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 15 毫秒
1.
Ca(II) ions are crucial during proteolytic conversion of Factor XIII zymogen into the active enzyme Factor XIIIa. Factor XIII proteolyzed by thrombin or trypsin in the presence of 5 mM-EDTA resulted in rapid inactivation of transglutaminase activity. Factor XIIIa formed by thrombin or trypsin in the presence of 40 microM-Tb(III) ions, however, was indistinguishable from Factor XIIIa formed in the presence of 2-5 mM-Ca(II) ions with respect to molecular mass and transglutaminase activity. Thrombin treatment of Factor XIII in the presence of 1-5 microM-Tb(III) ions resulted in three fragments (76 kDa, 51 kDa and 19 kDa) with simultaneous loss of transglutaminase activity. Tb(III) ions at concentrations greater than 40 microM made platelet Factor XIII resistant to proteolysis by either thrombin or trypsin. Other lanthanide(III) ions [Ln(III) ions] tested [Ce(III), La(III) and Gd(III) ions] functioned similarly to Tb(III) ions during proteolytic activation of Factor XIII. Ln(III) ions (10-100 microM) were unable to replace the Ca(II) ions required for transglutaminase activity of Factor XIIIa. Tb(III) ions also inhibited in a non-competitive manner the transglutaminase activity of Factor XIIIa (Ki 71 microM) even when measured in the presence of 200-fold molar excess of Ca(II) ions. Factor XIII selectively bound to a Tb(III)-chelate affinity column, and could not be eluted by 100 mM-CaCl2. Binding of Tb(III) ions to Factor XIII was demonstrated by fluorescence emission due to Forster energy transfer. A 10(4)-fold molar excess of CaCl2, but not NaCl, partially quenched Tb(III) fluorescence. Low concentrations (5-20 microM) of Tb(III) ions also inhibited the binding of Factor XIII to des-A-fibrinogen by about 43%, whereas higher concentrations (40-100 microM) promoted binding. Conformational changes in Factor XIII consequent to the binding of Tb(III) ions could be responsible for the observed effects on protein structure and function.  相似文献   

2.
The rates of dissociation of 2 equiv of various metal ions [Ca(II), Cd(II), Pr(III), Nd(III), Sm(III), Eu(III), Gd(III), Tb(III), Dy(III), Ho(III), Er(III), Yb(III), and Lu(III)] from the primary CD and EF metal ion binding sites of parvalbumin (isotype pI = 4.75) from codfish (Gadus callarius L) were measured by stopped-flow techniques. The removal or replacement of metal ions was monitored by changes in sensitized Tb(III) luminescence or in intrinsic protein tryptophan fluorescence as quenching ions [Eu(III) or Yb(III)] were bound or removed or as the apoprotein was formed. In experiments wherein the bound metal ions were removed by mixing the parvalbumin with an excess of 1,2-diaminocyclohexanetetraacetic acid (DCTA), the kinetic traces were best fit by a double exponential with koff rate constants of 1.07 and 5.91 s-1 for Ca(II), 1.54 and 10.5 s-1 for Cd(II), and approximately 0.05 and approximately 0.5 s-1 for all of the trivalent lanthanide ions. In experiments wherein the bound metal ions were exchanged with an excess of a different metal ion, pseudo-first-order rate constants were proportional to the concentration of excess attacking metal ion for both the fast and slow processes in most experiments. In these cases, extrapolation of the rate constants to zero concentration of attacking metal ion gave values which agree well with the DCTA scavenging results. This finding demonstrates that the off rate constants do not depend on the occupancy of the neighboring site and therefore implies that there is no significant cooperativity in metal ion binding between the two sites in parvalbumin.  相似文献   

3.
Pulsed dye laser excitation spectroscopy of the 7F0----5D0 transition of Eu(III) reveals only a single peak as this ion is titrated into apocalmodulin. A titration based on the intensity of this transition shows that the first two Eu(III) ions bind quantitatively to two tight sites, followed by weaker binding (Kd = 2 microM) to two additional sites under conditions of high ionic strength (0.5 M KC1). This excitation experiment is also shown to be a general method for measuring contaminating levels of EDTA down to 0.2 microM in proton solutions. Experiments with Tb(III) using both direct laser excitation and indirect sensitization of Tb(III) luminescence through tyrosine residues in calmodulin also give evidence for two tight and two weaker binding sites (Kd = 2-3 microM). The indirect sensitization results primarily upon binding to the two weaker sites, implying that Tb(III) binds first to domains I and II, which are remote from tyrosine-containing domains III and IV. The 7F0----5D0 excitation signal of Eu(III) was used to measure the relative overall affinities of the tripositive lanthanide ions, Ln(III), across the series. Ln(III) ions at the end of the series are found to bind more weakly than those at the beginning and middle of the series. Eu(III) excited-state lifetime measurements in H2O and D2O reveal that two water molecules are coordinated to the Eu(III) at each of the four metal ion binding sites. Measurements of F?rster-type nonradiative energy-transfer efficiencies between Eu(III) and Nd(III) in the two tight sites were carried out by monitoring the excited-state lifetimes of Eu(III) in the presence and absence of the energy acceptor ion Nd(III).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
C L Wang  P C Leavis  J Gergely 《Biochemistry》1984,23(26):6410-6415
The stepwise addition of Tb3+ to calmodulin yields a large tyrosine-sensitized Tb3+ luminescence enhancement as the third and fourth ions bind to the protein [Wang, C.-L. A., Aquaron, R. R., Leavis, P. C., & Gergely, J. (1982) Eur. J. Biochem. 124, 7-12]. Since the only tyrosine residues in calmodulin are located within binding sites III and IV, these results suggest that Tb3+ binds first to sites I and II. Recent NMR studies have provided evidence that Ca2+, on the other hand, binds preferentially to sites III and IV. Kinetic studies using a stopped-flow apparatus also show that the preferential binding of Ca2+ and lanthanide ions is different. Upon rapid mixing of 2Ca-calmodulin with two Tb3+ ions, there was a small and rapid tyrosine fluorescence change, but no Tb3+ luminescence was observed, indicating that Tb3+ binds to sites I and II but not sites III and IV. When two Tb3+ ions are mixed with 2Dy-calmodulin, Tb3+ luminescence rises rapidly as Tb3+ binds to the empty sites III and IV, followed by a more gradual decrease (k = 0.4 s-1 as the ions redistribute themselves over the four sites. These results indicate that (i) both Tb3+ and Dy3+ prefer binding to sites I and II of calmodulin and (ii) the binding of Tb3+ to calmodulin is not impeded by the presence of two Ca2+ ions initially bound to the protein. Thus, the Ca2+ and lanthanide ions must exhibit opposite preferences for the four sites of calmodulin: sites III and IV are the high-affinity sites for Ca2+, whereas Tb3+ and Dy3+ prefer sites I and II.  相似文献   

5.
Bowen LM  Muller G  Riehl JP  Dupureur CM 《Biochemistry》2004,43(48):15286-15295
Type II restriction enzymes are homodimeric systems that bind four to eight base pair palindromic recognition sequences of DNA and catalyze metal ion-dependent phosphodiester cleavage. While Mg(II) is required for cleavage in these enzymes, in some systems Ca(II) promotes avid substrate binding and sequence discrimination. These properties make them useful model systems for understanding the roles of alkaline earth metal ions in nucleic acid processing. We have previously shown that two Ca(II) ions stimulate DNA binding by PvuII endonuclease and that the trivalent lanthanide ions Tb(III) and Eu(III) support subnanomolar DNA binding in this system. Here we capitalize on this behavior, employing a unique combination of luminescence spectroscopy and DNA binding assays to characterize Ln(III) binding behavior by this enzyme. Upon excitation of tyrosine residues, the emissions of both Tb(III) and Eu(III) are enhanced severalfold. This enhancement is reduced by the addition of a large excess of Ca(II), indicating that these ions bind in the active site. Poor enhancements and affinities in the presence of the active site variant E68A indicate that Glu68 is an important Ln(III) ligand, similar to that observed with Ca(II), Mg(II), and Mn(II). At low micromolar Eu(III) concentrations in the presence of enzyme (10-20 microM), Eu(III) excitation (7)F(0) --> (5)D(0) spectra yield one dominant peak at 579.2 nm. A second, smaller peak at 579.4 nm is apparent at high Eu(III) concentrations (150 microM). Titration data for both Tb(III) and Eu(III) fit well to a two-site model featuring a strong site (K(d) = 1-3 microM) and a much weaker site (K(d) approximately 100-200 microM). Experiments with the E68A variant indicate that the Glu68 side chain is not required for the binding of this second Ln(III) equivalent; however, the dramatic increase in DNA binding affinity around 100 microM Ln(III) for the wild-type enzyme and metal-enhanced substrate affinity for E68A are consistent with functional relevance for this weaker site. This discrimination of sites should make it possible to use lanthanide substitution and lanthanide spectroscopy to probe individual metal ion binding sites, thus adding an important tool to the study of restriction enzyme structure and function.  相似文献   

6.
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   

7.
The binding of Gd(III) to rabbit IgG (immunoglobulin G) and the Fab (N-terminal half of heavy and light chain), (Bab')2 (N-terminal half of heavy and light chains joined by inter-chain disulphide bond), Fc (C-terminal half of heavy-chain dimer)and pFc' (C-terminal quarter of heavy-chain dimer) fragments was demonstrated by measurements of the enhancement of the solvent-water proton relaxation rates in the appropriate Gd(III) solutions. At pH 5.5 there are six specific Gd(III)-binding sites on the IgG. These six sites can be divided into two classes; two very 'tight' sites on the Fc fragment (Kd approx. 5 muM) and two weaker sites on each Fab region (Kd approx. 140 muM). Ca(II) does not apparently compete for these metal-binding sites. The metal-binding parameters for IgG can be explained as the sum of the metal binding to the isolated Fab and Fc fragments, suggesting that there is no apparent interaction between the Fab and Fc regions in the IgG molecule. The binding of Gd(III) to Fab and Fc fragments was also monitored by measuring changes in the electron-spin-resonance spectrum of Gd(III) in the presence of each fragment and also by monitoring the effects of Gd(III) on the protein fluorescence at 340 nm (excitation 295 nm). The fluorescence of Tb(III) solutions of 545 nm (excitation 295 nm) is enhanced slightly on addition of Fab or Fc.  相似文献   

8.
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   

9.
We attempted to establish whether lanthanide ions, when added to sarcoplasmic reticulum (SR) membranes in the absence of nucleotide, compete with Ca2+ for binding to the transport sites of the Ca(2+)-ATPase in these membranes, or whether they bind to different sites. Equilibrium measurements of the effect of lanthanide ions on the intrinsic fluorescence of SR ATPase and on 45Ca2+ binding to it were performed either at neutral pH (pH 6.8), i.e. when endogenous or contaminating Ca2+ was sufficient to nearly saturate the ATPase transport sites, or at acid pH (pH 5.5), which greatly reduced the affinity of calcium for its sites on the ATPase. These measurements did reveal apparent competition between Ca2+ and the lanthanide ions La3+, Gd3+, Pr3+, and Tb3+, which all behaved similarly, but this competition displayed unexpected features: lanthanide ions displaced Ca2+ with a moderate affinity and in a noncooperative way, and the pH dependence of this displacement was smaller than that of the Ca2+ binding to its own sites. Simultaneously, we directly measured the amount of Tb3+ bound to the ATPase relative to the amount of Ca2+ and found that Tb3+ ions only reduced significantly the amount of Ca2+ bound after a considerable number of Tb3+ ions had bound. Furthermore, when we tested the effect of Ca2+ on the amount of Tb3+ bound to the SR membranes, we found that the Tb3+ ions which bound at low Tb3+ concentrations were not displaced when Ca2+ was added at concentrations which saturated the Ca2+ transport sites. We conclude that the sites on SR ATPase to which lanthanide ions bind with the highest affinity are not the high affinity Ca2+ binding and transport sites. At higher concentrations, lanthanide ions did not appear to be able to replace Ca2+ ions and preserve the native structure of their binding pocket, as evaluated in rapid filtration measurements from the effect of moderate concentrations of lanthanide ions on the kinetics of Ca2+ dissociation. Thus, the presence of lanthanide ions slowed down the dissociation from its binding site of the first, superficially bound 45Ca2+ ion, instead of specifically preventing the dissociation of the deeply bound 45Ca2+ ion. These results highlight the need for caution when interpreting, in terms of calcium sites, experimental data collected using lanthanide ions as spectroscopic probes on SR membrane ATPase.  相似文献   

10.
Luminescence measurements show that apoferritin binds three Tb(III) atoms per subunit in accordance with crystallographic evidence. Fe(II) competes with Tb(III) for at least some of the binding sites. This competition may be the molecular basis for the inhibition of iron incorporation into apoferritin brought about by Tb(III). Ca(II), which is generally replaced by Tb(III) in Ca(II) binding proteins, does not compete with the lanthanide for binding to apoferritin.  相似文献   

11.
N Coruh  J P Riehl 《Biochemistry》1992,31(34):7970-7976
A number of different experimental techniques have been used to probe the details of structural changes on the binding of Ca(II) to the large number of known calcium-binding proteins. The use of luminescent lanthanide(III) ions, especially terbium(III) and europium(III), as substitutional replacement for calcium(II), has led to a number of useful experiments from which important details concerning the metal ion coordination sites have been obtained. This work is concerned with the measurement of the circularly polarized luminescence (CPL) from the 5D4----7F5 transition of Tb(III) bound to the calcium binding sites of bovine trypsin, bovine brain calmodulin, and frog muscle parvalbumin. It is demonstrated that it is possible to make these polarization measurements from very dilute solutions (less than 20 microM) and monitor structural changes as equivalents of Tb(III) are added. It is shown that the two proteins that belong to the class of "EF-hand" structures (calmodulin and parvalbumin) possess quite similar CPL line shapes, whereas Tb(III) bound to trypsin has a much different band structure. CPL results following competitive and consecutive binding of Ca(II) and Tb(III) bound to calmodulin are also reported and yield information concerning known differences between the sequence of binding of these two species.  相似文献   

12.
J Bruno  W D Horrocks  R J Zauhar 《Biochemistry》1992,31(31):7016-7026
The effects of minor differences in the amino acid sequences between a vertebrate (bovine testes) and an invertebrate (octopus) calmodulin on metal ion binding were investigated via laser-induced Eu3+ and Tb3+ luminescence. Amino acid substitutions at residues which are coordinated to the metal ion do not produce any detectable changes in the 7F0----5D0 excitation spectrum of the Eu3+ ion bound to octopus calmodulin relative to bovine testes calmodulin; only minor differences in the excited-state lifetime values in D2O solution are observed. The dissociation constants for Eu3+ (1.0 +/- 0.2 microM) and Tb3+ (5 +/- 1 microM) from the weak lanthanide binding sites (III and IV, numbered from the amino terminus) of octopus calmodulin were measured using luminescence techniques. Both values agree well with those reported previously for bovine testes calmodulin [Mulqueen, P. M., Tingey, J. M., & Horrocks, W. D., Jr. (1985) Biochemistry 24, 6639-6645]. The measured dissociation constant of Eu3+ bound in the tight lanthanide binding sites (I and II) is 6 +/- 2 nM for octopus calmodulin and 12 +/- 2 nM for bovine testes calmodulin. The distances between sites I and II (12.4 +/- 0.5 A) and sites III and IV (11.7 +/- 0.8 A) were determined from F?rster-type energy transfer in D2O solutions of octopus calmodulin containing bound Eu3+ donor and Nd3+ acceptor ions. F?rster theory parameters for nonradiative energy transfer between Tyr138 and Tb3+ ions bound at sites III and IV of octopus calmodulin were comprehensively evaluated, including a dynamics simulation of the orientation factor kappa 2. This theory is found to account quantitatively for the observed energy-transfer efficiency as evaluated from the observed sensitized Tb3+ emission.  相似文献   

13.
Monoclonal antibodies to various domains of human protein C were characterized, and the cross-reactivity of these antibodies with other vitamin K-dependent proteins was explored. Three antibodies, JTC-1, -2, and -3 reacted with protein C only in the presence of Ca2+ and were shown to bind to the light chain of protein C. It is suggested that these antibodies recognize a gamma-carboxyglutamic acid domain-related conformational change induced by metal ions, evidenced by the fact that half-maximal binding was observed at calcium concentration of 0.5, 0.6, and 0.7 mM, respectively, by the fact that these antibodies, even in the presence of Ca2+, do not react with gamma-carboxyglutamic acid domainless protein C, and by the fact that Zn2+ and Tb3+ support binding in essentially the same way. Each cell line was stabilized by recloning five times. In addition each antibody had a single isoelectric point and was of the IgG1 kappa class. The interaction of antibodies JTC-1, -2; and -3 with protein C-Ca2+ was characterized by a single class of binding sites with Kd of 3.98 X 10(-9) M, 4.01 X 10(-9) M, and 6.76 X 10(-9) M, respectively. However, antibodies JTC-1, -2, and -3 bound to prothrombin-Ca2+ with Kd of 7.81 X 10(-9) M, 2.0 X 10(-7) M, and higher than 1.0 X 10(-5) M, respectively. In addition they had weak affinity for factor X in the presence of Ca2+. The results indicate that the antibodies JTC-1, -2, and -3 are conformation-specific monoclonal antibodies directed against an at least partially common metal ion-induced three-dimensional structure in protein C, prothrombin, and factor X.  相似文献   

14.
This work reports Eu(III) and Tb(III) luminescence titrations in which the lanthanide ions were used as spectroscopic probes for Ca(II) ions to determine the metal binding ability of Ac-NESVKEEGGW-NH(2) and Ac-NESVKEDGGW-NH(2). These decapeptides correspond to the putative calcium binding region of the plant antifungal proteins SI-alpha1 from Sorghum bicolor and of Zeathionin from Zea mays, respectively. The luminescence spectra for the Eu(III)-decapeptide system (red emission) with the excitation at the Trp band at 280 nm showed an enhancement of the intensities of the 5D(0)-->7F(J) transitions (where J=0-4) with increments of Eu(III) ion concentration. The photoluminescence titration data of the terbium ion (green emission) in the decapeptide solutions showed intensification of the 5D(4)-->7F(J) transitions (J=0-6), similar to that observed for the Eu(III) ion. Thus, energy transfer from Ac-NESVKEEGGW-NH(2) and Ac-NESVKEDGGW-NH(2) to the trivalent lanthanide ions revealed that these peptides are capable of binding to these metal ions with association constants of the order of 10(5) M(-1). The amino acid derivative Ac-Trp-OEt also transferred energy to Tb(III) and Eu(III) ions as judged from the quenching of tryptophan luminescence. However, the energy transfers were significantly lower. Taken together the luminescence titration data indicated that Ac-NESVKEEGGW-NH(2) and Ac-NESVKEDGGW-NH(2) bind efficiently to both trivalent lanthanide ions and that these ions may be used as probes to distinguish an anionic peptide from a neutral amino acid derivative.  相似文献   

15.
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   

16.
alpha-Lactalbumin (alpha-LA) is a calcium binding protein that also binds Mn(II), lanthanide ions, A1(III), Zn(II), Co(II). The structural implications of cation binding were studied by high-resolution proton (200 MHz) NMR and photochemically induced dynamic nuclear polarization (CIDNP) spectroscopy. Marked changes were observed in the NMR spectra of the apoprotein upon addition of a stoichiometric amount of calcium to yield Ca(II)-alpha-LA, manifested particularly in ring current shifted aliphatic peaks and in several shifts in the aromatic region, all of which were under slow exchange conditions. The CIDNP results showed that two surface-accessible tyrosine residues, assigned as Tyr-18 and -36, became inaccessible to the solvent upon addition of 1:1 Ca(II) to apo-alpha-lactalbumin, while Tyr-103 and Trp-104 remained completely accessible in both conformers. The proton NMR spectra of apo-alpha-LA and A1(III)-alpha-LA were extremely similar, which was also consistent with intrinsic fluorescence results [Murakami, K., & Berliner, L. J. (1983) Biochemistry 22, 3370-3374]. The paramagnetic cation Mn(II) bound to the strong calcium binding site on apo-alpha-LA but also to the weak secondary Ca(II) binding site(s) on Ca(II)-alpha-LA. It was also found that Co(II) bound to some secondary sites on Ca(II)-alpha-LA that overlapped the weak calcium site. All of the lanthanide shift reagents [Pr(III), Eu(III), Tb(III), Dy(III), Tm(III), Yb(III)] bound under slow exchange conditions; their relative affinities for apo-alpha-lactalbumin from competitive binding experiments were Dy(III), Tb(III), and Pr(III) greater than Ca(II) greater than Yb(III).  相似文献   

17.
Metal binding to the iron storage protein apoferritin is the first step in the process by which iron accumulates within the protein shell. In the present study, the stoichiometry of metal binding to apoferritin in solution has been examined using the probe ions Mn(II), VO(IV), and Cd(II) in conjunction with EPR spectroscopic and cadmium ion selective electrode measurements. Binding studies were carried out with the individual ions, in competition with one another, and in competition with Fe(II), Fe(III), and Tb(III). All three probe ions show binding stoichiometries near 0.3 and 0.7 metal ion per subunit, close to the theoretically predicted values of 0.33 and 0.67 for the binding of one and two metal ions, respectively, per three subunits. These results in conjunction with other data are consistent with the binding of one, and possibly two, metal ions within each of the eight hydrophilic channels which are located on 3-fold axes leading to the interior of the protein. Pairs of cadmium binding sites have been located in these channels by x-ray crystallography (Rice, D. W., Ford, G. C., White, J. L., Smith, J. M. A., and Harrison, P. M. (1983) Adv. Inorg. Biochem. 5, 39-49). The possibility that some metal binding occurs elsewhere on the protein is not precluded by the present data, however. In competition experiments between various metal ions, approximately 0.3 metal ion per subunit is readily displaced implying common binding sites in the channels for all of them. The stoichiometry of Mn(II) displacement by Fe(II) is less clear. Oxidation of Fe(II) to Fe(III) by molecular oxygen in the presence of Mn(II) regenerates some Mn(II) binding on the protein, suggesting migration of iron(III) to other protein sites, or perhaps to core.  相似文献   

18.
Phospholipid-binding properties of bovine factor V and factor Va.   总被引:5,自引:0,他引:5  
J W Bloom  M E Nesheim  K G Mann 《Biochemistry》1979,18(20):4419-4425
Factor V and factor Va binding to single bilayer phospholipid vesicles was investigated by light-scattering intensity measurements. This technique allows the measurement of free and phospholipid-bound protein concentrations from which equilibrium constants can be obtained. As controls, the Ca2+-dependent phospholipid binding of prothrombin and factor X were also studied. The average values obtained for the dissociation constants (Kd) and lipid to protein ratio at saturation, moles/mole (n), for prothrombin (Kd = 2.3 X 10(-6) M, n = 104) and factor X (Kd = 2.5 X 10(-6) M, n = 46) binding to vesicles containing 25% Folch fraction III and 75% phosphatidylcholine in the presence of 2 mM Ca2+ were in agreement with those reported in the literature. The average factor V and factor Va values for the dissociation constants and lipid to protein ratio at saturation (moles/mole) were Kd = 7.2 X 10(-8) M and n = 270 for factor V and Kd = 4.4 X 10(-7) M and n = 76 for factor Va. In contrast to prothrombin and factor X, factor V and factor Va demonstrated Ca2+-independent lipid binding. In addition, the number of factor V and factor Va molecules bound per vesicle was found to be dependent both on the phosphatidylserine content of the vesicle and the ionic strength of the buffer.  相似文献   

19.
Optical stopped-flow techniques have been used to determine the dissociation rate constants (koff) for the lanthanide(III) ions from carp (pI 4.25) parvalbumin. For most of the 13 different lanthanides studied, the release kinetics were diphasic, composed of both a fast phase (whose rate varied across the series, La3+ leads to Lu3+, between the limits -1.2 less than or equal to log kFAST less than or equal to -0.7) and a slower phase (whose rate varied across the series, La3+ leads to Lu3+, between the limits -1.2 greater than or equal to log kSLOW greater than or equal to -2.9). In addition, the La3+- and Lu3+-induced changes in the 270-MHz proton nuclear magnetic resonance spectrum of parvalbumin were used to calculate the dissociation constants for these specific lanthanides from the two high-affinity Ca2+ binding sites. The KD for one site appears to remain constant across the lanthanide series, determined to be 4.8 X 10(-11) M for both La3+ and Lu3+. The other site, however, is evidently quite sensitive to the nature of the bound Ln3+ ion and shows a strong preference for La3+ (KD,La = 2.0 X 10(-11) M; KD,Lu = 3.6 X 10(-10) M). We conclude from these observations that reports of nearly indistinguishable CD/EF binding site affinities for parvalbumin complexes of the middle-weight lanthanides (i.e., Eu3+, Gd3+, and Tb3+) are quite reasonable in view of the crossover in relative CD/EF site affinities across the lanthanide series.  相似文献   

20.
myo-Inositol monophosphatase catalyzes dephosphorylation of the synthetic substrate antranyloyl-2′-AMP. Binding of this fluorescent substrate to Tb(III)-monophosphatase was monitored by luminescence spectroscopy. The anthraniloyl chromophore excited at 330 nm sensitizes the long lived luminescence of enzyme bound Tb(III) at 490, 545, 585 and 620 nm. Assuming a mechanism of radiationless energy transfer, the actual distance of separation between the donor anthraniloyl moiety and the acceptor Tb(III) was calculated to be R = 10Å. The binding studies support the earlier observation of Bone et al. (Proc. Natl. Acad. Sci. USA 89 (1992) 10031–10035) that the substrate and the lanthanide Gd(III) interact with a common binding domain of the protein. The catalytic activity of the monophosphatase is completely dependent upon Mg(II) ions which elicit changes in the secondary structure of the protein as revealed by circular dichroism measurements. Binding of Mg(II) ions tend to stabilize the secondary structure of the phosphatase against guanidinium-HCl denaturation.  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