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.
S T Ferreira 《Biochemistry》1989,28(26):10066-10072
The fluorescence properties of the single tryptophan residue in whiting parvalbumin were used to probe the dynamics of the protein matrix. Ca2+ binding caused a blue-shift in the emission (from lambda max = 339 to 315 nm) and a 2.5-fold increase in quantum yield. The fluorescence decay was nonexponential in both Ca2(+)-free and Ca2(+)-bound parvalbumin and was best described by Lorentzian lifetime distributions centered around two components: a major long-lived component at 2-5 ns and a small subnanosecond component. Raising the temperature from 8 to 45 degrees C resulted in a decrease in both the center (average) and width (dispersion) of the major lifetime distribution component, whereas the center, width, and fractional intensity of the fast component increased with temperature. Arrhenius activation energies of 1.3 and 0.3 kcal/mol were obtained in the absence and in the presence of Ca2+, respectively, from the temperature dependence of the center of the major lifetime distribution component. Direct anisotropy decay measurements of local tryptophan rotations yielded an activation energy of 2.3 kcal/mol in Ca2(+)-depleted parvalbumin and indicated a correlation between rotational rates and lifetime distribution parameters (center and width). Ca2+ binding produced a decrease in the width of the major lifetime distribution component and a decrease in tryptophan rotational mobility within the protein. There was a rough correlation between these two parameters with changes in Ca2+ and temperature, so that both measurements may be taken to indicate that the structure of Ca2(+)-bound parvalbumin was more rigid than in Ca2(+)-depleted parvalbumin.  相似文献   

2.
The possible structural changes in the major isotype of parvalbumin from the toad (Bufo bufo japonicus) skeletal muscle caused by Ca2+ and Mg2+ binding have been analyzed by microcalorimetric titrations. Parvalbumin was titrated with Ca2+ in both the absence and presence of Mg2+ and with Mg2+ in the absence of Ca2+, at pH 7.0, and at 5 degrees, 15 degrees, and 25 degrees C. The two sites in a molecule were equivalent on Mg2(+)-Ca2+ exchange, but distinguishable on Ca2+ and Mg2+ binding. The reactions of parvalbumin with Ca2+ are exothermic at every temperature in both the absence and presence of Mg2+, but those with Mg2+ are always endothermic except for the binding to site 1 at 25 degrees C. The magnitudes of the hydrophobic and internal vibrational contributions to the heat capacity and entropy changes of parvalbumin on Ca2+ and Mg2+ binding and Mg2(+)-Ca2+ exchange have been estimated by the empirical method of Sturtevant [Sturtevant, J. M. (1977) Proc. Natl Acad. Sci. USA 74, 2236-2240]. Although no major conformational changes were noted between Ca2(+)- and Mg2(+)-bound forms of toad parvalbumin, the conformational difference was larger in Ca2+ (or Mg2+) binding to site 1 than site 2. This may indicate that the metal-free form is much less stable than any form with Ca2+ (or Mg2+) bound at one site at least. On Mg2(+)-Ca2+ exchange, the vibrational as well as hydrophobic entropy is only slightly increased in a parallel manner. In contrast, on Ca2+ (or Mg2+) binding, the hydrophobic entropy increases but the vibrational entropy decreases; the former indicates the sequestering of nonpolar groups from the surface to the interior of a molecule, and the latter suggests that the overall structures are tightened on Ca2+ (or Mg2+) binding but loosened on Mg2(+)-Ca2+ exchange. Despite the clear distinctions in the thermodynamic features, the conformational changes of toad parvalbumin are essentially the same as those of the two isotypes of bullfrog parvalbumins on Ca2+ binding and Mg2(+)-Ca2+ exchange.  相似文献   

3.
The calcium-induced conformational changes of the 108-amino acid residue proteins, cod III parvalbumin and oncomodulin, were compared using tryptophan as a sensitive spectroscopic probe. As native oncomodulin is devoid of tryptophan, site-specific mutagenesis was performed to create a mutant protein in which tryptophan was placed in the identical position (residue 102) as the single tryptophan residue in cod III parvalbumin. The results showed that in the region probed by tryptophan-102, cod III parvalbumin experienced significantly greater changes in conformation upon decalcification compared to the oncomodulin mutant, F102W. Addition of 1 eq of Ca2+ produced greater than 90% of the total fluorescence response in F102W, while in cod III parvalbumin, only 74% of the total was observed. Cod III parvalbumin displayed a negligible response upon Mg2+ addition. In contrast, F102W did respond to Mg2+, but the response was considerably less when compared to Ca2+ addition. Time-resolved fluorescence showed that the tryptophan in both proteins existed in at least two conformational states in the presence of Ca2+ and at least three conformational states in its absence. Comparison with quantum yield measurements indicated that the local electronic environment of the tryptophan was significantly different in the two proteins. Collectively, these results demonstrate that both cod III parvalbumin and oncomodulin undergo Ca2(+)-specific conformational changes. However, oncomodulin is distinct from cod III parvalbumin in terms of the electronic environment of the hydrophobic core, the magnitude of the Ca2(+)-induced conformational changes, and the number of calcium ions required to modulate the major conformational changes.  相似文献   

4.
Fluorescence titration and fluorescence stopped-flow studies were performed on carp muscle parvalbumin components 1, 2, 3, and 5 (the latter three components were modified with a SH-directed fluorescent reagent, dansyl-L-cysteine). Apparent binding constants (Kapp) of Ca2+ to these components decrease in the order of component 2 (Kapp = 2.8 +/- 0.9 X 10(8) M-1) greater than component 1 (Kapp = 1.25 +/- 0.25 X 10(8) M-1) greater than component 3 = component 5 (Kapp = 4.0 +/- 0.5 X 10(7) M-1) in 30 mM KCl, 50 mM Na-cacodylate-HCl, pH 7.0 at 20 degrees C. The rate constant of the conformational change of parvalbumin induced by Ca2+ binding or removal decreases in the order of component 2 greater than component 1 greater than component 5 greater than or equal to component 3; that is, component 2 undergoes the fastest conformational change and component 3 the slowest in response to the rapid free Ca2+ concentration ([Ca2+]) change in the protein solution. The fluorescence titration curves and [Ca2+]-dependences of the rate constants are analyzed by a simple two-state model, (partially unfolded state) k1 in equilibrium k2 (folded state). It is shown that the equilibrium constant K = k1/k2 depends on the second power of [Ca2+], the rate constant k1 on the first power of [Ca2+] and k2 on the inverse first power of [Ca2+], respectively.  相似文献   

5.
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   

6.
The microenvironments of the histidines in three isoforms of Ca(II)-bound parvalbumin (carp, pI = 4.25; pike, pI = 5.00; rat, pI = 5.50) have been examined with 1H NMR techniques to probe their protonation characteristics and photochemically induced dynamic nuclear polarizability (photo-CIDNP). The histidine at position 26 (or 25), present in all three of these proteins, shows absolutely no photo-CIDNP enhancement of its C2H or C5H resonances. Nor does this nonpolarizable histidine possess a normal pKa: values range only from 4.20 for carp to 4.32 for pike to 4.44 for rat. The C2H and C5H resonances of the histidine in this carp isoform split into doublets as the pH is lowered. The magnitude of this splitting depends on the magnetic field strength, temperature, and pH; however, the line intensities within each doublet are temperature-independent. Although the crystal structure of carp parvalbumin indicates that His-26 is exposed to solvent [Kretsinger, R. H., & Nockolds, C. E. (1973) J. Biol. Chem. 248, 3313-3326], we conclude that in solution this residue, in its unprotonated state, is part of the hydrophobic core of the protein. In contrast, His-48 in rat parvalbumin and His-106 in pike III parvalbumin show dramatic photo-CIDNP enhancements of their C2H, C5H, and beta-CH2 1H NMR resonances. Combined with its nearly normal pKa, 6.14, and exchange-broadened C2H resonance, the photo-CIDNP enhancement results for His-48 indicate that its microenvironment differs little from random-coil exposure, consistent with its presumed position on the solvent surface of helix C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Vanadate-sensitized photocleavage of the Ca2(+)-ATPase of rabbit sarcoplasmic reticulum was observed upon illumination of sarcoplasmic reticulum vesicles or the purified Ca2(+)-ATPase by ultraviolet light in the presence of 1 mM monovanadate or decavanadate. The site of the photocleavage is influenced by the Ca2+ concentration of the medium. When the [Ca2+] is maintained below 10 nM by EGTA, the vanadate-catalyzed photocleavage yields fragments of approximately equal to 87 and approximately equal to 22 kDa, while in the presence of 2-20 mM Ca, polypeptides of 71 and 38 kDa are obtained as the principal cleavage products. These observations indicate that the site of the vanadate-catalyzed photocleavage is altered by changes in the conformation of Ca2(+)-ATPase. Selective tryptic proteolysis, at Arg-505-Ala-506, combined with covalent labeling of Lys-515 by fluorescein 5'-isothiocyanate and with the use of anti-ATPase antibodies of defined specificity, permitted the tentative allocation of the sites of photocleavage to the A fragment near the T2 cleavage site in the absence of Ca2+, and to the B fragment between Lys-515 and Asp-659 in the presence of 2-20 mM Ca2+. The loss of ATPase activity during illumination is accelerated by calcium in the presence of vanadate. The vanadate-catalyzed photocleavage in the presence of Ca2+ is consistent with the existence of an ATPase-Ca2(+)-vanadate complex (Markus et al. (1989) Biochemistry 28, 793-799).  相似文献   

8.
Rat hearts were depleted of Ca2+ (less than 10(-9) M) for 10 min, followed by 15 min of Ca2+-repletion. The calcium paradox injury occurs during Ca2+-repletion, after a period of calcium depletion. The calcium paradox injury was assessed by percent recovery (hemodynamics, [Ca2+]i, and energy levels) during Ca2+-repletion. A decrease in Na+ concentration during Ca2(+)-depletion did not allow for recovery during Ca2(+)-repletion, however 2.5% and 5% ethanol during Ca2(+)-depletion allowed for an approximate 50% recovery during Ca2(+)-repletion. A combination of ethanol (2.5% or 5%) with a low extracellular Na+ concentration (88 mM) allowed for complete recovery. Ethanol prevented a depletion of diastolic [Ca2+]i during Ca2(+)-depletion, and allowed for a return of normal diastolic [Ca2+]i during Ca2(+)-repletion. Ethanol modulates the activity of the Na+/Ca2+ exchanger and protects against the Ca2(+)-paradox injury.  相似文献   

9.
A Mrakovci?  S Oda  E Reisler 《Biochemistry》1979,18(26):5960-5965
Evidence for salt-induced changes in myosin light chains [dissociated by treatment with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB)], troponin-C (TnC), and parvalbumin was obtained from chymotryptic digestion, circular dichroism, fluorescence, and difference absorption studies. High salt (0.6 M NaCl) protects the DTNB light chain from proteolysis, increases its alpha-helical content, and quenches the tryptophan fluorescence. These effects are similar to the changes induced by Ca2+ but smaller in magnitude. TnC is affected by monovalent cations in a similar manner. Changes in the alpha-helical content resemble the effect of Ca2+. The enhancement of tyrosine fluorescence reflects conformational changes in the Ca2+-Mg2+ binding sites. The increase in the fluorescence of dansylaziridine-labeled TnC suggests perturbation of Ca2+-specific sites by salt. Cancellation of this effect by Mg2+ binding to the high-affinity sites is indicative of site-site interactions. In Whiting parvalbumin, salt-induced a perturbation of tryptophan absorption similar in nature to the Ca2+ effect.  相似文献   

10.
Upon substitution of Tb(III) for the most easily replaced Ca(II) from bovine cardiac TN-C, irradiation at 280 nm produces an emission at 545 nm from Tb(III) that is partially circularly polarized. Characteristics of these emission spectra produced by energy transfer from a tyrosyl side chain to a juxtaposed Tb(III) are virtually identical to those found in rabbit skeletal muscle TN-C and carp parvalbumin. A single homologous tyrosyl residue occurs in the two troponins and is in turn homologous to a phenylalanyl residue in parvalbumins. Addition of the other troponin subunits, TN-I and TN-T, to Tb(III)-TN-C weakens the total emission and completely quenches the circularly polarized emission.  相似文献   

11.
The ability of different ligands of glycoprotein (GP) IIb-IIIa (alphaIIb/beta3-integrin) to support platelet aggregation stimulated by activating anti-GP IIb-IIIa monoclonal antibody (monoAB) CRC54 has been investigated. Antibody CRC54 stimulated aggregation of washed platelets not only in the presence of fibrinogen, the main GP IIb-IIIa ligand, but also in the presence of von Willebrand factor (vWF). Unlike these ligands, fibronectin failed to support CRC54-induced aggregation. Fibrinogen and vWF dependent platelet aggregation was completely suppressed by GP IIb-IIIa antagonists--preparations Monafram (F(ab')2 fragments of monoAB that blocked GP IIb-IIIa receptor activity) and aggrastat (RGD-like peptidomimetic). However, aggregation stimulated in the presence of vWF was also completely inhibited by monoAB AK2 directed against GP Ib and capable of blocking its binding with vWF. CRC54-induced aggregation of platelets from patient with GP Ib deficiency in the presence of vWF was significantly lower than aggregation of platelets from normal donors and was not inhibited by anti-GP Ib antibody but still blocked by GP IIb-IIIa antagonist Monafram. Monafram also suppressed CRC54-stimulated platelet adhesion to plastic-adsorbed fibrinogen, vWF, and fibronectin. Unlike CRC54-induced platelet aggregation supported by fluid phase vWF, CRC54-induced adhesion to adsorbed vWF was not affected by anti-GP Ib antibody. Aggregation induced by CRC54 in the presence of fibrinogen and vWF was only partially suppressed by prostaglandin E1, an inhibitor of platelet activation, and was associated with serotonin release from platelet granules only when Ca2+ concentration was decreased from 1 mM (physiological level) to 0.1 mM. The data indicate that vWF supports CRC54-induced platelet aggregation via interaction with two receptors--GP IIb-IIIa and GP Ib. Aggregation induced by CRC54 in the presence of vWF or fibrinogen is only partially dependent on platelet activation and is accompanied with granule secretion only at low Ca2+ concentrations.  相似文献   

12.
Mammalian homologues of the Drosophila canonical transient receptor potential (TRP) proteins have been implicated to function as plasma membrane Ca(2+) channels. This study examined the role of TRPC1 in human neuroblastoma (SH-SY5Y) cells. SH-SY5Y cells treated with an exogenous neurotoxin, 1-methyl-4-phenylpyridinium ion (MPP(+)) significantly decreased TRPC1 protein levels. Confocal microscopy on SH-SY5Y cells treatment with MPP(+) showed decreased plasma membrane staining of TRPC1. Importantly, overexpression of TRPC1 reduced neurotoxicity induced by MPP(+). MPP(+)-induced alpha-synuclein expression was also suppressed by TRPC1 overexpression. Protection of SH-SY5Y cells against MPP(+) was significantly decreased upon the overexpression of antisense TRPC1 cDNA construct or the addition of a nonspecific transient receptor potential channel blocker lanthanum. Activation of TRPC1 by thapsigargin or carbachol decreased MPP(+) neurotoxicity, which was partially dependent on external Ca(2+). Staining of SH-SY5Y cells with an apoptotic marker (YO-PRO-1) showed that TRPC1 protects SH-SY5Y neuronal cells against apoptosis. Further, TRPC1 overexpression inhibited cytochrome c release and decreased Bax and Apaf-1 protein levels. Interpretation of the above data suggests that reduction in the cell surface expression of TRPC1 following MPP(+) treatment may be involved in dopaminergic neurodegeneration. Furthermore, TRPC1 may inhibit degenerative apoptotic signaling to provide neuroprotection against Parkinson's disease-inducing agents.  相似文献   

13.
Feng QL  Wu DM  Cui XL  Zhao HC  Lin YY  Zhao LY  Wu BW 《生理学报》2010,62(5):407-414
Considering that α-1 repeat region may be involved in the ion binding and translocation of Na(+)-Ca(2+) exchanger (NCX), it is possible that the antibodies against NCX α-1 repeat may have a crucial action on NCX activity. The aim of the present study is to investigate the effect of antibody against α-1 repeat (117-137), designated as α-1(117-137), on NCX activity. The antibody against the synthesized α-1(117-137) was prepared and affinity-purified. Whole-cell patch clamp technique was used to study the change of Na(+)-Ca(2+) exchange current (I(Na/Ca)) in adult rat cardiomyocytes. To evaluate the functional specificity of this antibody, its effects on L-type Ca(2+) current (I(Ca,L)), voltage-gated Na(+) current (I(Na)) and delayed rectifier K(+) current (I(K)) were also observed. The amino acid sequences of α-1(117-137) in NCX and residues 1?076-1?096 within L-type Ca(2+) channel were compared using EMBOSS Pairwise Alignment Algorithms. The results showed that outward and inward I(Na/Ca) were decreased by the antibody against α-1(117-137) dose-dependently in the concentration range from 10 to 160 nmol/L, with IC(50) values of 18.9 nmol/L and 22.4 nmol/L, respectively. Meanwhile, the antibody also decreased I(Ca,L) in a concentration-dependent manner with IC(50) of 22.7 nmol/L. No obvious effects of the antibody on I(Na) and I(K) were observed. Moreover, comparison of the amino acid sequences showed there was 23.8% sequence similarity between NCX α-1(117-137) and residues 1?076-1?096 within L-type Ca(2+) channel. These results suggest that antibody against α-1(117-137) is a blocking antibody to NCX and can also decrease I(Ca,L) in a concentration-dependent manner, while it does not have obvious effects on I(Na) and I(K).  相似文献   

14.
The component I isolated from carp red muscle has been characterized as a true parvalbumin, fairly different from carp parvalbumins described so far. The protein is antigenically related to the parvalbumin III from pike, which belongs to the so called parvalbumin lineage alpha. Immunological investigations on the location of the various carp parvalbumins reveal genuine variation in the pattern of these proteins according to organ and type of muscular tissue.  相似文献   

15.
Ca2+ and Mg2+ movements across the sarcoplasmic reticulum (SR) of frog skeletal muscle fibers were measured in situ by electron probe microanalysis of muscles rapidly frozen following a tetanus. At 400 ms following a 1.2-s tetanus at room temperature, the force had relaxed to base-line, and 0.3 mmol of Ca2+/liter of cytoplasmic H2O had been pumped by the SR, indicating that the in situ pumping of the SR Ca-ATPase is sufficiently high to account for the removal of Ca2+ from the Ca2+-specific sites of troponin (0.18 mmol of Ca2+-specific sites/liter of cytoplasmic H2O) and for the rate of relaxation from a tetanus at room temperature. The half-time of the return of the total 1.0 mmol of Ca2+/liter of cytoplasmic H2O released during a tetanus was 1.1 s, comparable to the slow Koff rate of Ca2+ from (carp) parvalbumin (1.0 s-1) and consistent with the hypothesis that the return of this Ca2+ to the terminal cisternae is rate-limited by the Ca2+ off-rate from parvalbumin. The return of the Mg2+ taken up by the terminal cisternae during a tetanus to resting levels was significantly slower than the time course of the Ca2+ movements, suggesting that the Mg2+ permeability of the SR in situ is low and may be transiently increased during tetanic stimulation.  相似文献   

16.
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   

17.
A basic, dimeric myotoxic protein, myotoxin II, purified from Bothrops asper venom has a similar molecular weight and is immunologically cross-reactive with antibodies raised to previously isolated B. asper phospholipases A2, except that it shows only 0.1% of the phospholipase activity against L-alpha-phosphatidylcholine in the presence of Triton X-100. Its 121 amino acid sequence, determined by automated Edman degradation, clearly identifies it as a Lys-49 phospholipase A2. Key amino acid differences between myotoxin II and phospholipase active proteins in the Ca2(+)-binding loop region, include Lys for Asp-49, Asn for Tyr-28, and Leu for Gly-32. The latter substitution has not previously been seen in Lys-49 proteins. Other substitutions near the amino terminus (Leu for Phe-5 and Gln for several different amino acids at position 11) may prove useful for identifying other Lys-49 proteins in viperid and crotalid venoms. Myotoxin II shows greater sequence identity with other Lys-49 proteins from different snake venoms (Agkistrodon piscivorus piscivorus, Bothrops atrox, and Trimeresurus flavoviridis) than with another phospholipase A2 active Asp-49 molecule isolated from the same B. asper venom. This work demonstrates that phospholipase activity per se, is not required in phospholipase molecules for either myotoxicity or edema inducing activities.  相似文献   

18.
Relative to the same site in oncomodulin, the CD ion-binding domain of rat parvalbumin exhibits much greater affinity for Ca2+ and Mg2+. As part of an effort to understand the structural basis for these differences, site-specific variants of oncomodulin have been prepared in which the amino acid residues at positions 52, 54, 57, 59, and 60 have been replaced with the residues present at the corresponding positions in rat parvalbumin. The proteins resulting from the single-site substitutions at residues 52, 54, and 57 are indistinguishable from the wild-type protein on the basis Eu3+ luminescence spectroscopy, and none of the three variants displays increased affinity for Ca2+. By contrast, the substitutions at residues 59 and 60 perturb both the Eu3+ luminescence parameters and the Ca2+ and Mg2+ affinities, and these differences are amplified when both replacements are simultaneously incorporated into the protein. The Eu3+ 7F0----5D0 spectrum of the double variant (D59E/G60E) at pH 5.0, with a maximum at 5796 A and pronounced shoulder at 5791 A, strongly resembles that obtained with pike parvalbumin. Consistent with this increased parvalbumin-like character, KCa is decreased from 0.78 microM (for the wild-type protein) to 0.41 microM, and KMg is decreased from 3.5 to 0.74 mM. Nevertheless, the affinity of the CD ion-binding domain in D59E/G60E for Ca2+ remains almost 2 orders of magnitude lower than the corresponding site in rat parvalbumin, strongly suggesting that residues besides those present in the binding loop are involved in dictating the metal ion-binding properties of the oncomodulin CD site.  相似文献   

19.
Lifetimes of phenylalanine, tyrosine and tryptophan self-fluorescence of three Ca2+-binding proteins (parvalbumins pI 4.47 and 3.95 and bovine alpha-lactalbumin) in the Ca2+-saturated state and without Ca2+ were measured on a device functioning in a channel of synchrotron radiation of the Lebedev Physical Institute electron accelerator C-60 with a single photon counting system. The decay curve of phenylalanine fluorescence of Ca2+-saturated parvalbumin pI 4.47 is two-exponential, which results from the presence of two subsystems of phenylalanine residues in this protein. Radiation of these subsystems is almost independent of one another. Detachment of Ca2+ from protein disturbs these subsystems. In case of tyrosine fluorescence of carp parvalbumin pI 3.95 a change in the quantum yield value of the stationary fluorescence induced by elimination of Ca2+ proceeds without a change of fluorescence lifetime. This seems to be related to the existence of static quenching of fluorescence in this case at the expense of complex formation between the chromophore and some adjacent quenching groups. Detachment of Ca2+ from alpha-lactalbumin induces conformational changes in its structure. The latter result in a transition of a number of tryptophane residues from its interior to the surface of the globule which is reflected in an increase of fluorescence quantum yield duration. It is concluded that in Ca2+-saturated alpha-lactalbumin some tryptophane residues are located near the quenching groups (dynamic quenching), most likely the disulfide bridges.  相似文献   

20.
Photoaffinity heterobifunctional cross-linking reagents are described that incorporate a 4-azidosalicylate group at one terminus and an N-hydroxysuccinimidyl ester at the other terminus with a "linking arm" of variable length separating the photoactive and electrophilic termini. Exposure of calmodulin (CaM) to succinimidyl N-[2-[(4-azidosalicyloyl)oxy]ethyl]suberamate (4C) led to monoadducts at Lys-21, Lys-75, and Lys-94. Separation of the monoadducts from CaM and polyadducts, radioiodination, and photolysis in the presence of human erythrocyte plasma membrane Ca2+, Mg2(+)-ATPase led to calcium-dependent cross-linking with 8% cross-linking efficiency.  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