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.
Laser nephelometry is a suitable technique for the quantitative determination and differentiation of both lectins and glycoconjugates in the low-picomolar range. Simultaneously this method renders possible investigations on the specificity and mode of interaction between lectins and different ligands. The results demonstrate that the degree of co-operativity between concanavalin A and the respective glycoconjugate is dependent on the presence of hydrophobic binding sites and can be substantially altered by conformational changes of the ligand. The transition from apotransferrin to Fe3+-transferrin induces a transformation of the sigmoidal-shaped binding curve to a hyperbolic one. Hence, at low concentrations, Fe3+-transferrin is bound far better than apotransferrin, whereas maximal binding is nearly identical. After removal of N-acetylneuraminate, concanavalin A is less efficient in differentiating between the Fe3+-charged and Fe3+-free (apo) forms of transferrin.  相似文献   

2.
There is an increasing interest in the use of lanthanides in medicine. However, the mechanism of their accumulation in cells is not well understood. Lanthanide cations are similar to ferric ions with regard to transferrin binding, suggesting transferrin-receptor mediated transport is possible; however, this has not yet been confirmed. In order to clarify this mechanism, we investigated the binding of Yb3+ to apotransferrin by UV-Vis spectroscopy and stopped-flow spectrophotometry, and found that Yb3+ binds to apotransferrin at the specific iron sites in the presence of bicarbonate. The apparent binding constants of these sites showed that the affinity of Yb3+ is lower than that of Fe3+and binding of Yb3+ in the N-lobe is kinetically favored while the C-lobe is thermodynamically favored. The first Yb3+ bound to the C-lobe quantitatively with a Yb/apotransferrin molar ratio of < 1, whereas the binding to the other site is weaker and approaches completeness by a higher molar ratio only. As demonstrated by 1H NMR spectra, Yb3+ binding disturbed the conformation of apotransferrin in a manner similar to Fe3+. Flow cytometric studies on the uptake of fluorescein isothiocyanate labeled Yb3+-bound transferrin species by K562 cells showed that they bind to the cell receptors. Laser scanning confocal microscopic studies with fluorescein isothiocyanate labeled Yb3+-bound transferrin and propidium iodide labeled DNA and RNA in cells indicated that the Yb3+ entered the cells. The Yb3+-transferrin complex inhibited the uptake of the fluorescein labeled ferric-saturated transferrin (Fe2-transferrin) complex into K562 cells. The results demonstrate that the complex of Yb3+-transferrin complex was recognized by the transferrin receptor and that the transferrin-receptor-mediated mechanism is a possible pathway for Yb3+ accumulation in cells.  相似文献   

3.
Effects of protease inhibitors on liver regeneration   总被引:2,自引:0,他引:2  
The oxidation of Fe2+ to Fe3+ by oxygen at pH 7.45 is a first order reaction with a 25 minute half life. In the presence of apotransferrin the oxidation rate is greatly enhanced and Fe3+-transferrin is formed. The apotransferrin mediated reaction reaches 50% completion in one minute; it does not follow simple first order kinetics. Iron-saturated transferrin does not exhibit the rate enhancement effect suggesting that the specific metal binding sites are the loci of the iron oxidation. Addition of H2O2, an agent which rapidly oxidizes Fe2+ to Fe3+, during the reaction of Fe2+ with apotransferrin greatly decreases the yield of Fe3+-transferrin. It is postulated that the basis of the rate enhancement effect is the binding of Fe2+ to the metal binding site of the transferrin molecule, followed by a rapid oxidation of the iron to the trivalent form.  相似文献   

4.
The sedimentation behavior of 125I-labeled gastrin has been studied as a function of Fe3+ ion concentration and pH. Both sedimentation velocity and sedimentation equilibrium experiments indicated that high-molecular-weight Fe3+-gastrin complexes were formed at pH 5.0 and pH 7.4. Self-association of gastrin alone was observed at pH values below 5.0. 125I-labeled gastrin bound to human serum apotransferrin at pH 7.4. Scatchard analysis of the gastrin-apotransferrin complex gave a Kd of approximately 6.4 microM at 37 degrees C, with two binding sites per molecule of apotransferrin. No significant binding of gastrin to diferric transferrin was observed under the same conditions. The binding of gastrin to apotransferrin was inhibited by NaCl. The results are consistent with the hypothesis that gastrin and transferrin act synergistically in the uptake of dietary iron by the gastrointestinal tract.  相似文献   

5.
A detailed study of the exchange of Fe3+ between pyrophosphate and human serum transferrin was undertaken to test the hypothesis of a generalized reaction route for exchange of Fe3+ between transferrin and chelators. The initial rate of Fe3+ transfer from pyrophosphate to apotransferrin-CO2-3 is highly sensitive to the pyrophosphate to iron ratio with a maximal rate being observed at a ratio of 3:1, consistent with the presence of slowly reactive polymeric species at ratios less than 3:1 as revealed by EPR and kinetic measurements. At a ratio of 4:1 the reaction is distinctly biphasic. The rapid first phase results in the formation of an intermediate postulated as a mixedligand complex of the type PPi-Fe3+-transferrin-CO2-3. The intermediate has a distinct EPR spectrum and an absorption spectrum similar to that of Fe3+-transferrin-CO2-3, but with a spectral maximum at 450 nm rather than 465 nm. The second phase principally arises from the slow reaction of polymeric iron-pyrophosphate with the apoprotein and has contributions from the breakdown of the intermediate formed in the first phase. The rate of formation of the intermediate shows a hyperbolic dependence on NaHCO3 and apotransferrin concentrations, the latter suggesting a rate-limiting labilization of Fe3+(PPi)3, perhaps to form species of the type Fe3+(PPi)2, prior to attack by apotransferrin-CO2-3. Multimixing stopped flow spectrophotometry was employed to test the chemical reactivity of the Fe3+ to reduction at various times during the first phase. Surprisingly, a diminution of reactivity of 1000-fold was noted after only 2% of the first phase was completed, indicating a fast initial reaction which is not observed by normal rapid flow spectrophotometry. This initial reaction may involve the binding of iron-pyrophosphate to allosteric sites on the protein. The kinetics of iron removal from Fe3+-transferrin-CO2-3 by PPi are consistent with a rate-limiting conformational change in the protein as proposed earlier.  相似文献   

6.
Reticulocyte binding of Fe(III)_-transferrin and transferrin complexes with other metal ions have been compared by different investigators. The functional relevance of this comparison is not clear, therefore transferrin complexes with Fe(III), Cu(II), Mn(II) and Zn(II) have been studied further by DEAE-cellulose chromatography and by measurement of transferrin and metal uptakes by rabbit reticulocytes.Human Fe-transferrin behaved as a weaker anion than apotransferrin during DEAE-cellulose chromatography; since Fe-transferrin has a higher negative charge than apotransferrin and behaves a as stronger anion in electrophoretic systems, the chromatographic result was the opposite of that anticipated. The lower affinity of human Fe-transferrin for DEAE-cellulose is probably caused by a redistribution of charged groups on the surface of transferrin molecules when Fe(III) ions are bound and is therefore considered to be dependent on molecular conformation. Apotransferrin and divalent metal-transferrin complexes were found to have nearly equal affinities for DEAE-cellulose, thus the effect on surface charge of human transferrin molecules induced by binding Fe(III) appeared to be limited to that metal ion.Iron uptake by reticulocytes was associated with increased binding of transferrin to the cell surface: uptake of divalent metals occured without a concomitant increase in transferrin uptake or evidence of a specific metal-transfer process. Cu-transferrin was rapidly dissociated during incubation with cells.The effect of Fe(III)_binding on human transferrin molecules was to alter the molecular affinity for charged surfaces, namely DEAE-cellulose and reticulocyte membranes. This was less apparent with rabbit transferrin. Transferrin complexes with divalent metals behaved as apotransferrin in the process of association with reticulocytes.  相似文献   

7.
Harris (Biochemistry 24 (1985) 7412) reports that inorganic anions bind to human apotransferrin in such a way as to perturb the ultraviolet spectrum. The locus of binding is thought to involve the specific metal/anion-binding sites since no perturbation is observed with Fe3+-transferrin-CO3(2-). Paradoxically, we were unable to demonstrate the formation of Fe3+-transferrin-inorganic anion complexes despite the presence of high concentrations of SO4(2-), H2PO4-, Cl-, ClO4- or NO3-. Similar results were found for human lactoferrin. Electron paramagnetic resonance spectroscopy and visible spectrophotometry were used to monitor the results. An attempt to form the H2PO4- complex by displacement of glycine from Fe3+-transferrin-glycine resulted only in the disruption of the ternary complex. A series of inorganic anions varied in their ability to release iron from Fe3+-transferrin-CO3(2-) at pH 5.5, the approximate pH of endosomes where iron release takes place within cells. The order of effectiveness was H2P2O7(2-) much greater than H2PO4- greater than SO4(2-) greater than NO3- greater than Cl- greater than ClO4-. The rate of iron removal from Fe3+-transferrin-CO3(2-) at pH 5.5 by a 4-fold excess of pyrophosphate was greatly enhanced by physiological NaCl concentration. Iron removal was complete within 10 min, the approximate time for iron release from Fe3+-transferrin-CO3(2-) in developing erythroid cells. Thus, inorganic anions may have a significant effect on the release of iron under physiological conditions despite the fact that such inorganic anions cannot act as synergistic anions. The results are discussed in relation to a special role for the carboxylate group in allowing ternary complex formation.  相似文献   

8.
A series of new iron(III) and copper(II) complexes of bovine serum transferrin (BTf), with carbonate and/or oxalate as the synergistic anion, are presented. The complexes [Fe(2)(CO(3))(2)BTf], [Fe(2)(C(2)O(4))(2)BTf], [Cu(2)(CO(3))(2)BTf] and [Cu(C(2)O(4))BTf] were prepared by standard titrimetric techniques. The oxalate derivatives were also obtained from the corresponding carbonate complexes by anion-displacement. The site-preference of the transition metal-oxalate synergism has facilitated the preparation and isolation of the mononuclear complex [Cu(C(2)O(4))BTf], the mixed-anion complexes [Cu(2)(CO(3))(C(2)O(4))BTf] and [Fe(2)(CO(3))(C(2)O(4))BTf] and the mixed-metal complex [FeCu(C(2)O(4))(2)BTf]. The sensitivity of electron paramagnetic resonance (EPR) spectroscopy to the nature of the synergistic anions at the specific-binding sites of the transferrins has made this physical technique particularly indispensable to this study. None of the other members of the transferrin family of proteins has ever been demonstrated to bind the ferric and cupric ions one after the other, each occupying a separate specific-binding site of the same transferrin molecule, as a response to the coordination restrictions imposed by the oxalate ion. The bathochromic shift of the visible p(pi)-d(pi*) CT band for iron(III)-BTf and the hypsochromic shift of the p(pi)-d(sigma*) CT band for copper(II)-BTf, on replacing carbonate by oxalate as the associated anion, are consistent with the relative positions of these anionic ligands in the spectrochemical series and the nature of the d-type acceptor orbitals involved in the CT transitions. The binding and spectroscopic properties of bovine serum transferrin--a serum transferrin--very nearly mirror those of human serum transferrin, but differ significantly from those of human lactoferrin.  相似文献   

9.
Binding of vanadate to human serum transferrin   总被引:1,自引:0,他引:1  
Human serum transferrin specifically and reversibly binds 2 equiv of vanadate at the two metal-binding sites of the protein. The vanadium(V)-transferrin complex can be formed either by the addition of vanadate to apotransferrin or by the air oxidation of the vanadyl(IV)-transferrin complex. The formation of the vanadium complex can be blocked by loading the apotransferrin with iron(III), and bound vanadium can be displaced from the protein by the subsequent addition of either gallium(III) or iron(III). The binding constant for the second equiv of vanadate is 10(6.5) in 0.1 M hepes, pH 7.4 at 25 degrees C. The binding constant for the first equiv of vanadate is probably very similar, although no quantitative value could be determined. Although transferrin reacts with the vanadate anion, studies on the transferrin model compound ethylenebis(o-hydroxyphenylglycine) indicate that at pH 9.5, the vanadium is binding at the metal-binding site as a dioxovanadium(V) cation coordinated to two phenolic residues at each binding site. This bound cation appears to be protonated over the pH range 9.5-6.5, as shown by changes in the difference uv spectrum of the transferrin complex, to produce an oxohydroxo species. Further decreases in the pH lead to dissociation of the vanadium-transferrin complex.  相似文献   

10.
D C Harris 《Biochemistry》1977,16(3):560-564
Transferrin, the serum serum iron-transport protein which can bind two metal ions at physiologic pH, binds just one Fe3+, VO2+, or Cr3+ ion at pH 6.0. Fe3+ and VO2+ appear to be bound at the same site, designated A, based on electron paramagnetic resonance (EPR) spectra of VO2+-transferrin and (Fe3+)1(VO2+)1-transferrin. The EPR spectra of (Cr3+)1(VO2+)1-transferrin and of (Cr3+), (FE3+)1-transferrin indicate that that Cr3+ is bound to site B at pH 6.0. Transferrin was labeled at site A with 59Fe at pH 6.0 and at site B with 55Fe at pH 7.5. When the pH of the resulting preparation was lowered to 6.3 and the dissociated iron was separated by gel filtration, about ten times as much 55Fe as 59Fe was lost. The same EPR and isotopic-labeling experiments showed that Fe3+ added to transferrin at pH 7.5 binds to site A with about 90% selectivity.  相似文献   

11.
Vanadium associates with serum transferrin of rats administered vanadyl(IV) sulfate or ammonium metavanadate(V) by gastric intubation. Low molecular weight species account for only 3% of the vanadium present in plasma. The element distributes between the two major isotransferrins in proportion to their concentrations. Rat apotransferrin binds both vanadium(IV) and vanadium(V), forming 2:1 metal-protein complexes in both instances. Although the two isotransferrins apparently differ in their physiological properties, they exhibit identical vanadyl(IV) (VO2+) EPR spectra, indicating identical or very similar metal binding sites for both proteins. In contrast to other transferrins, the two sites of the rat protein are spectroscopically indistinguishable and exhibit a VO2+ EPR spectrum similar to that of the C-terminal metal binding site of human serum transferrin. VO2+ EPR signals are observed with liver, spleen, and kidney tissue samples from animals maintained on a vanadium-supplemented diet. These signals arise from a specific intracellular VO2+ complex with the iron storage protein ferritin.  相似文献   

12.
1. The electrophoretically fast (F) and slow (S) fragments obtained by tryptic cleavage of bovine iron-saturated transferrin differed in carbohydrate content and peptide 'maps'. 2. A fragment capable of binding one Fe3+ ion per molecule was isolated after brief tryptic digestion of bovine apotransferrin and shown closely to resemble the S fragment obtained from the iron-saturated protein. 3. Fragments F and S are probably derived from the N- and C-terminal halves of the transferrin molecule respectively. 4. Bovine transferrin could donate iron to rabbit reticulocytes, but the monoferric fragments possessed little iron-donating ability.  相似文献   

13.
Thermodynamics of anion binding to human serum transferrin   总被引:1,自引:0,他引:1  
W R Harris 《Biochemistry》1985,24(25):7412-7418
The binding of phosphate, bicarbonate, sulfate, and vanadate to human serum transferrin has been evaluated by two difference ultraviolet spectroscopic techniques. Direct titration of apotransferrin with bicarbonate, phosphate, and sulfate produces a strong negative absorbance near 245 nm, while titration with vanadate produces a positive absorbance in this region. Least-squares refinement of the absorbance data indicates that two anions of sulfate, phosphate, and vanadate bind to each transferrin molecule but that there is detectable binding of only a single bicarbonate anion. A second method used to study the thermodynamics of anion binding was competition equilibrium between anions for binding to the transferrin. The equilibrium constant for binding of the first equivalent of vanadate was determined by competition vs. phosphate and sulfate, while the equilibrium constant for binding of the second equivalent of bicarbonate was determined by competition vs. vanadate. Anion binding was described by two equilibrium constants for the successive binding of two anions per transferrin molecule: K1 = [A-Tr]/[A][Tr] and K2 = [A-Tr-A]/[A][A-Tr] where [A] represents the free anion concentration, [Tr] represents apotransferrin concentration, and [A-Tr] and [A-Tr-A] represent the concentrations of 1:1 and 2:1 anion-transferrin complexes, respectively. The results were the following: for phosphate, log K1 = 4.19 +/- 0.03 and log K2 = 3.25 +/- 0.21; for sulfate, log K1 = 3.62 +/- 0.07 and log K2 = 2.79 +/- 0.20; for vanadate, log K1 = 7.45 +/- 0.10 and log K2 = 6.6 +/- 0.30; for bicarbonate, log K1 = 2.66 +/- 0.07 and log K2 = 1.8 +/- 0.3.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The ability of transferrin to potentiate oxygen free radical-mediated endothelial cell injury was assessed. 51Cr-labeled endothelial cells derived from rat pulmonary arteries (RPAECs) were incubated with hydrogen peroxide (H2O2) in the presence and absence of holosaturated human transferrin, and the effect of transferrin on H2O2-mediated endothelial cell toxicity was determined. Addition of holosaturated transferrin potentiated H2O2-mediated RPAEC cytotoxicity at concentrations of H2O2 greater than 10 microM, suggesting that transferrin may provide a source of iron for free radical-mediated endothelial cell injury. Free radical-mediated injury is dependent on non-protein-bound iron. The ability of RPAECs to facilitate the release of iron from transferrin was assessed. We determined that RPAECs facilitate the release of transferrin-derived iron by reduction of transferrin-bound ferric iron (Fe3+) to ferrous iron (Fe2+). The reduction and release of transferrin-derived Fe2+ were inhibited by apotransferrin and chloroquine, indicating a dependence on receptor-specific binding of transferrin to the RPAEC cell surface, with subsequent endocytosis, acidification, and reduction of transferrin-bound Fe3+ to Fe2+. The release of transferrin-derived Fe2+ was potentiated by diethyldithiocarbamate, an inhibitor of intracellular superoxide dismutase (SOD). In contrast, exogenous SOD did not alter iron release, suggesting that intracellular superoxide anion (O2-) may play an important role in mediating the reduction and release of transferrin-derived iron. Results of this study suggest that transferrin may provide a source of iron for oxygen free radical-mediated endothelial cell injury and identify a novel mechanism by which endothelial cells may mediate the reduction and release of transferrin-derived iron.  相似文献   

15.
The single polypeptide chain of conalbumin strongly binds two Fe(III) or two Cu(II) ions to yield intense absorption in the visible region similar to that shown by the related protein transferrin. Comparison of the metal-ion-binding sites in the two proteins is made by exploiting the sensitivity to ligand geometry of circular dichroism (CD). For the Fe(III) proteins strong similarities of the CD spectra outweigh marginal differences. For Cu(II) conalbumin an additional negative extremum near 506 nm appears between two positive ones at 634 and 410 nm suggesting greater subtraction of oppositely signed CD components leading to lesser magnitudes for the two positive peaks than are found in Cu(II)-transferrin. The two Fe(III)-binding sites within conalbumin are compared by noting the strong similarities of the CD and MCD of proteins with Fe(III) in one site and Ga(III) in the other site, and vice versa, with the protein containing Fe(III) in both sites. Due to features of the amino acid sequences of the single protein chains, the four strong metal ion binding sites in conalbumin and transferrin cannot be identical in all particulars, yet CD spectra of their metal ion complexes are closely similar. From a study of model phenolate complexes and the wavelength maxima of visible absorption in the Fe(III), Cu(II), and Co(III) proteins near 465, 440, and 405 nm, respectively, these strong absorption bands are identified as ligand to metal ion electron-transfer transitions. It is suggested that tyrosyl residues are the donors in the electron transfer transitions and that they lock in the metal ions after being keyed into position by binding of bicarbonate or other anions.  相似文献   

16.
Equilibrium constants for the binding of anions to apotransferrin, to the recombinant N-lobe half transferrin molecule (Tf/2N), and to a series of mutants of Tf/2N have been determined by difference UV titrations of samples in 0.1 M Hepes buffer at pH 7.4 and 25 degrees C. The anions included in this study are phosphate, sulfate, bicarbonate, pyrophosphate, methylenediphosphonic acid, and ethylenediphosphonic acid. There are no significant differences between anion binding to Tf/2N and anion binding to the N-lobe of apotransferrin. The binding of simple anions like phosphate appears to be essentially equivalent for the two apotransferrin binding sites. The binding of pyrophosphate and the diphosphonates is inequivalent, and the studies on the recombinant Tf/2N show that the stronger binding is associated with the N-terminal site. Anion binding constants for phosphate, pyrophosphate, and the diphosphonates with the N-lobe mutants K206A, K296A, and R124A have been determined. Anion binding tends to be weakest for the K296A mutant, but the variation in log K values among the three mutants is surprisingly small. It appears that the side chains of K206, K296, and R124 all make comparable contributions to anion binding. There are significant variations in the intensities of the peaks in the difference UV spectra that are generated by the titrations of the mutant apoproteins with these anions. These differences appear to be related more to variations in the molar extinction coefficients of the anion-protein complexes rather than to differences in binding constants.  相似文献   

17.
Structure-function relationships for transferrins are discussed in the light of recent X-ray crystal structure determinations. A common folding pattern into two lobes, each comprising two domains is adopted; this allows the tight, but reversible binding of iron. Uptake and release of iron involve substantial domain movements which open and close the binding clefts. The iron binding sites are similar and the key role of the CO3(2-) anion bound with each Fe3+ can now be understood; structural differences near the iron binding sites suggest reasons for the different binding properties of serum transferrin and lactoferrin. The glycan moieties do not appear to affect the protein structure or metal binding properties; they are not clearly seen in the X-ray analyses but have been modelled. The accommodation of different metals and anions is illustrated by the crystal structures of Cu2+ and oxalate-substituted lactoferrins; Al3+ binding is of particular interest. New results on transferrin-receptor interactions with transferrin, and melanotransferrin and an invertebrate transferrin (both of which have defective C-terminal binding sites), emphasize possible functional differences between the two lobes. The availability of site-specific mutants of both transferrin and lactoferrin now offers the opportunity to probe the structural determinants of iron binding, iron release, and receptor binding.  相似文献   

18.
铽(Ⅲ)与人血清脱铁转铁蛋白结合的荧光光谱研究   总被引:5,自引:0,他引:5  
在pH7.40.1mol/LHepes及室温条件下,使用荧光光谱进行了Tb3+对人血清脱铁转铁蛋白的滴定.结果表明Tb3+与人血清脱铁转铁蛋白结合后,其549nm处的荧光强度增强约105倍.在549nm处Tb3+-脱铁转铁蛋白络合物的摩尔荧光强度是(9.65±0.05)×104mol-1L,Tb3+可占据脱铁转铁蛋白的两个金属离子结合部位,优先占据脱铁转铁蛋白的C端结合部位,条件平衡常数是lgKC=9.96±0.20,lgKN=6.37±0.16.Tb3+与R3+E(RE=Nd、Sm、Eu和Gd)间的线性自由能关系表明稀土离子占据脱铁转铁蛋白的C端结合部位时受离子大小的影响  相似文献   

19.
The 34-kDa periplasmic iron-transport protein (FBP) from Neisseria gonorrhoeae (nFBP) contains Fe(III) and (hydrogen)phosphate (synergistic anion). It has a characteristic ligand-to-metal charge-transfer absorption band at 481 nm. Phosphate can be displaced by (bi)carbonate to give Fe.CO(3).nFBP (lambda(max) 459 nm). The local structures of native Fe-PO(4)-nFBP and Fe.CO(3).nFBP were determined by EXAFS at the FeK edge using full multiple scattering analysis. The EXAFS analysis reveals that both phosphate and carbonate ligands bind to FBP in monodentate mode in contrast to transferrins, which bind carbonate in bidentate mode. The EXAFS analysis also suggests an alternative to the crystallographically determined position of the Glu ligand, and this in turn suggests that an H-bonding network may help to stabilize monodentate binding of the synergistic anion. The anions oxalate, pyrophosphate, and nitrilotriacetate also appear to serve as synergistic anions but not sulfate or perchlorate. The oxidation of Fe(II) in the presence of nFBP led to a weak Fe(III).nFBP complex (lambda(max) 471 nm). Iron and phosphate can be removed from FBP at low pH (pH 4.5) in the presence of a large excess of citrate. Apo-FBP is less soluble and less stable than Fe.nFBP and binds relatively weakly to Ga(III) and Bi(III) but not to Co(III) ions, all of which bind strongly to apo-human serum transferrin.  相似文献   

20.
P B O'Hara  S H Koenig 《Biochemistry》1986,25(6):1445-1450
A human serum transferrin complex was prepared in which Gd(III) was substituted for Fe(III) at the two metal-binding sites. Characteristic changes upon metal binding in both the UV absorption of ligated tyrosines and the solvent proton longitudinal magnetic relaxation rates demonstrated 2/1 metal stoichiometry and pH-dependent binding constants. Binding studies were complicated both by binding of Gd(III) to nonspecific sites on transferrin at pH less than or equal to 7 and by complexation of the Gd(III) by the requisite bicarbonate anion at pH greater than or equal to 6.0. A unique Gd(III) electron spin resonance spectrum, with a prominent signal at g = 4.96, was observed for the specific Gd(III)-transferrin complex. The major features of this spectrum were fit successfully by a model Hamiltonian which utilized crystal field parameters similar to those determined for Fe(III) in transferrin [Aasa, R. (1970) J. Chem. Phys. 52, 3919-3924]. The magnetic field dependence of the solvent proton relaxation rate was measured as a function of both pH and metal ion concentration. An observed biphasic dependence of the relaxation rate on metal concentration is attributed to either sequential metal binding to the two iron-binding sites with different relaxation properties or random binding to two sites that are similar but show conformationally induced changes in relaxation properties as the second metal is bound. The increase in the solvent proton relaxation rate with pH is consistent with a model in which a proton of a second coordination sphere water molecule is hydrogen bonded to a metal ligand which becomes deprotonated at pH 8.5.  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