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 62 毫秒
1.
Lacking the extraordinary thermal stability of its metal-bound forms, apo-alpha-parvalbumin from rat muscle assumes two distinct conformations in aqueous solution. At 25 degrees C, its highly structured form predominates (Keq = 5.7; delta G degree = -4.3 kJ X mol-1); as deduced from both 1H NMR and circular dichroism (CD) spectroscopy, this conformation is exceedingly similar to those of its Mg(II)-, Ca(II)-, and Lu(III)-bound forms. The temperature dependences of several well-resolved aromatic and upfield-shifted methyl 1H NMR resonances and several CD bands indicate that the native, highly helical structure of rat apo-alpha-parvalbumin is unfolded by a concerted mechanism, showing no indication of partially structured intermediates. The melting temperature, TM, of rat apo-alpha-parvalbumin is 35 +/- 0.5 degrees C as calculated by both spectroscopic techniques. By 45 degrees C, rat apo-alpha-parvalbumin unfolds entirely, losing the tertiary structure that characterizes its folded form: not only are the ring-current-shifted aromatic and methyl 1H NMR resonances leveled, but the 262- and 269-nm CD bands are also severely reduced. As judged by the decrease in the negative ellipticity of the 222-nm CD band, this less-structured form of rat apo-alpha-parvalbumin shows an approximate 50% loss in apparent alpha-helical content compared to its folded state. Several changes in the 1H NMR spectrum of rat apo-alpha-parvalbumin were exceptionally informative probes of the specific conformational changes that accompany metal ion binding and metal ion exchange. In particular, the line intensities of the ortho proton resonance of Phe-47, the unassigned downfield-shifted alpha-CH resonances from the beta-sheet contacts between the metal-binding loops, the C2H resonance of His-48, and the epsilon-CH3 resonance of an unassigned Met residue were monitored as a function of added metal to determine the stability constants of several metal ion-parvalbumin complexes. We conclude that Mg(II) binds to the CD and EF sites independently, its affinity for the EF site being almost twice that for the CD site. Mg(II)----Ca(II) exchange showed that the CD-site Mg(II) is displaced first, in contrast to Lu(III)'s preferential displacement of the EF-site Ca(II) as determined from the Ca(II)----Lu(III) exchange experiments.(ABSTRACT TRUNCATED AT 400 WORDS)  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.
High resolution 1H nuclear magnetic resonance spectroscopy and optical stopped-flow techniques have been used to study the metal binding properties of a site-specific mutant of bacterial recombinant oncomodulin in which glutamate has replaced a liganding aspartate at position 59 in the CD calcium-binding site. In particular we have followed the replacement of calcium by lutetium in bacterial recombinant oncomodulin and D59E oncomodulin to provide a measure of the protein's preferences for metal ions of different ionic radii. The result of the Asp----Glu substitution is to make the mutant oncomodulin more similar to rat parvalbumin in terms of its relative CD- and EF-domain affinities for lutetium(III), that is to increase its affinity for metal ions with smaller ionic radii. This finding supports the original hypothesis that the presence of Asp at sequence position 59 is an important factor in the reduced preference of the CD site of oncomodulin for smaller metals such as magnesium (Williams, T. C., Corson, D. C., Sykes, B. D., and MacManus, J. P. (1987) J. Biol. Chem. 262, 6248-6256). However, our studies show that both the CD and the EF sites are affected by this single residue substitution suggesting that many factors play a role in the metal binding affinity and interaction between the two sites.  相似文献   

4.
113Cd-n.m.r. studies were used to investigate the binding of the lanthanide ions La3+, Gd3+, Tb3+, Yb3+ and Lu3+ to parvalbumins. It was shown that lanthanide ions with a smaller ionic radius bind sequentially to Cd2+-saturated parvalbumin, whereas those with a larger ionic radius bind with similar affinity to both the CD site and the EF site. The smallest ion, Lu3+, does in fact not compete significantly with Cd2+ for the CD site in carp parvalbumin, but appears to bind only to the EF site. This preference of the smaller lanthanide ions for the EF site was used to assign the n.m.r. signals for protein-bound 113Cd. By using Cd n.m.r. and Tb3+ fluorescence it was also shown for alpha-lineage parvalbumin from pike that these proteins possess a third site that can bind lanthanide ions. This site is, however, much weaker than in the beta-lineage parvalbumins. It was used to assign the 113Cd resonances from protein-bound Cd2+ ions in the spectrum of pike pI5.0 parvalbumin.  相似文献   

5.
Ca2+ binding to the wild type recombinant oncomodulin was studied by equilibrium flow dialysis in the absence and presence of 1, 2, and 10 mM Mg2+. Direct Mg2(+)-binding experiments were carried out by the Hummel-Dryer gel filtration technique. These studies revealed that in the absence of Mg2+ oncomodulin binds two Ca2+ with KCa = 2.2 x 10(7) and 1.7 x 10(6) M-1, respectively. In the absence of Ca2+ the protein binds only one Mg2+ with KMg = 4.0 x 10(3) M-1.Mg2+ antagonizes Ca2+ binding at the high affinity site according to the rule of direct competition. Ca2+ binding to the low affinity site is only slightly affected by Mg2+, so that in the presence of 2-3 mM Mg2+ the two sites have apparently an equal affinity for Ca2+. Microcalorimetry showed that, in spite of the different affinities of the two Ca2(+)-binding sites, delta H0 for the binding of each Ca2+ is identical and exothermic for -18.9 kJ/site. It follows that the entropy gain upon binding of Ca2+ is +77.1 J K-1 site-1 for the high affinity Ca2(+)-Mg2+ site and +56.0 J K-1 site-1 for the low affinity Ca2(+)-specific site. Mg2+ binding is endothermic for +13 kJ/site with an entropy change of +111 J K-1 site-1. The thermodynamic characteristics of the Ca2(+)-Mg2+ site resemble most those of site II (the so-called EF domain) of toad alpha-parvalbumin. The characteristics of Ca2+ binding to the specific site (likely the CD domain) are different from those of the Ca2+ specific sites in troponin C and in calmodulin and suggest that in oncomodulin hydrophobic forces do not play a predominant role in the binding process at the specific site.  相似文献   

6.
Replacement of the aspartate residue at position 59 of rat oncomodulin by glutamate by oligonucleotide-directed mutagenesis has afforded a protein which more closely resembles rat parvalbumin, at least judged by its interaction with the luminescent lanthanide ion Eu3+. The single-peak 7F0----5D0 spectrum observed at pH 5.0 with the fully bound wild-type protein is replaced by one which clearly shows two features at 5791 and 5796 A, arising from Eu3+ ions bound at the CD and EF sites, respectively. Furthermore, the pH dependence of the spectrum is substantially altered; the pKa observed for the CD domain, in which aspartate 59 residues, is shifted upward from pH 6.0 for the wild-type recombinant protein to pH 6.8 in the D59E mutant. Moreover, the maximum in the high-pH spectrum is shifted from 5781 to 5784 A. All three changes are indicative of a CD binding domain having increased parvalbumin-like character. Interestingly, however, the D59E substitution has only a modest effect on the Ca2+- and Mg2+-binding properties of the CD domain. For the wild-type protein, KCa = 7.8 x 10(-7) M and KMg = 3 x 10(-3) M. These affinities are more than an order of magnitude weaker than those seen for various parvalbumins and substantiate previous claims for calcium specificity made for the oncomodulin CD domain. Replacement of aspartate 59 by glutamate resulted in minor increases in affinity of the CD domain for Ca2+ (KCa = 5.5 x 10(-7) M) and Mg2+ (KMg = 1 x 10(-3) M). These findings strongly suggest that residues in oncomodulin besides aspartate 59 are important determinants of the observed calcium specificity of the CD calcium-binding domain. The consequences of the substitution at residue 59 appear to be confined to the CD domain. For the EF site in wild-type recombinant oncomodulin, KCa = 4.2 x 10(-8) M and KMg = 1.6 x 10(-4) M. The corresponding values for the D59E site-specific variant are identical within experimental error (KCa = 4.2 x 10(-8) M and KMg = 1.8 x 10(-4) M).  相似文献   

7.
The binding of cations by parvalbumins was studied by the proton relaxation enhancement (PRE) method using the paramagnetic probes Gd(III) and Mn(II). Gd(III) appears as a specific probe of the primary sites CD and EF with the following binding parameters: n = 2, KdGd = 0.5 x 10(-11) M and epsilon b = 2.3. The low value of epsilon b is the result of a nearly complete dehydration of the protein bound ions. Competition experiments between Gd(III) and various diamagnetic cations show the following order of affinity for the EF and CD sites: Mg2+ less than Zn2+ less than Sr2+ less than Ca2+ less than Cd2+ less than La3+ less than or equal to Gd3+. Mn 2+ is a specific probe of a secondary site with the following binding parameters: n = 1, KdMn = 0.6 x 10(-3) M and epsilon b = 17. The high value of epsilon b suggests that the protein bound Mn(II) has retained most of its hydration shell. Competition experiments between (Mn(II) and different cations show similar affinities for this site: Ca2+ less than or equal to Mg2+ less than or equal to Cd2+ less than or equal to Mn2+. This secondary site is located near the EF primary site.  相似文献   

8.
The 7F0----5D0 transition of Eu3+ was used to probe the metal-binding domains of rat oncomodulin and rat parvalbumin. Two distinct differences between the two proteins were observed. The first relates to the pH-dependent behavior of their 7F0----5D0 spectra, a phenomenon noted previously for other paravalbumins. In the case of rat parvalbumin, the spectral features associated with both metal-binding sites titrate concomitantly (pK alpha = 8.2); however, in the case of oncomodulin, the two sites titrate sequentially (pK alpha = 6.3 for the CD site; pK alpha = 8.3 for EF site). The proteins also contrast with regard to their discrimination for Eu3+ over Ca2+. The CD and EF sites in rat parvalbumin both display a large preference for Eu3+: (KCa/KEu)CD = 143 +/- 11 and (KCa/KEu)EF = 191 +/- 30. However, in the case of oncomodulin, although the EF site of oncomodulin greatly prefers the trivalent lanthanide ion (KCa/KEu = 300 +/- 80), the CD site exhibits a relatively minor preference (KCa/KEu = 11 +/- 1).  相似文献   

9.
A series of modified parvalbumins, differing only in length of alpha-helix F at the C-terminus, was prepared by carboxypeptidase-mediated digestions of the beta-lineage parvalbumin (pI = 4.25) from carp (N; 108 residues). Removal of Ala-108 to form the N-1 derivative (des-Ala108,Lys107-parvalbumin) only slightly alters the protein's ability to chelate Ca(II) or lanthanides(III). Analysis of the kinetics of their Yb(III) off-rates by optical stopped-flow techniques, determination of their Lu(III)-binding constants by high-resolution 1H NMR methods, and inspection of their solution structures by Yb(III)-shifted 1H NMR techniques indicate N-1 and N-2 are very similar to N (0.1-0.2 M KCl; pH 6-7; 23-55 degrees C). However, removal of the next one or two residues, Val-106 or Val-106/Leu-105, to generate the N-3 and N-4 derivatives severely alters the metal ion binding characteristics of the protein. Although two Yb(III) off-rates are observed for N-3, both are faster than that for the unmodified protein: kCD by a factor of 2 and kEF by a factor of 2200. Removal of Ala-104 and Ala-104/Thr-103 to give a mixture of N-5 and N-6 derivatives eliminates the slow-release site altogether, the single observable koff being 20-30 times faster than release of Yb(III) from the CD site of native parvalbumin. Removal of the C-terminal alpha-helix by digestion through Phe-102 to give N-7 destabilizes the entire protein structure as judged both by the random-coil appearance of its 1H NMR spectrum and by its aberrant kinetics. Although one abnormally fast koff is still observed at micromolar concentrations, Ln(III) chelation tends to precipitate N-7 at higher parvalbumin concentrations (1-3 mM). In contrast to the critical instability of the N-3 through N-7 derivatives, the remarkable stability of the N-1 and N-2 forms of carp parvalbumin may be attributed to the maintainance of two key structural features: an ion pair bond between the negatively charged C-terminal carboxyl function and the protonated epsilon-NH3+ of Lys-27 and hydrophobic interactions of the inner side of helix F with residues in the protein's core.  相似文献   

10.
The Eu(III)7F0-->5D0 excitation spectra of the parvalbumins are highly pH-dependent. Below pH 6.0, they exhibit a sharp, partially resolved doublet centered near 5,795 A. However, as the pH is raised, the spectrum becomes increasingly dominated by a much broader signal near 5,784 A. This behavior has been traced to the Eu(III) ion bound at the CD site, but the identity of the moiety undergoing deprotonation remains uncertain. Site-specific mutagenesis studies on the parvalbumin-like protein known as oncomodulin now suggest that the species in question is a liganding serine hydroxyl group. Specifically, replacement of serine-55 by aspartate (the residue present at the corresponding position in the EF site) affords a protein that retains two functional lanthanide binding sites, but fails to undergo the pH-dependent spectral alteration. By contrast, replacement of aspartate-59 by glycine (the corresponding EF site residue) fails to abolish the pH-dependent behavior.  相似文献   

11.
Close coorelation of atomic absorption measurements for Ca(II) contents indicates that from pH 5.8-7.4 a twentyfold excess of EGTA1 removes but one of two Ca(II) from carp parvalbumin. Thus binding of the two Ca(II) appears to be noncooperative. The maximum in emission intensity observed at a nonintegral 1.4-1.7 equivs of added Tb(III) is shown to be due to quenching by excess Tb(III). The emission intensity at the maximum increased 40% upon dialysis to remove Tb(III) not bound in the CD or EF sites. Atomic absorption results show that both Ca(CD) and Ca(EF) of native parvalbumin are easily replaced by Tb(III). Emission of Tb(EF) is not quenched by Tb(CD), but by solution Tb(III) bound at a third site, perhaps the single water molecule bound to Tb(EF). Labeling of the single sulfhydryl group with a trifluoroacetonyl gorup yields a protein with ultraviolet circular dichroism, emission, and circularly polarized emission spectra closely similar to those of native parvalbumin.  相似文献   

12.
The substitution of trivalent lanthanide ions for Ca(II) in the Ca(II)-DEPENDENT ACTIVATION OF BOVINE Factor X by the coagulant protein of Russell's viper venom was studied at pH 6.8. Factor X contains two high affinity metal binding sites which bind Gd(III), Sm(III), and Yb(III) with a Kd of about 4 X 10-7 M and four to six lower affinity metal binding sites which bind Gd(III), Sm(III) with a Kd of about 1.5 X 10-5M. In comparison, 1 mol of Factor X binds 2 mol of Ca(II) with a Kd of 3 X 10-4M and weakly binds many additional Ca(II) ions. No binding of Gd(III) to the venom protein was observed. Dy(III), Yb(III), Tb(III), Gd(III), Eu(III), La(III), AND Nd(III) cannot substitute for Ca(II) in the Ca(II)-dependent activation of Factor X by the venom protein at pH 6.8. Kinetic data consistent with the models of competitive inhibition of Ca(II) by Nd(III) yielded a Ki of 1 to 4 X 10-6M. The substitution of lanthanide ions for Ca(II) to promote protein complex formation of Factor X-metal-venom protein without the activation of Factor X facilitated the purification of the coagulant protein from crude venom by affinity chromatography. Using a column containing Factor X covalently bound to agarose which was equilibrated in 10 mM Nd(III), Tb(III), Gd(III), or La(III), the coagulant protein was purified 10-fold in 40% yield from crude venom and migrated as a single band on gel electrophoresis in sodium dodecyl sulfate. These data suggest that lanthanide ions complete with Ca(II) for the metal binding sites of Factor X and facilitate the formation of a nonproductive ternary complex of venom protein-Factor X-metal. Tb(III) fluorescence, with emission maxima at 490 and 545 nm, is enhanced 10,000-fold in the presence of Factor X. The study of the participation of an energy donor intrinsic to Factor X in energy transfer to Tb(III) may be useful in the characterization of the metal binding sites of Factor X.  相似文献   

13.
13C NMR spectra are presented for the calcium binding protein parvalbumin (pI 4.25) from carp muscle in several different metal bound forms: with Ca2+ in both the CD and EF calcium binding sites, with Cd2+ in both sites, with 113Cd2+ in both sites, and with 113Cd2+ in the CD site and Lu3+ in the EF site. The different metals differentially shift the 13C NMR resonances of the protein ligands involved in chelation of the metal ion. In addition, direct 13C-113Cd spin-spin coupling is observed which allows the assignment of protein carbonyl and carboxyl 13C NMR resonances to ligands directly interacting with the metal ions in the CD and EF binding sites. The displacement of 113Cd2+ from the EF site by Lu3+ further allows these resonances to be assigned to the CD or EF site. The occupancy of the two sites in the two cadmium species and in the mixed Cd2+/Lu3+ species is verified by 113Cd NMR. The resolution in these 113Cd NMR spectra is sufficient to demonstrate direct interaction between the two metal binding sites.  相似文献   

14.
Henzl MT  Agah S  Larson JD 《Biochemistry》2003,42(12):3594-3607
We have examined the metal ion-binding domains from rat alpha and beta parvalbumin. We find that the CD-EF fragments differ markedly in their tendency to self-associate. Whereas Ca(2+)-free alpha CD-EF is monomeric, the Ca(2+)-free beta peptide dimerizes weakly (K(2) = 2400 +/- 200 M(-1)). In buffer containing 1.0 mM Ca(2+), the apparent dimerization constant for beta CD-EF (191,000 +/- 29,000 M(-1)) is more than 50 times that of alpha (3400 +/- 200 M(-1)). Alpha CD-EF binds two Ca(2+) with positive cooperativity. Titration calorimetry data afford binding constants of 3.7(0.1) x 10(3) M(-1) and 8.6(0.2) x 10(4) M(-1). Beta CD-EF also binds two Ca(2+) cooperatively but with lower affinity. Equilibrium dialysis yields Adair constants of 4.2(0.1) x 10(3) and 6.1(0.2) x 10(3) M(-1). Significantly, the difference in Ca(2+) affinity is substantially smaller than that observed for the full-length proteins-suggesting that the AB domain can modulate divalent ion affinity. Analysis of beta calorimetry data requires explicit consideration of the self-association behavior. Data collected at low CD-EF concentration are consistent with preferential occupation of the EF site, dimerization of singly bound monomers, and cooperative filling of the CD sites. At higher concentrations, apo-protein dimerization can apparently precede cooperative occupation of the EF sites. In the presence of Ca(2+), alpha CD-EF exhibits higher thermal stability, consistent with its higher Ca(2+) affinity. However, the beta melting temperature shows greater concentration dependence, consistent with its greater tendency to dimerize. Neither fragment exhibits a sigmoidal melting curve in the Ca(2+)-free state, suggesting that the apo-peptides are disordered.  相似文献   

15.
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   

16.
Parvalbumin (PV) and the homologous protein oncomodulin (OM) contain three EF-hand motifs, but the first site (AB) cannot bind Ca2+. Here we aimed to recreate the putative ancestral proteins [D19-28E]PV and [D19-28E]OM by replacing the 10-residue-long nonfunctional loop in the AB site by a 12-residue canonical loop. To create an optical conformational probe we also expressed the homologs with a F102W replacement. Unexpectedly, in none of the proteins did the mutation reactivate the AB site. The AB-remodeled parvalbumins bind two Ca2+ ions with strong positive cooperativity (nH = 2) and moderate affinity ([Ca2+]0.5 = 2 microM), compared with [Ca2+]0.5 = 37 nM and nH = 1 for the wild-type protein. Increasing Mg2+ concentrations changed nH from 2 to 0.65, but without modification of the [Ca2+]0. 5-value. CD revealed that the Ca2+ and Mg2+ forms of the remodeled parvalbumins lost one-third of their alpha helix content compared with the Ca2+ form of wild-type parvalbumin. However, the microenvironment of single Trp residues in the hydrophobic cores, monitored using intrinsic fluorescence and difference optical density, is the same. The metal-free remodeled parvalbumins possess unfolded conformations. The AB-remodeled oncomodulins also bind two Ca2+ with [Ca2+]0.5 = 43 microM and nH = 1.45. Mg2+ does not affect Ca2+ binding. Again the Ca2+ forms display two-thirds of the alpha-helical content in the wild-type, while their core is still strongly hydrophobic as monitored by Trp and Tyr fluorescence. The metal-free oncomodulins are partially unfolded and seem not to possess a hydrophobic core. Our data indicate that AB-remodeled parvalbumin has the potential to regulate cell functions, whereas it is unlikely that [D19-28E]OM can play a regulatory role in vivo. The predicted evolution of the AB site from a canonical to an abortive EF-hand may have been dictated by the need for stronger interaction with Mg2+ and Ca2+, and a high conformational stability of the metal-free forms.  相似文献   

17.
The luminescent isomorphous Ca2+ analogue, Tb3+, can be bound in the 12-amino acid metal binding sites of proteins of the EF hand family, and its luminescence can be enhanced by energy transfer from a nearby aromatic amino acid. Tb3+ can be used as a sensitive luminescent probe of the structure and function of these proteins. The effect of changing the molecular environment around Tb3+ on its luminescence was studied using native Cod III parvalbumin and site-directed mutants of both oncomodulin and calmodulin. Titrations of these proteins showed stoichiometries of fill corresponding to the number of Ca2+ binding loops present. Tryptophan in binding loop position 7 best enhanced Tb3+ luminescence in the oncomodulin mutant Y57W, as well as VU-9 (F99W) and VU-32 (T26W) calmodulin. Excitation spectra of Y57F, F102W, Y65W oncomodulin, and Cod III parvalbumin revealed that the principal Tb3+ luminescence donor residues were phenylalanine or tyrosine located in position 7 of a loop, despite the presence of other nearby donors, including tryptophan. Spectra also revealed conformational differences between the Ca2+- and Tb(3+)-bound forms. An alternate binding loop, based on Tb3+ binding to model peptides, was inserted into the CD loop of oncomodulin by cassette mutagenesis. The order of fill of Tb3+ in this protein reversed, with the mutated loop binding Tb3+ first. This indicates a much higher affinity for the consensus-based mutant loop. The mutant loop inserted into oncomodulin had 32 times more Tb3+ luminescence than the identical synthetic peptide, despite having the same donor tryptophan and metal binding ligands. In this paper, a ranking of sensitivity of luminescence of bound Tb3+ is made among this subset of calcium binding proteins. This ranking is interpreted in light of the structural differences affecting Tb3+ luminescence enhancement intensity. The mechanism of energy transfer from an aromatic amino acid to Tb3+ is consistent with a short-range process involving the donor triplet state as described by Dexter (Dexter, D. L. (1953) J. Chem. Phys. 21, 836). This cautions against the use of the F?rster equation in approximating distances in these systems.  相似文献   

18.
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   

19.
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   

20.
The -X glutamate in a 33-residue model peptide comprising the CD site of carp parvalbumin 4.25 (ParvCD) was replaced with aspartate (ParvCD-XD) and the effect on calcium-dependent dimerization and calcium affinity assessed. The peptide ParvCD demonstrates a 10(5)-fold lower calcium affinity than the same site in the native protein. Both the ParvCD and ParvCD-XD model peptides fail to bind magnesium. The low calcium affinity and failure of the model ParvCD site to bind magnesium may be due to higher enthalpic costs of chelation by the -X glutamate. Replacement of the -X glutamate with an aspartate resulted in a twofold increase in the calcium affinity of both the monomer and dimer forms and a twofold increase in the calcium dependent dimerization of the peptide. A -X glutamate to aspartate replacement in 33-residue model peptides corresponding to bovine brain calmodulin site 3 (R. M. Procyshyn and R. E. Reid, Arch. Biochem. Biophys. 311, 425-429, 1994) and in Escherichia coli d-galactose-binding protein (S. K. Drake, K. L. Lee, and J. J. Falke, Biochemistry 35, 6697-6705, 1996) agree with results in the ParvCD site. However, in rat oncomodulin a -X glutamate to aspartate replacement increases calcium affinity (R. C. Hapak, P. J. Lammers, W. A. Palmisano, E. R. Birnbaum, and M. T. Henzl, J. Biol. Chem. 264, 18751-18760, 1989). The different effect of a -X glutamate to aspartate substitution in the different sites suggests site-specific factors dictating the thermodynamic contribution of the -X glutamate to calcium affinity.  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