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 46 毫秒
1.
Washing spinach PSII oxygen-evolution complex (OEC) with 2 mmol/L EGTA or extraction medium caused a 28.4% and 25.0% loss of oxygen evolution activities respectively, but the loss of polypeptide components of OEC did not take place, whereas washing with 1 mol/L NaCI caused both a 90.0% loss of oxygen evolution activity and loss of 17, 23kD polypeptides. Adding 5–10 mmol/L CaC12 could restore oxygen evolution activities of OEC by various washing to a great extent, but had no effect on control OEC, whereas adding 5–10 mmol/L EGTA had no effect on the OEC by various' washing, but caused the loss of oxygen evolution mixtures, which could induce the release of of 17, 23kD polypeptides from OEC, caused 54.3% loss of oxygen evolution activity, under this circumstance, adding 2 mmol/L of EGTA could only maintain a weak oxygen evolution activity of OEC, but adding 10 mmol/L of CaCl2 could restore oxygen evolution activity of OEC to the control level. These findings' suggest a two way loose binding of Ga2+ to PSⅡ OEC in one way Ca2+ is loose bound to the surface of PSⅡOEC and in other, the Ca2+-binding site is wrapped by 17, 23kD polypeptides. Both of them have effect on oxygen evolution activity of PSⅡ OEC. By way, Mn2+ can antagonize the restoration of oxygen evolution activity by Ca2+ to the NaCl-washing PSⅡ OEC.  相似文献   

2.
A 5 min exposure of photosystem II to a pH 3 citric acid solution is a simple method for selective removal of Ca(2+) from the O(2)-evolving complex. The resulting preparation retains the 23 and 17 kDa extrinsic polypeptides, but the activity of this material is only 10-20% of that of an untreated control sample. Biochemical characterization of citrate-treated photosystem II reveals that some reaction centers lose the extrinsic proteins during citrate treatment. Furthermore, a comparison of photosystem II preparations treated with citrate, or depleted of 23 and 17 kDa extrinsic polypeptides by high-salt treatment, shows that low concentrations of a small reductant, NH(2)OH, which has little effect on the activity of intact photosystem II, can reduce and inhibit the Mn cluster in both types of preparations. In contrast, a large reductant, hydroquinone, cannot access the majority of O(2)-evolving centers in citrate-treated preparations, while 23 and 17 kDa-depleted material is rapidly inactivated by the reductant. Incubation of the citrate-treated samples in high ( approximately 60 mM) concentrations of CaCl(2) restores 50% of the lost activity; this Ca(2+)-reconstituted activity is chelator-insensitive, indicating that rebinding of Ca(2+) restores the structural integrity of the O(2)-evolving complex. A characterization of Ca(2+) and Cl(-) affinities in steady-state activity assays shows that citrate-treated preparations exhibit a Cl(-) requirement similar to that of polypeptide-depleted photosystem II, while Ca(2+) reactivation of O(2) evolution appears to occur at two structurally distinct sites. One site exhibits a high Ca(2+) affinity, similar to that found in polypeptide-depleted samples, but a second, lower-affinity site also exists, with a K(M) that is approximately 10 times greater than that of the high-affinity site, which is associated with centers that retain the extrinsic polypeptides. These data indicate that citrate-induced Ca(2+) depletion causes release of the 23 and 17 kDa extrinsic polypeptides from some photosystem II reaction centers, and also modifies the structure of the polypeptide-retaining O(2)-evolving centers so that the Mn cluster is exposed to small, but not large, reductants. This change may be due to subtle modifications to the structure of the photosystem II extrinsic proteins that produces a new pathway between the solvent and the Mn cluster or, alternatively, to the opening of an existing channel in the intrinsic lumenal polypeptide domain, between the solvent and the Mn cluster, that is normally occluded by a bound Ca(2+) atom.  相似文献   

3.
Calcium is an essential cofactor in the oxygen-evolving complex (OEC) of photosystem II (PSII). The removal of Ca2+ or its substitution by any metal ion except Sr2+ inhibits oxygen evolution. We used steady-state enzyme kinetics to measure the rate of O2 evolution in PSII samples treated with an extensive series of mono-, di-, and trivalent metal ions in order to determine the basis for the affinity of metal ions for the Ca2+-binding site. Our results show that the Ca2+-binding site in PSII behaves very similarly to the Ca2+-binding sites in other proteins, and we discuss the implications this has for the structure of the site in PSII. Activity measurements as a function of time show that the binding site achieves equilibrium in 4 h for all of the PSII samples investigated. The binding affinities of the metal ions are modulated by the 17 and 23 kDa extrinsic polypeptides; their removal decreases the free energy of binding of the metal ions by 2.5 kcal/mol, but does not significantly change the time required to reach equilibrium. Monovalent ions are effectively excluded from the Ca2+-binding site, exhibiting no inhibition of O2 evolution. Di- and trivalent metal ions with ionic radii similar to that of Ca2+ (0.99 A) bind competitively with Ca2+ and have the highest binding affinity, while smaller metal ions bind more weakly and much larger ones do not bind competitively. This is consistent with a size-selective Ca2+-binding site that has a rigid array of coordinating ligands. Despite the large number of metal ions that competitively replace Ca2+ in the OEC, only Sr2+ is capable of partially restoring activity. Comparing the physical characteristics of the metal ions studied, we identify the pK(a) of the aqua ion as the factor that determines the functional competence of the metal ion. This suggests that Ca2+ is directly involved in the chemistry of water oxidation and is not only a structural cofactor in the OEC. We propose that the role of Ca2+ is to act as a Lewis acid, binding a substrate water molecule and tuning its reactivity.  相似文献   

4.
Absorbance and fluorescence changes of oxacarbocyanine dyes during ATP-induced Ca2+ transport in rabbit sarcoplasmic reticulum were analyzed. The response of the probes is complex and contains contributions from the binding of Ca2+ and ATP to the membrane. In a medium of 0.12 M KCl and 5 mM MgCl2, the fluorescence of Di-O-C5(3) is decreased by Ca2+ or ATP with apparent dissociation constants of 0.2 and 5 micron, respectively. This suggests that oxacarbocyanines respond to binding of Ca2+ and ATP at the active site of Ca2+ transport ATPase. The effect of ATP is observed in the absence of divalent cations. Further changes in the fluorescence or absorbance of cyanine dyes occur at millimolar concentrations of Ca2+ or during ATP-induced Ca2+ uptake, which can be related to Ca2+ binding to low affinity, relatively nonspecific binding sites on the membrane, that can also bind K+ and Mg2+. The optical changes due to Ca2+ accumulation are most pronounced in media of 0.25 M sucrose and much reduced in 0.12 M KCl and 5 mM MgCl2, in accord with competition by K+ and Mg2+ for the low affinity Ca2+ binding sites. These effects must be taken into account in the evaluation of the magnitude and direction of membrane potential in sarcoplasmic reticulum vesicles during Ca2+ uptake and release.  相似文献   

5.
Dependence of the red blood cell calcium pump on the membrane potential   总被引:4,自引:0,他引:4  
(1) It is shown that the rate of calcium extrusion from intact human red cells is faster at a membrane potential of approximately +50 mV (inside) than at approximately -50 mV. (2) The positive potential applied was the chloride potential of KCl cells in a K-gluconate medium when the Ca2+ sensitive K+ channel was blocked by 0.3mM quinidine. The negative potential resulted from the high K+ permeability in Ca2+ loaded cells (the cells were loaded to a Ca2+ activity in the cell water of about 50 microM). (3) It is further demonstrated that the Ca2+ affinity of the pump ATPase is decreased both at the internal (high affinity) and external (low affinity) site by increasing the proton concentration. Acidification thus inhibits internally and stimulates externally. (4) An indirect effect of the membrane potential on the pump activity via the accompanying pH shifts on either side of the membrane could be ruled out by choosing Ca2+ concentrations which are fully activating at the internal Ca2+ binding site at pH 6.5 and not yet inhibitory at the external Ca2+ binding site at pH 8. (5) The result is compatible with the assumption that the human red cell Ca-pump is exchanging Ca2+ for protons, yet is electrogenic by virtue of a stoichiometry of 1H+:1Ca2+ for this exchange.  相似文献   

6.
Lowering the extracellular K+ content from 6 to 0.6 mM causes a rise, and elevation from 6 to 8.5 mM a fall of 45Ca++ efflux from the vascular smooth muscle cells of the arteria carotis communis of cattle. In contrast, a level of 17 mM K+ has no influence. Removal of extracellular calcium does not block these effects. 10(-4) M ouabain also induces a rise in Ca++ efflux, additional potassium reduction then being without effect; 10(-9) M ouabain is of no influence. The 45Ca++ efflux kinetics correlates with the activity of the isolated Na,K-ATPase. Tonus increases of the vascular strips by 10(-4) M ouabain and potassium deficiency cannot be blocked by 4 mM lanthanum or removal of extracellular calcium. Unlike sodium, potassium stimulates the active Ca++ binding and the activity of the Ca-ATPase of the microsomal fraction. The ative Ca++ binding of the mitochondria is stimulated by both ions. It is postulated that the activity of the plasma membrane Na,K-pump is able to regulate the tonus of big arteries through alteration of Ca++ storage processes.  相似文献   

7.
Bovine cytochrome c oxidase in 0.1% dodecylmaltoside, 50 mM KCl and 10 mM Tris-HCl, pH 7.4 is monodisperse with an apparent Mr 360,000 (dimer) as estimated by filtration on Ultrogel AcA 34. In the absence of added KCl the apparent Mr is 160,000 (monomer). The dimeric enzyme has a high and a low affinity site for cytochrome c; the monomeric, only the high affinity site. The results are consistent with the existence of one active site per monomer, having high affinity for cytochrome c. Since in a dimer the two sites are in close proximity, the binding of the first molecule of cytochrome c to the first site hinders the binding of the second molecule to the second site. The kinetic data fit with a model of homotropic negative cooperativity. The effect of salts on the cytochrome c oxidase kinetics is also present in isolated bovine heart mitochondria.  相似文献   

8.
We have studied the calcium-binding properties of two high affinity calcium-binding proteins from squid optic lobes: one, squid calmodulin (SCaM), similar to bovine brain calmodulin (BCaM), the other, squid calcium-binding protein (SCaBP), distinct (Head, J.F., Spielberg, S., and Kaminer, B. (1983) Biochem J. 209, 797-802). Equilibrium dialysis measurements on the squid proteins (and BCaM) were made at 100 mM KCl in the presence and absence of 3 mM Mg2+, and at 400 mM KCl in the presence of 3 mM Mg2+, which more closely resembles the conditions in the squid. SCaM, SCaBP, and BCaM each bind a maximum of 4 Ca2+ ions/molecule of protein under the ionic conditions tested. SCaBP has a higher affinity than SCaM or BCaM for Ca2+ at 100 mM KCl in the absence of Mg2+. However, in the presence of Mg2+, half-maximal binding to SCaBP occurs at a similar pCa value to that observed with calmodulin. Increasing the KCl concentration reduces the affinity of all three proteins for Ca2+. UV absorption measurements showed that the binding of 4 Ca2+ ions/molecule is necessary to complete spectral changes in SCaBP, compared to two for the calmodulins. While Ca2+ causes perturbations in aromatic chromophores in SCaM and SCaBP, Mg2+ causes a significant perturbation only in SCaBP. These Mg2+-induced changes differ qualitatively from those induced by Ca2+.  相似文献   

9.
Receptor fractions were prepared from follicle-rich ovaries (for FSH), luteal cell-rich ovaries (for LH and PRL), and adrenals (for PRL) of rats. Divalent metal ions, Mg++, Ca++, and Mn++ showed inhibitory effects on the binding of LH and FSH to their receptors. The binding of the former was more sensitive to these ions than the latter. On the other hand they showed bell-shaped promotive effects on PRL-ovarian receptor binding, the maximal effects being observed at 10-20 mM. Besides these ions, Ba++ also had a promotive effect, while other divalent metal ions such as Zn++, Cd++, Ni++, and Co++ showed inhibitory effects on PRL-ovarian receptor binding at 5 mM. Mg++ and Ca++ also promoted PRL-adrenal receptor binding, while Mn++ promoted the binding at 10 mM but inhibited it at higher concentrations. Association constant (Ka) and binding capacity (Bmax) of PRL receptors of the ovary and the adrenal were significantly different (ovary: Ka = 0.69 X 10(10) M-1, Bmax = 62 fmol/mg protein, adrenal: Ka = 0.21 X 10(10) M-1, Bmax = 99 fmol/mg protein). Ka of the ovarian PRL receptor was not influenced by these divalent ions, while that of the adrenal receptor was doubled by Ca and Mn ions, Bmax of the latter was also increased. A cooperative effect of Mg and Ca ions was observed on Ka and Bmax of the adrenal receptor. The sizes of the PRL binding sites of these organs revealed by affinity labelling were 17K and 40K in the ovary, and 40K and 110K in the adrenal. These results indicate the different properties of receptors in these different target organs.  相似文献   

10.
Centrifuge transport, equilibrium dialysis, and electron paramagnetic resonance studies on the binding of Mn2+ to myosin revealed two sets of noninteracting binding sites which are characterized at low ionic strength (0.016 M KCl) by affinity constants of 10(6) M-1 (Class I) and 10(3) M-1 (Class II), respectively. At 0.6 M KCl concentration, the affinity of Mn2+ for both sets of sites is reduced. The maximum number of binding sites is 2 for the high affinity and 20 to 25 for the low affinity set. Other divalent metal ions displace Mn2+ from the high affinity sites in the following order of effectiveness: Ca greater than Mg = Zn = Co greater than Sr greater than Ni. The inhibitory effects of Mg2+ and Ca2+ upon the Mn2+ binding are competitive with inhibitor constants of 0.75 to 1 mM which is similar to that of the low affinity divalent metal ion binding sites. Exposure of myosin to 37 degrees partially inhibits Mn2+ binding to Class I parallel with inhibition of ATPase activity. The binding of Mn2+ to the high affinity binding sites is not significantly influenced by ADP or PPi, although Mn2+ increases the affinity of ADP binding to myosin at high ionic strength.  相似文献   

11.
The roles of Ca(2+) in H(2)O oxidation may be as a site of substrate binding, and as a structural component of the photosystem II O(2)-evolving complex. One indication of this dual role of the metal is revealed by probing the Mn cluster in the Ca(2+) depleted O(2) evolving complex that retains extrinsic 23- and 17-kDa polypeptides with reductants (NH(2)OH and hydroquinone) [Biochemistry 41 (2002) 958]. Calcium appears to bind to photosystem II at a site where it could bind substrate H(2)O. Equilibration of Ca(2+) with this binding site is facilitated by increased ionic strength, and incubation of Ca(2+) reconstitution mixtures at 22 degrees C accelerates equilibration of Ca(2+) with the site. The Ca(2+) reconstituted enzyme system regains properties of unperturbed photosystem II: Sensitivity to NH(2)OH inhibition is decreased, and Cl(-) binding with increased affinity can be detected. The ability of ionic strength and temperature to facilitate rebinding of Ca(2+) to the intact O(2) evolving complex suggests that the structural environment of the oxidizing side of photosystem II may be flexible, rather than rigid.  相似文献   

12.
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   

13.
Calcium binding to troponin C and troponin was examined by a metallochromic indicator method under various conditions to obtain a further understanding of the regulatory roles of these proteins in muscle contraction. Troponin C has four Ca binding sites, of which 2 sites have a high affinity of 4.5 X 10(6) M-1 for Ca2+ and the other 2 sites have a low affinity of 6.4 X 10(4) M-1 in a reaction medium consisting of 100 mM KCl, 20 mM MOPS-KOH pH 6.80 and 0.13 mM tetramethylmurexide at 20 degrees C. Magnesium also binds competitively to both the high and low affinity sites: the apparent binding constants are 1,000 M-1 and 520 M-1, respectively. Contrary to the claim by Potter and Gergely (J. Biol. Chem. 250, 4628-4633, 1975), the low affinity sites are not specific only for Ca2+. The high and low affinity sites of troponin C showed different dependence on the ionic strength: the high affinity sites were similar to GEDTA, while the low affinity sites were similar to calmodulin, which has a steeper ionic strength dependence than GEDTA. Ca binding to troponin C was not affected by change of pH between 6.5 and 7.2. Troponin I enhanced the apparent affinity of troponin C for Ca2+ to a value similar to that for troponin. Trifluoperazine also increased Ca binding to troponin C. Troponin has four Ca binding sites as does troponin C, but the affinities are so high that the precise analysis was difficult by this method. The apparent binding constants for Ca2+ and Mg2+ were determined to be 3.5 X 10(6) M-1 and 440 M-1, respectively, for low affinity sites under the same conditions as for troponin C, being independent of change in pH between 6.5 and 7.2. The competitive binding of Mg2+ to the low affinity sites of troponin is consistent with the results of Kohama (J. Biochem. 88, 591-599, 1980). The estimate for the high affinity sites is compatible with the reported results.  相似文献   

14.
We have established a new binding assay in which 125I-labeled synaptic vesicles are incubated with brain spectrin covalently immobilized on cellulosic membranes in a microfiltration apparatus. We obtained saturable, high affinity, salt- (optimum at 50-70 mM NaCl) and pH- (optimum at pH 7.5-7.8) dependent binding. Nonlinear regression analysis of the binding isotherm indicated one site binding with a Kd = 59 micrograms/ml and a maximal binding capacity = 1.9 micrograms vesicle protein per microgram spectrin. The fact that the binding of spectrin was via synapsin was demonstrated in three ways. (a) Binding of synaptic vesicles to immobilized spectrin was eliminated by prior extraction with 1 M KCl. When the peripheral membrane proteins in the 1 M KCl extract were separated by SDS-PAGE, transferred to nitrocellulose paper and incubated with 125I-brain spectrin, 96% of the total radioactivity was associated with five polypeptides of 80, 75, 69, 64, and 40 kD. All five polypeptides reacted with an anti-synapsin I polyclonal antibody, and the 80- and 75-kD polypeptides comigrated with authentic synapsin Ia and synapsin Ib. The 69- and 64-kD polypeptides are either proteolytic fragments of synapsin I or represent synapsin IIa and synapsin IIb. (b) Pure synapsin I was capable of competitively inhibiting the binding of radioiodinated synaptic vesicles to immobilized brain spectrin with a Kl = 46 nM. (c) Fab fragments of anti-synapsin I were capable of inhibiting the binding of radioiodinated synaptic vesicles to immobilized brain spectrin. These three observations clearly establish that synapsin I is a primary receptor for brain spectrin on the cytoplasmic surface of the synaptic vesicle membrane.  相似文献   

15.
Two kinds of ATP binding sites were found on the ATPase molecule in deoxycholic acid-treated sarcoplasmic reticulum. One was the catalytic site (1 mol/mol active site) and its affinity was high. Upon addition of Ca2+, all the ATP bound to the catalytic site disappeared at 75 mM KCl, while a significant amount of ATP remained bound to the site at 0–2 mM KCl. The latter binding was found to be due to the formation of a slowly exchanging enzyme-ATP complex, which is in equilibrium with phosphoenzyme + ADP. The other binding site was the regulatory one (1 mol/mol active site) and its affinity was low, changing only insignificantly upon addition of Ca2+. The ATP binding to the regulatory site shifted the equilibrium between the slowly exchanging complex and EP toward EP.  相似文献   

16.
M Miyao  Y Inoue 《Biochemistry》1991,30(22):5379-5387
The Mn cluster that catalyzes photosynthetic oxygen evolution was removed from the photosystem II (PSII) complex by treating PSII membranes with 1.0 mM NH2OH with concomitant inactivation of oxygen evolution. The cluster was reconstituted by incubating the treated membranes with 1.0 mM Mn2+, 20 mM Ca2+, 10 microM 2,6-dichlorophenolindophenol, and Cl- under illumination with continuous or flashing light to restore the oxygen-evolving capacity. This light-dependent activation (photoactivation) of oxygen evolution did not occur to a significant extent at 3 mM Cl-, but markedly accelerated at higher Cl- concentrations without showing a saturation phenomenon even at 1 M Cl-. At 10 mM Cl- only about 10% of the oxygen-evolving activity before NH2OH treatment was restored by 5-min illumination with continuous light, whereas at 600 mM Cl- about 60% of the original activity was recovered. This acceleration resulted from at least two different actions of Cl-: (1) stabilization of the intermediate state involved in the photoactivation process and (2) increase in the quantum yield of photoactivation. The stabilization of the intermediate was saturated at about 150 mM Cl-, whereas the increase in yield did not show saturation. The Cl(-)-induced increase in quantum yield did not involve any changes in the affinity of either Mn2+ binding or Ca2+ binding for photoactivation, but was rather ascribed to a protective effect of Cl- against inhibition of photoactivation by high concentrations of Mn2+. We also found that removal of the extrinsic 33-kDa protein from the PSII complex increased the Cl- requirement for photoactivation.  相似文献   

17.
T Mine  S Kimura  H Osawa  E Ogata 《Life sciences》1986,38(25):2285-2292
Cobalt ions (2 mM) inhibited the glycogenolysis induced by phenylephrine and glucagon in perfused rat liver. Cobalt ions also inhibited 45Ca++ efflux from prelabelled livers induced by phenylephrine and glucagon. In addition, they inhibited the rise in tissue levels of cyclic AMP caused by glucagon, but did not inhibit the stimulation of 45Ca++ efflux or glycogenolysis by cyclic AMP or dibutyryl cyclic AMP. The specific binding of glucagon and alpha-agonist to hepatocytes was not inhibited by cobalt ions. These data suggest that cobalt ions, presumably through their high affinity for calcium binding sites on membranes inhibit the stimulation of glycogenolysis by phenylephrine and glucagon in distinct ways; one by inhibiting calcium mobilization and the other by inhibiting cyclic AMP production. Therefore, it is conceivable that membrane-bound calcium plays an important role in stimulating Ca++ mobilization by phenylephrine, and cyclic AMP production by glucagon.  相似文献   

18.
A Abbott  W J Ball 《Biochemistry》1992,31(45):11236-11243
Monoclonal antibody M7-PB-E9 binds the sheep kidney Na+,K(+)-ATPase alpha-subunit with high affinity (Kd = 3 nM) and inhibits enzyme turnover in competition with ATP, and, like ATP, in the presence of Mg2+, it stimulates the rate of ouabain binding [Ball, W. J. (1984) Biochemistry 23, 2275-2281]. In this study, covalent attachment of fluorescein 5'-isothiocyanate (FITC) at (or near) the enzyme's ATP binding site did not alter the antibody's affinity for alpha nor did bound antibody alter the anisotropy of (r = 0.36) or the solvent accessibility of iodide to bound FITC. Further, in its E1Na+ conformation (4 mM NaCl), the enzyme's affinity for the ATP congener eosin was unaltered by the bound antibody (Kd = 9 nM). In contrast, partial E2 conformations induced by KCl lowered eosin affinities (0.2 mM KCl, Kd = 28 nM; 0.4 mM, Kd = 86 nM), and M7-PB-E9 reduced these affinities further (Kd = 66 and 130 nM, respectively). By monitoring the fluorescence changes of the FITC-labeled enzyme, the antibody was found to assist several ligand-induced conformational transitions from E1 (E1Na+ or E1Tris) to E2 (E2K+, E2-P(i)Mg2+, or E2Mg2+.ouabain) states, and inhibit the E2K(+)-->E1Na+ transition. Antibody binding alone, however, did not appear to significantly alter enzyme conformation. The antibody therefore is not directed against the ATP site but binds to a region of alpha distinct from any ligand binding site and which plays an important role in the E1<-->E2 transitions.  相似文献   

19.
Ca2+ binding by Myxicola neurofilament proteins   总被引:1,自引:0,他引:1  
Titrimetric, 45Ca dialysis, and autoradiographic methods were used to examine how axoplasmic proteins from the giant neuron of the marine annelid Myxicola infundibulum bind calcium. Following the autoradiographic method of Maruyama et al., the 150-160 kD neurofilament subunits were identified as prominent intracellular Ca-binding peptides. Using equilibrium dialysis, extracts of axoplasmic proteins (greater than 50% neurofilament subunits) were examined in 300 mM KCl at different concentrations of free Ca and Mg, and at different pH. Axoplasmic proteins showed a high affinity Ca binding site (K1/2 3-6 microM, capacity 3-7 mumole g-1 protein) at pH 6.8 or pH 7.5. Changing the Mg concentration from 0 to 5 mM had no effect on the Ca binding. Elevating the dialysis pH from 7.0 to 9.0 reduced the apparent number of binding sites for Ca. Using microelectrodes to record the free Ca, microtitrations of axoplasmic proteins were completed by adding small amounts of CaCl2 to 100 microliters volumes of protein solutions. In a medium containing ionic constituents closely resembling those of the Myxicola axon, a Ca binding capacity of 5.0 mumole g-1 protein and a K1/2 of approximately 1 microM were measured.  相似文献   

20.
Photosystem II enriched membranes were depleted of Ca2+ and the 17- and 23-kDa polypeptides by treatment with NaCl and EGTA. The 17- and 23-kDa polypeptides were then reconstituted. This preparation was incapable of O2 evolution until Ca2+ was added. An EPR study revealed the presence of two new EPR signals. One of these is a modified S2 multiline signal with an isotropic g value of 1.96 with at least 26 hyperfine peaks (average spacing 55 G) distributed over approximately 1600 G. The other is a near-Gaussian signal with an isotropic g value of 2.004, which is attributed to a formal S3 state. Experiments involving the interconversion of these signals and the effect of Ca2+ and Sr2+ rebinding provide evidence for these assignments. From these results the following conclusions are drawn: (1) These results are consistent with our earlier demonstration that charge accumulation is blocked after formation of S3 when Ca2+ is deficient. (2) Binding of the 17- and 23-kDa polypeptides to photosystem II in the absence of Ca2+ results in the perturbation of the Mn cluster. This is taken as a further indication that the Ca2+-binding site is close to or even an integral part of the Mn cluster. (3) The S3 signal may arise from an organic free radical interacting magnetically with the Mn cluster. However, other possible origins for this signal, including the Mn cluster itself, must also be considered.  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