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.
Lee YH  Tanner JJ  Larson JD  Henzl MT 《Biochemistry》2004,43(31):10008-10017
In model peptide systems, Ca2+ affinity is maximized in EF-hand motifs containing four carboxylates positioned on the +x and -x and +z and -z axes; introduction of a fifth carboxylate ligand reduces the affinity. However, in rat beta-parvalbumin, replacement of Ser-55 with aspartate heightens divalent ion affinity [Henzl, M. T., et al. (1996) Biochemistry 35, 5856-5869]. The corresponding alpha-parvalbumin variant (S55D/E59D) likewise exhibits elevated affinity [Henzl, M. T., et al. (2003) Anal. Biochem. 319, 216-233]. To determine whether these mutations produce a variation on the archetypal EF-hand coordination scheme, we have obtained high-resolution X-ray crystallographic data for alpha S55D/E59D. As anticipated, the aspartyl carboxylate replaces the serine hydroxyl at the +z coordination position. Interestingly, the Asp-59 carboxylate abandons the role it plays as an outer sphere ligand in wild-type rat beta, rotating away from the Ca2+ and, instead, forming a hydrogen bond with the amide of Glu-62. Superficially, the coordination sphere in the CD site of alpha S55D/E59D resembles that in the EF site. However, the orientation of the Asp-59 side chain is predicted to stabilize the D-helix, which may contribute to the heightened divalent ion affinity. DSC data indicate that the alpha S55D/E59D variant retains the capacity to bind 1 equiv of Na+. Consistent with this finding, when binding measurements are conducted in K(+)-containing buffer, divalent ion affinity is markedly higher. In 0.15 M KCl and 0.025 M Hepes-KOH (pH 7.4) at 5 degrees C, the macroscopic Ca2+ binding constants are 1.8 x 10(10) and 2.0 x 10(9) M(-1). The corresponding Mg2+ binding constants are 2.7 x 10(6) and 1.2 x 10(5) M(-1).  相似文献   

2.
We have examined permeation by Ca(2+) and Ba(2+), and block by Mg(2+), using whole-cell recordings from alpha1G T-type calcium channels stably expressed in HEK 293 cells. Without Mg(o)(2+), inward currents were comparable with Ca(2+) and Ba(2+). Surprisingly, three other results indicate that alpha1G is actually selective for Ca(2+) over Ba(2+). 1) Mg(2+) block is approximately 7-fold more potent with Ba(2+) than with Ca(2+). With near-physiological (1 mM) Mg(o)(2+), inward currents were approximately 3-fold larger with 2 mM Ca(2+) than with 2 mM Ba(2+). The stronger competition between Ca(2+) and Mg(2+) implies that Ca(2+) binds more tightly than Ba(2+). 2) Outward currents (carried by Na(+)) are blocked more strongly by Ca(2+) than by Ba(2+). 3) The reversal potential is more positive with Ca(2+) than with Ba(2+), thus P(Ca) > P(Ba). We conclude that alpha1G can distinguish Ca(2+) from Ba(2+), despite the similar inward currents in the absence of Mg(o)(2+). Our results can be explained by a 2-site, 3-barrier model if Ca(2+) enters the pore 2-fold more easily than Ba(2+) but exits the pore at a 2-fold lower rate.  相似文献   

3.
The administration of selective alpha(1) (phenylephrine)-, beta (isoproterenol)-, or mixed (epinephrine) adrenergic agonists induces a marked Mg(2+) extrusion from perfused rat livers. In the absence of extracellular Ca(2+), phenylephrine does not induce a detectable Mg(2+) extrusion, isoproterenol-induced Mg(2+) mobilization is unaffected, and epinephrine induces a net Mg(2+) extrusion that is lower than in the presence of extracellular Ca(2+) and quantitatively similar to that elicited by isoproterenol. In the absence of extracellular Na(+), no Mg(2+) is extruded from the liver irrespective of the agonist used. Similar results are observed in perfused livers stimulated by glucagon or 8-chloroadenosine 3', 5'-cyclic monophosphate. In the absence of extracellular Na(+) or Ca(2+), adrenergic-induced glucose extrusion from the liver is also markedly decreased. Together, these results indicate that liver cells extrude Mg(2+) primarily via a Na(+)-dependent mechanism. This extrusion pathway can be activated by the increase in cellular cAMP that follows the stimulation by glucagon or a specific beta-adrenergic receptor agonist or, alternatively, by the changes in cellular Ca(2+) induced by the stimulation of the alpha(1)-adrenoceptor. In addition, the stimulation of the alpha(1)-adrenoceptor appears to activate an auxiliary Ca(2+)-dependent Mg(2+) extrusion pathway. Finally, our data suggest that experimental conditions that affect Mg(2+) mobilization also interfere with glucose extrusion from liver cells.  相似文献   

4.
Henzl MT  Agah S  Larson JD 《Biochemistry》2004,43(29):9307-9319
Introduction of a fifth carboxylate into the ligand array of the CD site (via the combined S55D and E59D mutations) or the EF site (G98D) of rat alpha-parvalbumin substantially increases divalent ion affinity. This behavior, in conflict with that seen in model peptide systems, agrees with existing data for rat beta-parvalbumin [Henzl et al. (1996) Biochemistry 35, 5856-5869]. The complete analysis of the S55D/E59D double variant necessitated characterization of alpha E59D. Whereas the D59E mutation has minimal influence on beta CD site affinity, E59D has a major impact on the alpha CD site, lowering the apparent association constant by a factor of 14. The thermodynamic consequences of exchanging the rat alpha CD and EF site ligand arrays, which differ at the +z and -x coordination positions, were also examined. When the alpha CD array is imported into the EF site, it acquires a low-affinity phenotype, in agreement with previous findings for beta [Henzl et al. (1998) Biochemistry 37, 9101-9111]. However, when the EF ligand array is introduced into the alpha CD binding loop, it retains a high-affinity signature. This result, contrary to that observed in beta, suggests that the influence of the parvalbumin CD site environment supersedes the intrinsic behavior of the ligand array, a conclusion further supported by the disparate impact of the beta D59E and alpha E59D mutations.  相似文献   

5.
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   

6.
15N relaxation data for Ca(2+)-bound rat beta-parvalbumin (a.k.a. oncomodulin) were analyzed using the Lipari-Szabo formalism and compared with existing data for rat alpha-parvalbumin. Although the average S(2) values for the two proteins are very similar (0.85 for alpha, 0.84 for beta), residue-by-residue inspection reveals systematic differences. alpha tends to have the lower S(2) value in helical regions; beta tends to have the lower value in the loop regions. Rat beta was also examined in the Ca(2+)-free state. The 59 assigned residues displayed an average order parameter (0.90) significantly greater than the corresponding residues in the Ca(2+)-loaded form. The pentacarboxylate variants of rat beta-S55D and G98D-also were examined in the Ca(2+)-bound state. Although both mutations significantly heighten Ca(2+) affinity, they utilize distinct energetic strategies. S55D improves the Ca(2+)-binding enthalpy; G98D improves the binding entropy. They also show disparate peptide backbone dynamics. Whereas beta G98D displays an average order parameter (0.87) slightly greater than that of the wild-type protein, beta S55D displays an average order parameter (0.82) slightly lower than wild-type beta. Furthermore, whereas just two backbone N-H bonds in beta G98D show internal motion on the 20-200-psec timescale, fully 52 of the 93 residues analyzed in beta S55D show this behavior. These findings suggest that the increased electrostatic repulsion attendant to introduction of an additional carboxylate into the CD site ligand array impedes backbone vibrational motion throughout the molecule.  相似文献   

7.
The helix-loop-helix (i.e., EF-hand) Ca(2+) ion binding motif is characteristic of a large family of high-affinity calcium ion binding proteins, including the parvalbumins, oncomodulins and calmodulins. In this work we describe a set of molecular dynamics computations on the major parvalbumin from the silver hake (SHPV-B) and on functional fragments of this protein, consisting of the first four helical regions (the ABCD fragment), and the internal helix-loop- helix region (the CD fragment). In both whole protein and protein fragments (i.e., ABCD and CD fragments), the 9th loop residue in the calcium ion binding site in the CD helix-loop-helix region (the so-called "gateway" position) has been mutated from glutamic acid to aspartic acid. Aspartic acid is one of the most common residues found at the gateway position in other (non-parvalbumin) EF- hand proteins, but has never been found at the gateway position of any parvalbumin. (Interestingly, aspartic acid does occur at the gateway position in the closely related rat and human oncomodulins.) Consistent with experimental observations, the results of our molecular dynamics simulations show that incorporation of aspartic acid at the gateway position is very disruptive to the structural integrity of the calcium ion coordination site in the whole protein. The aspartic acid mutation is somewhat less disruptive to the calcium ion coordination sites in the two parvalbumin fragments (i.e., the ABCD and CD fragments), presumably due to the higher degree of motional freedom allowable in these protein fragments. One problem associated with the E59D whole protein variant is a prohibitively close approach of the aspartate carboxyl group to the CD calcium ion observed in the energy-minimized (pre-molecular dynamics) structure. This steric situation does not emerge during energy-minimization of the wild-type protein. The damage to the structural integrity of the calcium ion coordination site in the whole protein E59D variant is not relieved during the molecular dynamics simulation. In fact, during the course of the 300 picosecond simulation, all of the calcium ion ligands leave the primary coordination sphere. In addition, the conserved hydrogen- bonds (in the short beta-sheet structure) that links the CD site to the symmetry-related EF site (in the non-mutated whole protein) is also somewhat disrupted in the E59D whole protein variant. These results suggest that the Ca(2+) ion binding deficiencies in the CD loop are related, at least in part, to the unique interaction that exists between the paired CD and EF hands in the whole protein. Our theoretical results correlate well with previous studies on engineered EF-hand proteins and with all of our experimental evidence on whole silver hake parvalbumin and enzymatically-generated parvalbumin fragments.  相似文献   

8.
Integrin receptor alpha(2)beta(1) requires micromolar Ca(2+) to bind to collagen and to the peptide GPC(GPP)(5)GFOGER(GPP)(5)GPC (denoted GFOGER-GPP, where O represents hydroxyproline), which contains the minimum recognition sequence for the collagen-binding alpha(2) I-domain (Knight, C. G., Morton, L. F., Peachey, A. R., Tuckwell, D. S., Farndale, R. W., and Barnes, M. J. (2000) J. Biol. Chem. 275, 35-40). Platelet adhesion to these ligands is completely dependent on alpha(2)beta(1) in the presence of 2 mm Mg(2+). However, we show here that this interaction was abolished in the presence of 25 microm EGTA. Adhesion of Glanzmann's thrombasthenic platelets, which lack the fibrinogen receptor alpha(IIb)beta(3), was also inhibited by micromolar EGTA. Mg(2+)-dependent adhesion of platelets was restored by the addition of 10 microm Ca(2+), but millimolar Ca(2+) was inhibitory. Binding of isolated alpha(2)beta(1) to GFOGER-GPP was 70% inhibited by 50 microm EGTA but, as with intact platelets, was fully restored by the addition of micromolar Ca(2+). 2 mm Ca(2+) did not inhibit binding of isolated alpha(2)beta(1) to collagen or to GFOGER-GPP. Binding of recombinant alpha(2) I-domain was not inhibited by EGTA, nor did millimolar Ca(2+) inhibit binding. Our data suggest that high affinity Ca(2+) binding to alpha(2)beta(1), outside the I-domain, is essential for adhesion to collagen. This is the first demonstration of a Ca(2+) requirement in alpha(2)beta(1) function.  相似文献   

9.
Henzl MT  Agah S  Larson JD 《Biochemistry》2004,43(34):10906-10917
Association of the parvalbumin AB and CD-EF domains was examined in Hepes-buffered saline, pH 7.4, employing fragments from rat alpha and beta. All of the interactions require Ca(2+). In saturating Ca(2+), the alpha AB/alpha CD-EF (alpha/alpha) complex displays an association constant of (7.6 +/- 0.4) x 10(7) M(-1). Ca(2+)-binding data for a mixture of the alpha fragments are compatible with an identical two-site model, yielding an average binding constant of (8.5 +/- 0.2) x 10(5) M(-1). The beta/beta interaction is significantly weaker, exhibiting an association constant of (3.0 +/- 0.6) x 10(6) M(-1). The Ca(2+)-binding constants for beta/beta are likewise diminished, at (1.0 +/- 0.1) x 10(5) and (2.3 +/- 0.2) x 10(4) M(-1). The magnitude of the apparent DeltaDeltaG(degree)' for Ca(2+) binding by alpha/alpha and beta/beta, at 3.4 kcal/mol, approaches that measured for the intact proteins (3.6 kcal/mol) and is substantially larger than the 1.5 kcal/mol value previously measured for the isolated CD-EF domains. This result suggests that the AB domain can modulate the Ca(2+) affinities of the CD and EF sites. Interestingly, the heterologous alpha/beta complex displays a larger association constant [(6.6 +/- 0.4) x 10(6) M(-1)] than the homologous beta/beta complex and heightened Ca(2+) affinity [binding constants of (1.3 +/- 0.1) x 10(6) and (8.8 +/- 0.2) x 10(4) M(-1)]. By contrast, beta/alpha associates more weakly than alpha/alpha and exhibits sharply reduced affinity for Ca(2+). Thus, the interaction between the beta AB domain and beta CD-EF domain may act to attenuate Ca(2+) affinity in the intact protein.  相似文献   

10.
By analyzing, after expression in yeast and purification, the intrinsic fluorescence properties of point mutants of rabbit Ca(2+)-ATPase (SERCA1a) with alterations to amino acid residues in Ca(2+)-binding site I (E(771)), site II (E(309)), in both sites (D(800)), or in the nucleotide-binding domain (W(552)), we were able to follow the conformational changes associated with various steps in the ATPase catalytic cycle. Whereas Ca(2+) binding to purified wild-type (WT) ATPase in the absence of ATP leads to the rise in Trp fluorescence expected for the so-called E2 --> E1Ca(2) transition, the Ca(2+)-induced fluorescence rise is dramatically reduced for the E(309)Q mutant. As this purified E(309)Q mutant retains the ability to bind Ca(2+) at site I (but not at site II), we tentatively conclude that the protein reorganization induced by Ca(2+) binding at site II makes the major contribution to the overall Trp fluorescence changes observed upon Ca(2+) binding to both sites. Judging from the fluorescence response of W(552)F, similar to that of WT, these changes appear to be primarily due to membranous tryptophans, not to W(552). The same holds for the fluorescence rise observed upon phosphorylation from P(i) (the so-called E2 --> E2P transition). As for WT ATPase, Mg(2+) binding in the absence of Ca(2+) affects the fluorescence of the E(309)Q mutant, suggesting that this Mg(2+)-dependent fluorescence rise does not reflect binding of Mg(2+) to Ca(2+) sites; instead, Mg(2+) probably binds close to the catalytic site, or perhaps near transmembrane span M3, at a location recently revealed by Fe(2+)-catalyzed oxidative cleavage. Mutation of W(552) hardly affects ATP-induced fluorescence changes in the absence of Ca(2+), which are therefore mostly due to membranous Trp residues, demonstrating long-range communication between the nucleotide-binding domain and the membranous domain.  相似文献   

11.
Mg(2+) serves as a competitive antagonist against Ca(2+) in the high-affinity Ca(2+) activation site (A-site) and as an agonist of Ca(2+) in the low-affinity Ca(2+) inactivation site (I-site) of the ryanodine receptor (RyR), which mediates Ca(2+)-induced Ca(2+) release (CICR). This paper presents the quantitative determination of the affinities for Ca(2+) and Mg(2+) of A- and I-sites of RyR in frog skeletal muscles by measuring [(3)H]ryanodine binding to purified alpha- and beta-RyRs and CICR activity in skinned fibers. There was only a minor difference in affinity at most between alpha- and beta-RyRs. The A-site favored Ca(2+) 20- to 30-fold over Mg(2+), whereas the I-site was nonselective between the two cations. The RyR in situ showed fivefold higher affinities for Ca(2+) and Mg(2+) of both sites than the purified alpha- and beta-RyRs with unchanged cation selectivity. Adenine nucleotides, whose stimulating effect was found to be indistinguishable between free and complexed forms, did not alter the affinities for cations in either site, except for the increased maximum activity of RyR. Caffeine increased not only the affinity of the A-site for Ca(2+) alone, but also the maximum activity of RyR with otherwise minor changes. The results presented here suggest that the rate of CICR in frog skeletal muscles appears to be too low to explain the physiological Ca(2+) release, even though Mg(2+) inhibition disappears.  相似文献   

12.
BACKGROUND: The EF-hand family is a large set of Ca(2+)-binding proteins that contain characteristic helix-loop-helix binding motifs that are highly conserved in sequence. Members of this family include parvalbumin and many prominent regulatory proteins such as calmodulin and troponin C. EF-hand proteins are involved in a variety of physiological processes including cell-cycle regulation, second messenger production, muscle contraction, microtubule organization and vision. RESULTS: We have determined the structures of parvalbumin mutants designed to explore the role of the last coordinating residue of the Ca(2+)-binding loop. An E101D substitution has been made in the parvalbumin EF site. The substitution decreases the Ca(2+)-binding affinity 100-fold and increases the Mg(2+)-binding affinity 10-fold. Both the Ca(2+)- and Mg(2+)-bound structures have been determined, and a structural basis has been proposed for the metal-ion-binding properties. CONCLUSIONS: The E101D mutation does not affect the Mg(2+) coordination geometry of the binding loop, but it does pull the F helix 1.1 A towards the loop. The E101D-Ca(2+) structure reveals that this mutant cannot obtain the sevenfold coordination preferred by Ca(2+), presumably because of strain limits imposed by tertiary structure. Analysis of these results relative to previously reported structural information supports a model wherein the characteristics of the last coordinating residue and the plasticity of the Ca(2+)-binding loop delimit the allowable geometries for the coordinating sphere.  相似文献   

13.
The ability of membrane voltage to activate high conductance, calcium-activated (BK-type) K(+) channels is enhanced by cytosolic calcium (Ca(2+)). Activation is sensitive to a range of [Ca(2+)] that spans over four orders of magnitude. Here, we examine the activation of BK channels resulting from expression of cloned mouse Slo1 alpha subunits at [Ca(2+)] and [Mg(2+)] up to 100 mM. The half-activation voltage (V(0.5)) is steeply dependent on [Ca(2+)] in the micromolar range, but shows a tendency towards saturation over the range of 60-300 microM Ca(2+). As [Ca(2+)] is increased to millimolar levels, the V(0.5) is strongly shifted again to more negative potentials. When channels are activated by 300 microM Ca(2+), further addition of either mM Ca(2+) or mM Mg(2+) produces similar negative shifts in steady-state activation. Millimolar Mg(2+) also produces shifts of similar magnitude in the complete absence of Ca(2+). The ability of millimolar concentrations of divalent cations to shift activation is primarily correlated with a slowing of BK current deactivation. At voltages where millimolar elevations in [Ca(2+)] increase activation rates, addition of 10 mM Mg(2+) to 0 Ca(2+) produces little effect on activation time course, while markedly slowing deactivation. This suggests that Mg(2+) does not participate in Ca(2+)-dependent steps that influence current activation rate. We conclude that millimolar Mg(2+) and Ca(2+) concentrations interact with low affinity, relatively nonselective divalent cation binding sites that are distinct from higher affinity, Ca(2+)-selective binding sites that increase current activation rates. A symmetrical model with four independent higher affinity Ca(2+) binding steps, four voltage sensors, and four independent lower affinity Ca(2+)/Mg(2+) binding steps describes well the behavior of G-V curves over a range of Ca(2+) and Mg(2+). The ability of a broad range of [Ca(2+)] to produce shifts in activation of Slo1 conductance can, therefore, be accounted for by multiple types of divalent cation binding sites.  相似文献   

14.
Calbindin D28k exhibits properties characteristic of a Ca2+ sensor   总被引:3,自引:0,他引:3  
Calbindin D(28k) is a member of the calmodulin superfamily of Ca(2+)-binding proteins and contains six EF-hands. The protein is generally believed to function as a Ca(2+) buffer, but the studies presented in this work indicate that it may also act as a Ca(2+) sensor. The results show that Mg(2+) binds to the same sites as Ca(2+) with an association constant of approximately 1.4.10(3) m(-1) in 0.15 m KCl. The four high affinity sites in calbindin D(28k) bind Ca(2+) in a non-sequential, parallel manner. In the presence of physiological concentrations of Mg(2+), the Ca(2+) affinity is reduced by a factor of 2, and the cooperativity, which otherwise is modest, increases. Based on the binding constants determined in the presence of physiological salt concentrations, we estimate that at the Ca(2+) concentration in a resting cell calbindin D(28k) is saturated to 40-75% with Mg(2+) but to less than 9% with Ca(2+). In contrast, the protein is expected to be nearly fully saturated with Ca(2+) at the Ca(2+) level of an activated cell. A substantial conformational change is observed upon Ca(2+) binding, but only minor structural changes take place upon Mg(2+) binding. This suggests that calbindin D(28k) undergoes Ca(2+)-induced structural changes upon Ca(2+) activation of a cell. Thus, calbindin D(28k) displays several properties that would be expected for a protein involved in Ca(2+)-induced signal transmission and hence may function not only as a Ca(2+) buffer but also as a Ca(2+) sensor. Digestion patterns resulting from limited proteolysis of the protein suggest that the loop of EF-hand 2, a variant site that does not bind Ca(2+), becomes exposed upon Ca(2+) binding.  相似文献   

15.
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   

16.
Many diseases such as cardiac arrhythmia, diabetes, and chronic alcoholism are associated with a marked decrease of plasma and parenchymal Mg(2+), and Mg(2+) administration is routinely used therapeutically. This study uses isolated rat hepatocytes to ascertain if and under which conditions increases in extracellular Mg(2+) result in an increase in intracellular Mg(2+). In the absence of stimulation, changing extracellular Mg(2+) had no effect on total cellular Mg(2+) content. By contrast, carbachol or vasopressin administration promoted an accumulation of Mg(2+) that increased cellular Mg(2+) content by 13.2 and 11.8%, respectively, and stimulated Mg(2+) uptake was unaffected by the absence of extracellular Ca(2+). Mg(2+) efflux resulting from stimulation of alpha- or beta-adrenergic receptors operated with a Mg(2+):Ca(2+) exchange ratio of 1. These data indicate that cellular Mg(2+) uptake can occur rapidly and in large amounts, through a process distinct from Mg(2+) release, but operating only upon specific hormonal stimulation.  相似文献   

17.
Ca2+ binding to rabbit skeletal calsequestrin was studied at physiological ionic strength by equilibrium flow dialysis, Hummel-Dryer gel filtration and microcalorimetry. 31 Ca(2+)-binding sites with a mean dissociation constant (KD) of 0.79 mM were titrated in the absence, and 23 sites with a KD of 0.88 mM in the presence of 3 mM Mg2+. No cooperativity was observed. For Mg2+ binding, the combination of gel filtration and microcalorimetry yielded a stoichiometry of 26 Mg2+/protein with a KD of 2mM. 1 mM Ca2+ decreased the stoichiometry to 20 Mg2+/protein. Binding of Ca2+ in the absence and presence of 3 mM Mg2+ was accompanied by a release of 2.0 and 2.7 H+/protein, respectively. Mg2+ binding did not lead to a significant proton release suggesting a qualitative difference in the Ca(2+)- and Mg(2+)-binding sites. After correction for proton release, the enthalpy change for Ca2+ binding was very low (-1.5 kJ/protein in the absence, and -15 kJ/protein in the presence of 3 mM Mg2+). The entropy change (+59 J/K.site in the absence and +56 J/K.site in the presence of Mg2+) was therefore virtually the sole driving force for Ca2+ binding. Mg2+ binding is slightly more exothermic (-12.6 kJ/protein), but as for Ca2+, the entropy change (+50 J/K.site) constituted the major driving force of the reaction. A fluorimetric study indicates that the conformation of tryptophan in Mg(2+)-saturated calsequestrin was clearly different from that in the Ca(2+)-saturated protein, but that the (Ca2+ + Mg2+)-saturated protein was not distinct from the Ca(2+)-saturated protein. Thus, in addition to the thermodynamic characterization of the Ca2+/calsequestrin interaction, our data indicate that Ca2+ and Mg2+ do not bind to the same sites on calsequestrin. The data also predict considerable proton fluxes upon Ca(2+)-Mg2+ exchange in vivo.  相似文献   

18.
We have used the highly selective alpha(4)beta(1) inhibitor 2S-[(1-benzenesulfonyl-pyrrolidine-2S-carbonyl)-amino]-4-[4-methyl-2S-(methyl-[2-[4-(3-o-tolyl-ureido)-phenyl]-acetyl]-amino)-pentanoylamino]-butyric acid (BIO7662) as a model ligand to study alpha(4)beta(1) integrin-ligand interactions on Jurkat cells. Binding of [(35)S]BIO7662 to Jurkat cells was dependent on the presence of divalent cations and could be blocked by treatment with an excess of unlabeled inhibitor or with EDTA. K(D) values for the binding of BIO7662 to Mn(2+)-activated alpha(4)beta(1) and to the nonactivated state of the integrin that exists in 1 mm Mg(2+), 1 mm Ca(2+) were <10 pm, indicating that it has a high affinity for both activated and nonactivated integrin. No binding was observed on alpha(4)beta(1) negative cells. Through an analysis of the metal ion dependences of ligand binding, several unexpected findings about alpha(4)beta(1) function were made. First, we observed that Ca(2+) binding to alpha(4)beta(1) was stimulated by the addition of BIO7662. From solution binding studies on purified alpha(4)beta(1), two types of Ca(2+)-binding sites were identified, one dependent upon and the other independent of BIO7662 binding. Second, we observed that the metal ion dependence of ligand binding was affected by the affinity of the ligand for alpha(4)beta(1). ED(50) values for the metal ion dependence of the binding of BIO7762 and the binding of a lower affinity ligand, BIO1211, differed by 2-fold for Mn(2+), 30-fold for Mg(2+), and >1000-fold for Ca(2+). Low Ca(2+) (ED(50) = 5-10 microm) stimulated the binding of BIO7662 to alpha(4)beta(1). The effects of microm Ca(2+) closely resembled the effects of Mn(2+) on alpha(4)beta(1) function. Third, we observed that the rate of BIO7662 binding was dependent on the metal ion concentration and that the ED(50) for the metal ion dependence of BIO7662 binding was affected by the concentration of the BIO7662. These studies point to an even more complex interplay between metal ion and ligand binding than previously appreciated and provide evidence for a three-component coupled equilibrium model for metal ion-dependent binding of ligands to alpha(4)beta(1).  相似文献   

19.
The effects of three mutations on the EF-hand Ca(2+)/Mg(2+) binding site of smooth muscle myosin regulatory light chain (RLC) were studied: D5S, in which an aspartate is replaced by a serine in position 5 of the loop; D9E, in which an aspartate is replaced by a glutamate in position 9; and D12E, in which the aspartate in position 12 is replaced by a glutamate. All possible combinations of the three mutations were produced. The single mutants D5S and D9E and the double mutant D5S/D9E have low affinity for Ca(2+). All the mutants containing mutation D12E are Ca(2+)-specific and have higher affinities than wild type, even when containing mutations D5S or D9E. All of the mutants studied have lower affinity for Mg(2+) than the wild-type protein. As expected, the changes in binding free energy that each mutant produces depend on the residues present at the other positions of the site, since the mutated positions are very close in the protein structure. Coupling energies are about the same for all pairs of mutants when binding Ca(2+), but can have different values when binding Mg(2+). D5S and D9E have a large negative coupling energy for Mg(2+) binding which suggests an interaction between these two positions. When mutation D12E is present, the coupling energy for Mg(2+) binding between D5S and D9E is much lower, suggesting that this interaction occurs only if an aspartate is in position 12. Glutamate in position 9 may be able to coordinate Mg(2+) directly in the double mutant D5S/D9E.  相似文献   

20.
Ryanodine, a plant alkaloid, is one of the most widely used pharmacological probes for intracellular Ca(2+) signaling in a variety of muscle and non-muscle cells. Upon binding to the Ca(2+) release channel (ryanodine receptor), ryanodine causes two major changes in the channel: a reduction in single-channel conductance and a marked increase in open probability. The molecular mechanisms underlying these alterations are not well understood. In the present study, we investigated the gating behavior and Ca(2+) dependence of the wild type (wt) and a mutant cardiac ryanodine receptor (RyR2) after being modified by ryanodine. Single-channel studies revealed that the ryanodine-modified wt RyR2 channel was sensitive to inhibition by Mg(2+) and to activation by caffeine and ATP. In the presence of Mg(2+), the ryanodine-modified single wt RyR2 channel displayed a sigmoidal Ca(2+) dependence with an EC(50) value of 110 nm, whereas the ryanodine-unmodified single wt channel exhibited an EC(50) of 120 microm for Ca(2+) activation, indicating that ryanodine is able to increase the sensitivity of the wt RyR2 channel to Ca(2+) activation by approximately 1,000-fold. Furthermore, ryanodine is able to restore Ca(2+) activation and ligand response of the E3987A mutant RyR2 channel that has been shown to exhibit approximately 1,000-fold reduction in Ca(2+) sensitivity to activation. The E3987A mutation, however, affects neither [(3)H]ryanodine binding to, nor the stimulatory and inhibitory effects of ryanodine on, the RyR2 channel. These results demonstrate that ryanodine does not "lock" the RyR channel into an open state as generally believed; rather, it sensitizes dramatically the channel to activation by Ca(2+).  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