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 375 毫秒
1.
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   

2.
When thermal denaturation of conalbumin solutions partially saturated with Fe(III) is observed by differential scanning calorimetry, four endotherms are observed between 40 and 100 degrees. The relative size of these four endotherms is determined by the Fe(III) to conalbumin ration. At a heating rate of 10 degrees/min, in Tris buffer at pH 7.5, observed endotherm temperature maxima and enthalpies of denaturation are: conalbumin, 63 degrees, 320 kcal/mol; intermediate I, 68 degrees, intermediate I, 77 degrees; Fe2-conalbumin, 84 degrees, 630 kcal/mol. These four endotherms are observed over a range of protein concentration from 7 to 100 mg/ml and are unchanged when excess bicarbonate is present. Stoichiometric calculations of both total protein and total iron indicate that each intermediate endotherm results from denaturation of conalbumin molecules containing only one ferric ion. These experimental results are thus consistent with the presence of two different monomeric one-iron conalbumin intermediates. They strongly suggest that the two iron binding sites of conalbumin are not equivalent.  相似文献   

3.
Chicken ovotransferrin, an iron binding protein, has two metal binding sites (amino (N) and carboxy (C) terminal sites). It binds Cu(II), Al(III), Co(II), and other metals, as well as Fe(III). In this study, the selectivity and cooperativity of the N and C sites on Al(III), Co(II), and Tb(III) binding were investigated. Metals were classified into two groups according to their site preference. Co(II) and Al(III) bound to the N site more preferably than to the C site, whereas Tb(III) bound to the C site more preferably. On Fe(III) binding, the binding constant of Fe(III) becomes larger when the other site is already occupied. Thus, positive cooperativity is seen. In the present study, the binding cooperativities of Co(II), Tb(III), and Al(III) as to the N and C sites were investigated. On Co(II) and Tb(III) binding, no cooperativity was observed, as in the case of Cu(II) [Yamamura, T. et al. (1985) in Proteins of Iron Storage and Transport (Spik, G., Montreuil, J., Crichton, R.R., & Mazurier, J., eds.) pp. 53-56, Elsevier Science Publ. B.V., Amsterdam]. In contrast, negative cooperativity was observed on Al(III) binding. Based on a model proposed by Yamamura et al. [Yamamura, T. et al. (1985) ibid.], the ratio of the binding constants, KC/KN, and the stacking coefficient, Kst, were estimated. KC/KN is 2.2 +/- 0.4 for the Tb(III) ion, 0.5 +/- 0.1 for the Co(II) ion, and 0.12 +/- 0.02 for the Al(III) ion. Kst (= 1 in a non-cooperative case) is 0.98 +/- 0.02 for the Tb(III) ion, 1.03 +/- 0.02 for the Co(II) ion, and 0.55 +/- 0.22 for the Al(III) ion.  相似文献   

4.
Metal binding to serum albumins is examined by oxidative protein-cleavage chemistry, and relative affinities of multiple metal ions to particular sites on these proteins were identified using a fast and reliable chemical footprinting approach. Fe(ii) and Cu(ii), for example, mediate protein cleavage at their respective binding sites on serum albumins, in the presence of hydrogen peroxide and ascorbate. This metal-mediated protein-cleavge reaction is used to evaluate the binding of metal ions, Na(+), Mg(2+), Ca(2+), Al(3+), Cr(3+), Mn(2+), Co(2+), Ni(2+), Zn(2+), Cd(2+), Hg(2+), Pb(2+), and Ce(3+) to albumins, and the relative affinities (selectivities) of the metal ions are rapidly evaluated by examining the extent of inhibition of protein cleavage. Four distinct systems Fe(II)/BSA, Cu(II)/BSA, Fe(II)/HSA and Cu(II)/HSA are examined using the above strategy. This metallomics approach is novel, even though the cleavage of serum albumins by Fe(II)/Cu(II) has been reported previously by this laboratory and many others. The protein cleavage products were analyzed by SDS PAGE, and the intensities of the product bands quantified to evaluate the extent of inhibition of the cleavage and thereby evaluate the relative binding affinities of specific metal ions to particular sites on albumins. The data show that Co(II) and Cr(III) showed the highest degree of inhibition, across the table, followed by Mn(II) and Ce(III). Alakali metal ions and alkaline earth metal ions showed very poor affinity for these metal sites on albumins. Thus, metal binding profiles for particular sites on proteins can be obtained quickly and accurately, using the metallomics approach.  相似文献   

5.
A phosphorescent trivalent cation, terbium [Tb(III)], has been used to study the binding of different polyvalent cations to the proteins of wheat (Triticum aestivum L.) root plasma membranes. The phosphorescence emission intensity of Tb(III) was enhanced after Tb(III) binding to wheat root plasma membranes as a result of nonradiative resonance energy transfer from the membrane protein tyrosine and phenylalanine residues. Complex, saturable Tb(III) binding was observed, suggesting multiple binding sites. Bound Tb(III) could be displaced by divalent cations in the general order: Mn(II) > Ca(II) > Mg(II). Al(III) was very effective in reducing the protein-enhanced Tb(III) phosphorescence at pH values below 5. Al(III) also altered the Tb(III) phosphorescence lifetime, suggesting Al(III)-induced changes in membrane protein conformation. The more Al(III)-sensitive wheat cultivar (Anza) bound Al(III) with higher affinity than the more tolerant cultivar (BH 1146). At pH 5.5 where Al(III) did not displace bound Tb(III), low levels of Al(III) reduced the ability of Mn(II) to decrease Tb(III) phosphorescence. The significance of these results is discussed with respect to the mechanisms of Al(III) tolerance in wheat and the potential beneficial effects of Al(III) in reducing Mn(II) phytotoxicity.  相似文献   

6.
The reduction of low-molecular-weight Cu(II) and Fe(III) complexes by soybean leghemoglobin alpha was characterized using both kinetic analysis and 1H-NMR experiments. Whereas Fe(III) (CN)6(3-) was reduced through an outer sphere transfer over the exposed heme edge, all other Cu(II) and Fe(III) complexes investigated were reduced via a site-specific binding of the metal to the protein. Reduction of all metal complexes was enhanced by decreasing pH while only Fe(III)NTA reduction kinetics were altered by changes in ionic strength. Rates of reduction for both Cu(II) and Fe(III) were also affected inversely by the effective binding constant of the metal chelate used. NMR data confirmed that both Cu(II)NTA and Fe(III)NTA were bound to specific sites on the protein. Cu(II) bound preferentially to distal His-61 and Fe(III) exerted its greatest effect on two surface lysine residues with epsilon proton resonances at 3.04 and 3.12 ppm. The Fe(III)NTA complex also had a mild but noticeable line broadening effect on the distal His-61 singlet resonance near 5.3 ppm. Like hemoglobin and myoglobin, leghemoglobin might function not only as an oxygen carrier, but also as a biological reductant for low-molecular-weight Cu(II) and Fe(III) complexes.  相似文献   

7.
Complexes of the natural siderophore, desferricoprogen (DFC), with several trivalent and divalent metal ions in aqueous solution were studied by pH-potentiometry, UV-Vis spectrophotometry and cyclic voltammetry. DFC was found to be an effective metal binding ligand, which, in addition to Fe(III), forms complexes of high stability with Ga(III), Al(III), In(III), Cu(II), Ni(II) and Zn(II). Fe(II), however, is oxidized by DFC under anaerobic conditions and Fe(III) complexes are formed. By comparing the results with those of desferrioxamine B (DFB), it can be concluded that the conjugated beta-double bond slightly increases the stability of the hydroxamate chelates, consequently increases the stability of mono-chelated complexes of DFC. Any steric effect by the connecting chains arises only in the bis- and tris-chelated complexes. With metal ions possessing a relatively big ionic radius (Cu(II), Ni(II), Zn(II), In(III)) DFC, containing a bit longer chains than DFB, forms slightly more stable complexes. With smaller metal ions the trend is the opposite. Also a notable difference is that stable trinuclear complex, [Cu(3)L(2)], is formed with DFC but not with DFB. Possible bio-relevance of the Fe(II)/Fe(III) results is also discussed in the paper.  相似文献   

8.
Immobilized metal ion affinity chromatography has been used to demonstrate and partially characterize Fe(III) binding sites on apoferritin. Binding of Fe(III) to these sites is influenced by pH, but not affected by high ionic strength. These results suggest that both ionic and coordinate covalent interactions are important in the formation of the Fe(III): apoferritin complex. This is, to our knowledge, the first demonstration of direct Fe(III) binding to apoferritin. Other immobilized metal ions, including Zn(II), Ni(II), Cu(II), Cr(III), Co(II), and Tb(III), displayed little or no adsorption of apoferritin. The analytical technique of immobilized metal ion affinity chromatography also shows great promise in the purification of apoferritin, ferritin, and other iron-binding proteins.  相似文献   

9.
10.
Electron paramagnetic resonance (epr) and ultraviolet difference spectroscopy of vanadyl conalbumin indicate a binding capacity of two vanadyl ions, VO2+, per protein molecule in the pH 8–11 range; the binding capacity drops in the pH 6–8 range with an apparent pKa′ = 6.6. Iron-saturated conalbumin does not bind vanadyl ions, which suggests common binding sites for iron and vanadium. Ultraviolet difference spectroscopy indicates 2–3 tyrosines are involved in the binding of each metal ion; pH titrations show that three protons are released per vanadyl ion bound by conalbumin. Room and liquid nitrogen temperature X-band (ca. 9.2–9.5 gHz) epr spectra show that the vanadyl ion binds in three magnetically distinct environments (A, B, and C) that arise from interconvertible metal site configurations. These configurations are probably examples of conformational substrates of the protein. Q-band (ca 34 gHz) epr spectra resolve the spectral features more clearly and show that two configurations (A and B) have axially symmetric epr parameters but angles of noncoincidence of 12° and 8°, respectively, between the z components of the g and nuclear hyperfine tensors. The third (C) configuration has rhombic magnetic symmetry and a 6° angle of noncoincidence. These observations demonstrate that the metal sites are of low symmetry and are flexible in their geometry about the metal.The isotropic g and nuclear hyperfine tensor values and the line widths used in computer-simulated epr spectra are consistent with four oxygen or three oxygen and one nitrogen donor atoms binding equatorially to the VO2+ group. The apparent stability constant indicates that vanadyl ion binds to conalbumin approximately twelve orders of magnitude more weakly than iron to human serotransferrin but still sufficiently strongly to overcome hydrolysis.  相似文献   

11.
Al(III) toxicity in living organisms is based on competition with other metal cations. Mg(II) is one of the most affected cations, since the size similarity dominates over the charge identity. The slow ligand exchange rates for Al(III) render the ion useless as a metal ion at the active sites of enzymes and provide a mechanism by which Al(III) inhibits Mg(II) dependent biochemical processes. Al(III) cation interactions with relevant bioligands have been studied in a protein-model environment in gas and aqueous phases using density functional theory methods. The protein model consists of the metal cation bound to two chosen bioligands (functional groups of the amino acid side chains, one of them being always an acetate) and water molecules interacting with the cation to complete its first coordination shell. Analogous Mg(II) complexes are calculated and compared with the Al(III) ones. Formation energies of the complexes are calculated in both phases and magnesium/aluminum exchange energies evaluated. The effect of different dielectric media is also analyzed. The presence of an acetate ligand in the binding site is found to promote both, complex formation and metal exchange reactions. In addition, buried binding sites (with low dielectric constant) of the protein favor metal exchange, whereas fully solvated environments of high dielectric constant require the presence of two anionic ligands for metal exchange to occur.  相似文献   

12.
Although metal ions can promote amyloid formation from many proteins, their effects on the formation of amyloid from transthyretin have not been previously studied. We therefore screened the effects of Cu(II), Zn(II), Al(III), and Fe(III) on amyloid formation from wild-type (WT) transthyretin as well as its V30M, L55P, and T119M mutants. Cu(II) and Zn(II) promoted amyloid formation from the L55P mutant of transthyretin at pH 6.5 but had little effect on amyloid formation from the other forms of the protein. Zn(II) promoted L55P amyloid formation at pH 7.4 but Cu(II) inhibited it. Cu(II) gave dose-dependent quenching of the tryptophan fluorescence of transthyretin and the fluorescence of 1-anilino-8-naphthalene sulfonate bound to it. Zn(II) gave dose-dependent quenching of the tryptophan but not the 1-anilino-8-naphthalene sulfonate fluorescence. Apparent dissociation constants for Cu(II) and Zn(II) binding at pH 7.4 of approximately 10 nM and approximately 1 microM (approximately 0.4 microM and approximately 5 microM at pH 6.5), respectively, were obtained from the quenching data. Zn(II) enhanced urea-mediated the dissociation of the L55P but not the WT transthyretin tetramer. Cu(II), depending on its concentration, either had no effect or stabilized the WT tetramer but could enhance urea-mediated dissociation of L55P.  相似文献   

13.
The kinetics and mechanism of binding of Cu-(II).bleomycin, Fe(III).bleomycin, and Cu(II).phleomycin to DNA were studied by using fluorometry, equilibrium dialysis, electric dichroism, and temperature-jump and stopped-flow spectrophotometry. The affinity of Cu(II).bleomycin for DNA was greater than that of metal-free bleomycin but less than that of Fe(III).bleomycin. Cu(II).bleomycin exhibited a two-step binding process, with the slow step indicating a lifetime of 0.1 s for the Cu(II).bleomycin.DNA complex. Fe(III).bleomycin binding kinetics indicated the presence of complexes having lifetimes of up to 22 s. DNA was lengthened by 4.6 A/molecule of bound Cu(II).bleomycin and by 3.2 A/bound Fe(III).bleomycin but not at all by Cu(II).phleomycin, suggesting that both bleomycin complexes intercalate while the phleomycin complex does not. However, phleomycin exhibited nearly the same specificity of DNA base release as bleomycin. These results suggest that the coordinated metal ion plays a major role in the binding of metal-bleomycin complexes to DNA but that intercalation is neither essential for DNA binding and degradation nor primarily responsible for the specificity of DNA base release by these drugs.  相似文献   

14.
Ghatak  H.  Mukhopadhyay  S.K.  Jana  T.K.  Sen  B.K.  Sen  S. 《Wetlands Ecology and Management》2004,12(3):145-155
Humic (HA) and fulvic (FA) acids isolated from mangrove sediments of Sundarban, the largest delta on earth in the estuarine phase of the river Ganges, were studied and attempts were made to characterize their binding sites by quenching of Synchronous fluorescence (SyF) bands with Fe (III) and Cu (II). A modified Stern-Volmer relationship applicable for static quenching was applied for the determination of conditional stability constants and the data were compared with those determined by potentiometric titration. In the excited state HA and FA showed different acidity constant compared to the ground state. Values of the conditional stability constant (log Kc) for Fe (III) and Cu (II) indicated that binding sites were bidentate in nature. FA were better chelators than the HA fractions. High energy binding sites of both FA & HA were occupied by Fe(III) and the low energy binding sites, mainly responsible for mobilization and immobilization of metal, were occupied by Cu(II).  相似文献   

15.
Apoconalbumin binds Mn(II) at two sites with association constants of K1 = 7 (+/- 1) X 10(4) and K2 = 0.4 (+/- 0.25) X 10(4) M-1. The binding is tighter in the presence of excess bicarbonate resulting in K1 = 1.8 (+/- 0.2) X 10(5) and K2 = 3 (+/- 2) X 10(4) M-1. The electron paramagnetic resonance spectrum (at both 9 and 35 GHz) of Mn(II) bound at the tight site reveals a rhombic distortion (lambda = E/D approximately equal to 0.25-0.31) in the protein ligand environment of the mental ion. An evaluation of the 1/pT1p, paramagnetic contribution to the longitudinal relaxation rate of solvent protons with Mn(II)-, Mn(III)-, and Fe(III)-derivatives of conalbumin revealed that the mental ion in each site of conalbumin is accessible to one water molecule. For Mn(II)-conalbumin and Mn(III)-conalbumin species, inner coordination sphere protons are rapidly exchanging with the bulk solvent, while slow exchange conditions prevail for Fe(III)-conalbumin.  相似文献   

16.
E E Snyder  B W Buoscio  J J Falke 《Biochemistry》1990,29(16):3937-3943
The molecular mechanisms by which protein Ca(II) sites selectively bind Ca(II) even in the presence of high concentrations of other metals, particularly Na(I), K(I), and Mg(II), have not been fully described. The single Ca(II) site of the Escherichia coli receptor for D-galactose and D-glucose (GGR) is structurally related to the eukaryotic EF-hand Ca(II) sites and is ideally suited as a model for understanding the structural and electrostatic basis of Ca(II) specificity. Metal binding to the bacterial site was monitored by a Tb(III) phosphorescence assay: Ca(II) in the site was replaced with Tb(III), which was then selectively excited by energy transfer from protein tryptophans. Photons emitted from the bound Tb(III) enabled specific detection of this substrate; for other metals binding was detected by competitive displacement of Tb(III). Representative spherical metal ions from groups IA, IIA, and IIIA and the lanthanides were chosen to study the effects of metal ion size and charge on the affinity of metal binding. A dissociation constant was measured for each metal, yielding a range of KD's spanning over 6 orders of magnitude. Monovalent metal ions of group IA exhibited very low affinities. Divalent group IIA metal ions exhibited affinities related to their size, with optimal binding at an effective ionic radius between those of Mg(II) (0.81 A) and Ca(II) (1.06 A). Trivalent metal ions of group IIIA and the lanthanides also exhibited size-dependent affinities, with an optimal effective ionic radius between those of Sc(III) (0.81 A) and Yb(III) (0.925 A). The results indicate that the GGR site selects metal ions on the basis of both charge and size.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
W Y Lin  C D Eads  J J Villafranca 《Biochemistry》1991,30(14):3421-3426
TNS, 2-p-toluidinylnaphthalene-6-sulfonate, has been used as a fluorescent probe to determine the binding constants of metal ions to the two binding sites of Escherichia coli glutamine synthetase (GS). TNS fluorescence is enhanced dramatically when bound to proteins due to its high quantum yield resulting from its interactions with hydrophobic regions in proteins. The fluorescence energy transfer from a hydrophobic tryptophan residue of GS to TNS has been detected as an excitation band centered at 280 nm. Therefore, TNS is believed to be bound to a hydrophobic site on the GS surface other than the active site and is located near a hydrophobic Trp residue of GS. GS binds lanthanide ions [Ln(III)] more tightly than either Mn(II) or Mg(II), and the binding constants of several lanthanide ions were determined to be in the range (2.1-4.6) x 10(10) and (1.4-3.0) x 10(8) M-1 to the two metal binding sites of GS, respectively. The intermetal distances between the two metal binding sites of GS were also determined by measuring the efficiencies of energy transfer from Tb(III) to other Ln(III) ions. The intermetal distances of Tb(III)-Ho(III) and Tb(III)-Nd(III) were 7.9 and 6.8 A, respectively.  相似文献   

18.
19.
1. Human lactoferrin and transferrin are capable of binding several transition metal ions [Fe(III), Cu(II), Mn(III), Co(III)] into specific binding sites in the presence of bicarbonate. 2. Increased conformational stability and increased resistance to protein unfolding is observed for these metal-ion complexes compared to the apoprotein form of these proteins. 3. Mn(III)-lactoferrin and transferrin complexes exhibit steeper denaturation transitions than the Co(III) complexes of these proteins suggesting greater cooperativity in the unfolding process. 4. The incorporation of Fe(III) into the specific metal binding sites offers the greatest resistance to thermal unfolding when compared to the other transition metal ions studied. 5. Non-coincidence of unfolding transitions is observed, with fluorescence transition midpoints being lower than those determined by absorbance measurements. 6. Fully denatured proteins in the presence of urea and alkyl ureas exhibit fluorescence wavelength maxima at 355-356 nm indicative of tryptophan exposure upon protein unfolding.  相似文献   

20.
P J Breen  E K Hild  W D Horrocks 《Biochemistry》1985,24(19):4991-4997
The binding of Ca(II) and members of the trivalent lanthanide ion, Ln(III), series to apoparvalbumin (isotype pI = 4.75) from codfish (Gadus callarius L) results in the development of a distinctive sharp feature in the UV absorption spectrum at about 290 nm. Titration curves obtained by monitoring the spectral change in this region reveal a change in slope after the addition of 1 equiv of metal ion and no further rise after 2 equiv has been added, consistent with sequential binding to the principal EF and CD sites. Laser-induced luminescence excitation spectra of the 7F0----5D0 transition of bound Eu(III) demonstrate the quantitative binding of this ion to the principal sites and disclose the presence of a subsidiary site at pH values greater than 6. Metal ion competition experiments monitored by means of this excitation transition show that the early members of the Ln(III) ion series bind more tightly than those at the end. Tryptophan-sensitized Tb(III) luminescence reveals that this ion binds sequentially to the EF and CD sites, in that order. The intrinsic tryptophan fluorescence of apoparvalbumin is increased in a stepwise fashion as Ca(II) or Ln(III) ions bind sequentially, with the exceptions of Eu(III) and Yb(III). The binding of the latter two ions causes quenching of the protein fluorescence via an energy-transfer process which involves low-lying charge-transfer bands. The distance dependences of the tryptophan to Tb(III) and tryptophan to Eu(III) energy-transfer processes are observed to be identical, consistent with a F?rster-type mechanism in both cases.  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