigned to typical amino acid sequences on the alpha and beta subunit. The protein affinity for calcium is much lower in the presence of magnesium and potassium. Potassium strongly antagonizes calcium binding on two calcium-binding sites responsible for most of the Ca2+-induced conformational changes on S100 proteins (probably site II alpha and site II beta). Zinc-binding studies in the absence of divalent cations revealed eight zinc-binding sites/mol of S100b protein dimer that we assumed to correspond to 4 zinc-binding sites/beta subunit. Zinc binding to S100b studied with UV spectroscopy methods showed that the occupation of the four higher affinity sites and the four lower affinity sites on the protein dimer were responsible for different conformational changes in S100b structure. Zinc binding on the higher affinity sites regulates calcium binding to S100b by increasing the protein affinity for calcium and decreasing the antagonistic effect of potassium on calcium binding. Zinc-binding studies on S100a and S100 alpha alpha protein showed that the Trp-containing S100 proteins bind zinc more weakly than S100b protein. Calcium-binding studies on zinc-bound S100a proved that calcium- and zinc-binding sites were distinct although there was no increase in zinc-bound S100a affinity for calcium, as in S100b protein. Finally we provide evidence that discrepancies between previously published results on the optical properties of S100b protein probably result from oxidation of the sulfhydryl groups in the protein.  相似文献   

19.
La ions in precipitated hydroxyapatites.   总被引:3,自引:0,他引:3  
Hydroxyapatites were synthesized by precipitation from an aqueous solution with La3+ (0-0.75%) and with carbonate (0-6.1%) at controlled pH 7.0. Uptake of La3+ was 90-95% complete. Relatively low Ca/P (1.54-1.63) ratios were attributed to nonstoichiometry. Carbonate in samples was identified by IR spectroscopy as B-type carbonate. Lattice parameters of the hexagonal apatite structure were not affected by the La3+ content. Noncarbonated samples heated to 800 degrees C transform partially to beta-Ca3(PO4)2. Thermogravimetric analysis showed release of 0.4 mol adsorbed and 1 mol crystalline water up to 400 degrees C and decomposition of carbonate up to 900 degrees C in the samples. Luminescence data obtained for Gd-containing hydroxyapatites prove that Gd3+ ions are not incorporated in the precipitated hydroxyapatite. These findings suggest that, in the La-containing samples, La3+ is surface absorbed and not incorporated in hydroxyapatite.  相似文献   

20.
Synopsis In order to study the reactions involved in some of the histochemical procedures used for demonstrating calcium in calcified tissues, it was considered appropriate to use well characterized synthetic hydroxyapatite in the first instance. In the first paper of this series (Speirs, 1970), it was found that many dyes not previously used in histochemistry were capable of staining hydroxyapatite; the purpose of the present paper is to describe the numerous experimental approaches that have been made in an attempt to elucidate the mechanisms involved in the adsorption of some of these dyes by hydroxyapatite. Dyes have been grouped according to their adsorption curves (in which dye uptake by solid was plotted against the concentration of dye in solution at equilibrium). From these graphs, predictions and calculations were made concerning the orientation of the dye molecules on the surface of hydroxyapatite, the type of bonding possibly involved and the area of surface covered by each molecule. These were then related to the dimensions and structure of the dye molecules. Saturation of surface sites was achieved in the adsorption of some dyes and the nature of these sites was investigated by studying (1) competition between several dyes for the surface, (2) the accessibility of surface calcium and phosphorus in stained and unstained hydroxyapatite, and (3) the release of32P from surface labelled hydroxyapatite during dye adsorption. Most of the dyes adsorbed from 95% ethanol were displaced relatively easily by treatment with 0.5 mM phosphate in ethanol, but those adsorbed from tris buffer, pH 7.45, were more stable when exposed to phosphate in tris. Treatment of stained hydroxyapatite with solvents containing 0.5 mM calcium reduced the rate of elution of the dyes. Convincing evidence for chelation, hydrogen bonding, ion exchange and physical adsorption processes as the mechanisms of adsorption has not been obtained. Future studies to investigate these processes are discussed.  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