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.
Z W Yang  C F Kong  J A Babitch 《Biochemistry》1988,27(18):7045-7050
In our previous work [Yang, Z. W., & Babitch, J. A. (1988) Biochemistry (preceding paper in this issue)] divalent cations were found to be more effective promoters of astroglial filament formation than were monovalent cations. To determine if one or more divalent cation binding sites were the basis for this difference, glial fibrillary acidic protein (GFAP) was attached to nitrocellulose membranes and bathed in 1 microM 45CaCl2 in 60 mM KCl, 0.5 mM MgCl2, and 10 mM imidazole hydrochloride, pH 7.4. After removal of unbound 45Ca2+, GFAP was observed to bind calcium. Flow dialysis experiments showed that GFAP, dissolved in 2 mM Tris-HCl, pH 7.5, contained three classes of binding sites and 0.61 +/- 0.08 (SD), 1.7 +/- 0.4, and 4.6 +/- 0.2 sites per GFAP molecule with dissociation constants of 0.66 +/- 0.01 microM, 6.6 +/- 0.3 microM, and 44 +/- 1 microM, respectively. After addition of 0.5 mM MgSO4 to the flow dialysis solution, the high- and low-affinity sites were not observed while the remaining sites (1.95 +/- 0.15 per GFAP molecule) had a Kd = 2.16 +/- 0.25 microM. This showed that the high- and low-affinity sites are "Ca2+-Mg2+" sites while sites with intermediate affinity are calcium specific. To locate the calcium-binding regions, GFAP peptides were examined for calcium binding by calcium-45 autoradiography. The calcium-specific binding areas were localized in coil I. Computer-assisted analysis of the GFAP sequence revealed several EF-hand-like areas which could be the calcium binding sites. We conclude that divalent cations may play both structural and regulatory roles in astroglial intermediate filaments.  相似文献   

2.
Numerous Ca2+-binding sites were localized ultrahistochemically at the inner surface of the erythrocyte membrane and small amounts at the outer surface. La3+-binding sites were demonstrated at the outer surface only. The results were discussed in relationship to the binding capacity of the filamentous matrix and the glycocalyx of the human erythrocyte membrane.  相似文献   

3.
Recently, Mills and Johnson [7] and our group [9] provided evidence that calmodulin contains, in addition to the four Ca2+-binding sites (capital sites), which are essential for drug- and enzyme-binding, a number of divalent cation-binding sites of different ion selectivity (auxiliary sites), which modulate drug-binding as well as the affinity of Ca2+ for the capital sites. In the present study, the number of auxiliary sites and their relationship to the capital sites were determined by equilibrium gel filtration and by flow microcalorimetry with Zn2+ and Mn2+ as selective probes for the auxiliary sites and with Cd2+ as a probe for both types of sites. In the absence of other divalent cations, 6 mol of Zn2+ bind to calmodulin with an identical affinity constant of 2,850 M-1 and a delta H0 of 106 kJ/mol calmodulin. In the presence of millimolar free Ca2+ calmodulin binds, in addition to four Ca2+, six Zn2+ with an affinity constant of 1,200 M-1 and a delta H0 of 47 kJ/mol calmodulin. The Zn2+-Ca2+ antagonism is governed by negative free energy coupling between the capital and auxiliary sites. In contrast, the Zn2+-Mg2+ antagonism follows the rule of straight competition at all six auxiliary sites. Mn2+ also binds exclusively to the auxiliary sites with affinity constants of 800 or 280 M-1 and delta H0 of 45 or 46 kJ/mol calmodulin in the absence and presence of saturating [Ca2+], respectively. Cd2+ binds to the capital sites with an affinity constant of 3.4 10(4) M-1 (delta H = 35 kJ/mol calmodulin) and to the auxiliary sites with ca. 100-fold lower affinity. The Zn2+ much greater than Mn2+ greater than or equal to Cd2+ greater than Mg2+ selectivity of the auxiliary sites corroborates the potencies of these cations in modulating drug binding. The auxiliary site-specific cations are unable to promote high-affinity complex formation between calmodulin and melittin.  相似文献   

4.
Y Hochman  A Lanir  C Carmeli 《FEBS letters》1976,61(2):255-259
Although 150 individual samples of milk from Italian water buffalo (Bubalus arnee) were examined by acid and alkaline gel electrophoresis, no polymorphism was observed for α-lactalbumin and β-lactoglobulin. After isolation and purification of these two proteins their amino acid compositions were determined and compared with those of the corresponding bovine proteins. The sequence alignments of 36 and 17 amino-acids from the N-terminal ends and 2 amino-acids from the C-terminal ends of buffalo α-lactalbumin and β-lactoglobulin, respectively, have been established. Our results indicate that buffalo α-lactalbumin differs from its cow B counterpart by a substitution Asn/Gly at position 17 and by another substitution, likely Glu/Gln or Asp/Asn, at an unknown position. Buffalo β-lactoglobulin is homologous to the bovine B variant. Three substitutions differentiate the two proteins: Ile/Leu and Val/Ile at positions 1 and 162 respectively; a further one, Gln/Ile, has not yet been located. According to these results the B variant of bovine β-lactoglobulin might be the wild type of the Bos genus.  相似文献   

5.
The actions of divalent cations on voltage-gated ion channels suggest that these cations bind to specific sites and directly influence gating kinetics. We have examined some chemical properties of the external divalent cation binding sites on neuronal potassium channels. Patch clamp techniques were used to measure the electrophysiological properties of these channels and Zn ions were used to probe the divalent cation binding site. The channel activation kinetics were greatly (three- to fourfold) slowed by low (2-5 mM) concentrations of Zn; deactivation kinetics were only slightly affected. These effects of Zn were inhibited by low solution pH in a manner consistent with competition between Zn and H ions for a single site. The apparent inhibitory pK for this site was near 7.2. Treatment of the neurons with specific amino acid reagents implicated amino, but no histidyl or sulfhydryl, residues in divalent cation binding.  相似文献   

6.
Summary The effects of divalent cation ionophores, A23187 and X-537A, on the electrical membrane properties were investigated by using the soma membrane of the X-organ of the crayfish. They reduced the amplitude and maximum rate of rise of Ca-action potential in lower concentration. As the concentration increased, a reduction of membrane resistance and hyperpolarization occurred simultaneously. Further increase resulted in membrane depolarization with a further decrease in resistance. The threshold concentration of X537A was 100 times higher than that of A23187. These effects were reversible only when the application period was relatively short, while a longer application resulted in an incomplete reversibility or in no reversibility at all. The ionophore effect was facilitated in high Ca medium and diminished in low Ca medium. In Sr medium, the same effects on the resistance and the membrane potential were barely observable. TEA reduced the effects of A23187 but did not completely inhibit the effects. The Na-action potential was also reduced by the higher concentration of the ionophore. From these results it is concluded that the divalent cation ionophores, A23187 and X537A, carry divalent cation, Ca ions in a physiological medium, into the neuron soma through the membrane and the consequent increase of the intracellular divalent cations induces K conductance increase and that higher concentration of the ionophore induces the increase in the conductance of the other ion species, such as Na.  相似文献   

7.
An X  Guo X  Sum H  Morrow J  Gratzer W  Mohandas N 《Biochemistry》2004,43(2):310-315
The erythrocyte membrane is a composite structure consisting of a lipid bilayer tethered to the spectrin-based membrane skeleton. Two complexes of spectrin with other proteins are known to participate in the attachment. Spectrin has also been shown to interact with phosphatidylserine (PS), a component of the lipid bilayer, which is confined to its inner leaflet. That there may be multiple sites of interaction with PS in the spectrin sequence has been inferred, but they have not hitherto been identified. Here we have explored the interaction of PS-containing liposomes with native alpha- and beta-spectrin chains and with recombinant spectrin fragments encompassing the entire sequences of both chains. We show that both alpha-spectrin and beta-spectrin bind PS and that sites of high affinity are located within 8 of the 38 triple-helical structural repeats which make up the bulk of both chains; these are alpha8, alpha9-10, beta2, beta3, beta4, beta12, beta13, and beta14, and PS affinity was also found in the nonhomologous N-terminal domain of the beta-chain. No other fragments of either chain showed appreciable binding. Binding of spectrin and its constituent chains to mixed liposomes of PS and phosphatidylcholine (PC) depended on the proportion of PS. Binding of spectrin dimers to PS liposomes was inhibited by single repeats containing PS binding sites. It is noteworthy that the PS binding sites in beta-spectrin are grouped in close proximity to the sites of attachment both of ankyrin and of 4.1R, the proteins engaged in attachment of spectrin to the membrane. We conjecture that direct interaction of spectrin with PS in the membrane may modulate its interactions with the proteins and that (considering also the known affinity of 4.1R for PS) the formation of PS-rich lipid domains, which have been observed in the red cell membrane, may be a result.  相似文献   

8.
9.
10.
The divalent metal ion binding sites of skeletal myosin were investigated by electron paramagnetic resonance (EPR) spectroscopy using the paramagnetic (Mn(II) ion as a probe. Myosin possesses two high affinity sites (K less than 1 muM) for Mn(II), which are located on the 5,5'-dithiobis(2-nitrobenzoate) (DTNB) light chains. Mn(II) bound to the isolated DTNB light chain gives rise to an EPR spectrum similar to that of Mn(II) bound to myosin and this indicates that the metal binding site comprises ligands from the DTNB light chain alone. Myosin preparations in which the DTNB light chain content is reduced by treatment with 5,5'-dithiobis(2-nitrobenzoate) show a corresponding reduction in the stoichiometry of Mn(II) binding, but the stoichiometry is recovered on reassociation of the DTNB light chain. Chymotryptic digestion of myosin filaments in the presence of ethylenediaminetetraacetic acid yields subfragment 1, but digestion in the presence of divalent metal ions produces heavy meromyosin. Myosin with a depleted DTNB light chain content gives rise to subfragment 1 on proteolysis, even in the presence of divalent metal ions. It is proposed that saturation of the DTNB light chain site with divalent ions protects this subunit against proteolysis, which, in turn, inhibits the cleavage of the subfragment 1-subfragment 2 link. Either the DTNB light chain is located near the region of the link and sterically blocks chymotryptic attack, or it is bound to the subfragment 1 moiety and affects the conformation of the link region. When the product heavy meromyosin was examined by sodium dodecyl sulfate gel electrophoresis, an apparent anomaly arose in that there was no trace of the 19 000-dalton band corresponding to the DTNB light chain. This was resolved by following the time course of chymotryptic digestion of the myosin heavy chain, the DTNB light chain, and the divalent metal binding site. The 19 000-dalton DTNB light chain is rapidly degraded to a 17 000-dalton fragment which comigrates with the alkali 2 light chain. The divalent metal site remains intact, despite this degradation, and the 17 000 fragment continues to protect the subfragment 1-subfragment 2 link. In the absence of divalent metal ions, the 17 000-dalton fragment is further degraded and attack of the subfragment 1 link ensues. Mn(II) bound to cardiac myosin gives an EPR spectrum basically similar to that of skeletal myosin, suggesting that their 19 000-dalton light chains are analogous with respect to their divalent metal binding sites, despite their chemical differences. The potential of EPR spectroscopy for characterizing the metal binding sites of myosin from different sources and of intact muscle fibers is discussed.  相似文献   

11.
The present work describes the results of a study aimed at identifying candidate cation binding sites on the extracellular region of bacteriorhodopsin, including a site near the retinal pocket. The approach used is a combined effort involving computational chemistry methods (computation of cation affinity maps and molecular dynamics) together with the Extended X-Ray Absorption Fine Structure (EXAFS) technique to obtain relevant information about the local structure of the protein in the neighborhood of Mn(2+) ions in different affinity binding sites. The results permit the identification of a high-affinity binding site where the ion is coordinated simultaneously to Asp212(-) and Asp85(-). Comparison of EXAFS data of the wild type protein with the quadruple mutant E9Q/E74Q/E194Q/E204Q at pH 7.0 and 10.0 demonstrate that extracellular glutamic acid residues are involved in cation binding.  相似文献   

12.
A sensitive high-resolution electron diffraction assay for change in structure is described and harnessed to analyze the binding of divalent cations to the purple membrane (PM) of Halobacterium halobium. Low-dose electron diffraction patterns are subject to a matched filter algorithm (Spencer, S. A., and A. A. Kossiakoff. 1980. J. Appl. Crystallogr. 13:563-571). to extract accurate values of reflection intensities. This, coupled with a scheme to account for twinning and specimen tilt in the microscope, yields results that are sensitive enough to rapidly quantitate any structure change in PM brought about by site-directed mutagenesis to the level of less than two carbon atoms. Removal of tightly bound divalent cations (mainly Ca2+ and Mg2+) from PM causes a color change to blue and is accompanied by a severely altered photocycle of the protein bacteriohodopsin (bR), a light-driven proton pump. We characterize the structural changes that occur upon association of 3:1 divalent cation to PM, versus membranes rendered purple by addition of excess Na+. High resolution, low dose electron diffraction data obtained from glucose-embedded samples of Pb2+ and Na+ reconstituted PM preparations at room temperature identify several sites with total occupancy of 2.01 +/- 0.05 Pb2+ equivalents. The color transition as a function of ion concentration for Ca2+ or Mg2+ and Pb2+ are strictly comparable. A (Pb2(+)-Na+) PM Fourier difference map in projection was synthesized at 5 A using the averaged data from several nominally untilted patches corrected for twinning and specimen tilt. We find six major sites located on helices 7, 5, 4, 3, 2 (nomenclature of Engelman et al. 1980. Proc. Natl. Acad. Sci. USA. 77:2023-2027) in close association with bR. These partially occupied sites (0.55-0.24 Pb2+ equivalents) represent preferential sites of binding for divalent cations and complements our earlier result by x-ray diffraction (Katre et al. 1986. Biophys. J. 50:277-284).  相似文献   

13.
A phosphorescent trivalent cation, terbium [Tb(III)], has been used to study the binding of different polyvalent cations to the proteins of wheat (Triticum aestivum L.) root plasma membranes. The phosphorescence emission intensity of Tb(III) was enhanced after Tb(III) binding to wheat root plasma membranes as a result of nonradiative resonance energy transfer from the membrane protein tyrosine and phenylalanine residues. Complex, saturable Tb(III) binding was observed, suggesting multiple binding sites. Bound Tb(III) could be displaced by divalent cations in the general order: Mn(II) > Ca(II) > Mg(II). Al(III) was very effective in reducing the protein-enhanced Tb(III) phosphorescence at pH values below 5. Al(III) also altered the Tb(III) phosphorescence lifetime, suggesting Al(III)-induced changes in membrane protein conformation. The more Al(III)-sensitive wheat cultivar (Anza) bound Al(III) with higher affinity than the more tolerant cultivar (BH 1146). At pH 5.5 where Al(III) did not displace bound Tb(III), low levels of Al(III) reduced the ability of Mn(II) to decrease Tb(III) phosphorescence. The significance of these results is discussed with respect to the mechanisms of Al(III) tolerance in wheat and the potential beneficial effects of Al(III) in reducing Mn(II) phytotoxicity.  相似文献   

14.
1. CO2-depletion of thylakoid membranes results in a decrease of binding affinity of the Photosystem II (PS II) inhibitor atrazine. The inhibitory efficiency of atrazine, expressed as I50-concentration (50% inhibition) of 2,6-dichlorophenolindophenol reduction, is the same in CO2-depleted as well as in control thylakoids. This shows that CO2-depletion results in a complete inactivation of a part of the total number of electron transport chains. 2. A major site of action of CO2, which had previously been located between the two electron acceptor quinone molecule B (or R) and Photosystem II inhibitor atrazine as suggested by the following observations: (a) CO2-depletion results in a shift of the binding constant (kappa b) of [14C]atrazine to thylakoid membranes indicating a decreased affinity of atrazine to membrane; (b) trypsin treatment, which is known to modify the Photosystem II complex at the level of B, strongly diminishes CO2 stimulation of electron transport reactions in CO2-depleted membranes; and (c) thylakoids from atrazine-resistant plants, which contain a Photosystem II complex modified at the inhibitor binding site, show an altered CO2-stimulation of electron flow. 3. CO2-depletion does not produce structural changes in enzyme complexes involved in Photosystem II function of thylakoid membranes, as shown by freeze-fracture studies using electron microscopy.  相似文献   

15.
Electron spin resonance linewidth measurements have been made on intact cell walls exchanged with various combinations of Mn2+ and Ca2+. These experiments were performed to find the Mn2+ nearest-neighbor distance and thereby determine whether carboxylate-Mn2+ complexes potentiate ion association at adjacent sites on cell wall polyuronides. Our results show that as the fraction of available binding sites occupied by Mn2+ increased from 2% to 27%, the nearest-neighbor distance parameter decreased only from 14 to 11 Å. These distances are close to polyuronide interanionic spacings. The small change in the distance parameter with concentration is evidence for sequential rather than random binding. Competitive ion-exchange with Ca2+ was found to reduce the Mn2+ spin-spin line broadening at similar total bound Mn2+ concentrations. This is expected only if Ca2+ competes at adjacent sites. The data presented offer strong support for the hypothesis that carboxylate groups near already occupied sites have a greater affinity for divalent cations than other sites along the polyuronide main chain.  相似文献   

16.
The chloroplast protein synthesis factor responsible for the translocation step of polypeptide synthesis on chloroplast ribosomes (chloroplast elongation factor G [EF-G]) has been detected in whole cell extracts and in isolated chloroplasts from Euglena gracilis. This factor can be detected by its ability to catalyze translocation on 70 S prokaryotic ribosomes such as those from E. coli. Chloroplast EF-G is present in low levels when Euglena is grown in the dark and can be induced more than 20-fold when the organism is grown in the light. The induction of this factor by light is inhibited by cycloheximide, a specific inhibitor of protein synthesis on cytoplasmic ribosomes. However, inhibitors of chloroplast protein synthesis such as streptomycin or spectinomycin have no effect on the induction of this factor by light. Furthermore, chloroplast EF-G can be partially induced by light in an aplastidic mutant (strain W3BUL) which has neither significant plastid structure nor detectable chloroplast DNA. These data strongly suggest that the genetic information for chloroplast EF-G resides in the nuclear genome, and that this protein is synthesized on cytoplasmic ribosomes prior to compartmentalization within the chloroplasts.  相似文献   

17.
The interaction of a number of lanthanide ions (namely terbium, praseodymium, erbium, lanthanum, gadolinium and europium) with Panulirus interruptus hemocyanin has been studied.Results from O2-binding experiments indicate that all these ions may substitute for calcium as allosteric effectors of hemocyanin. Addition of the lanthanides to deoxygenated Panulirus hemocyanin saturated with Tb3+ results in a quenching of the terbium luminescence. The highly efficient quenching observed in the case of Eu3+ may indicate energy-transfer between Tb3+ and Eu3+. Since energy-transfer between lanthanides is only effective over very short distances, the data suggest that some of the cation binding sites of Panulirus hemocyanin are clustered.  相似文献   

18.
T J Hallam  T J Rink 《FEBS letters》1985,186(2):175-179
Agonists such as thrombin, PAF (platelet-activating factor) and ADP are known to cause a larger elevation in [Ca2+]i in quin2-loaded platelets in the presence of extracellular Ca2+ than in its absence. The simplest interpretation of these observations is that in the presence of extracellular calcium there is an influx component across the cell surface. In the presence of Mn2+, a divalent cation which is known to avidly bind to quin2 and to quench its fluorescence, the agonists produce a small initial rise in quin2 fluorescence followed by a decrease in fluorescence to well below the resting level. The result indicates entry of Mn2+, presumably through some form of receptor-operated Ca2+ channel.  相似文献   

19.
Fluorescence resonance energy transfer was used to measure the distances between three nucleotide binding sites on solubilized chloroplast coupling factor from spinach and between each nucleotide site and two tyrosine residues which are important for catalytic activity. The nucleotide energy donor was 1,N6-ethenoadenosine di- or triphosphate, and the nucleotide energy acceptor was 2'(3')-(trinitrophenyl)adenosine diphosphate. The tyrosine residues were specifically labeled with 7-chloro-4-nitro-2,1,3-benzoxadiazole, which served as an energy acceptor. The results obtained indicate the three nucleotide binding sites form a triangle with sides of 44, 48, and 36 A. (The assumption has been made in calculating these distances that the energy donor and acceptor rotate rapidly relative to the fluorescence lifetime.) Two of the nucleotide sites are approximately equidistant from each of the two tyrosines: one of the nucleotide sites is about 37 A and the other about 41 A from each tyrosine. The third nucleotide site is about 41 A from one of the tyrosines and greater than or equal to 41 A from the other tyrosine.  相似文献   

20.
The location of binding sites on C1q for DNA   总被引:1,自引:0,他引:1  
Previous studies have suggested that C1q reacts with DNA via both the globular region of C1q (GR) and the collagen-like region of C1q (CLR). In this study, the binding of dsDNA and ssDNA to GR and CLR was quantitated by a solid-phase assay. Both dsDNA and ssDNA bound to the GR and CLR of C1q in an ionic strength-dependent manner. Under physiologic salt concentrations, however, dsDNA and ssDNA bound preferentially to CLR and not to GR. The binding of dsDNA to C1q was not affected by heat inactivation of C1q or its exposure to pH 4.45, which abolished the binding of heat-aggregated human IgG (AHG) with C1q. The preincubation of the solid-phase C1q with AHG did not decrease the binding of dsDNA or ssDNA to the solid-phase C1q. These results indicate that the major sites for binding DNA to C1q are located in the CLR of C1q and are not overlapping with those for AHG or immune complexes.  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