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.
The binding of Mg2+ to the ribitol teichoic acid of Staphylococcus aureus H walls was examined by equilibrium dialysis in solution and in the intact wall; the influence of alanyl ester groups on binding was determined. In solution the ribitol polymer had a lower affinity than did a glycerol teichoic acid and bound Mg2+ in the ratio Mg2+/P of 1:1. The presence of alanyl ester residues caused a decrease in the amount of cations bound in stoicheiometric proportion to the ratio Ala/P, but the affinity constant was unaltered. It is concluded that in solution the ribitol teichoic acid binds Mg2+ univalently to phosphate groups and univalently to a counter-ion. In the intact wall the binding of Mg2+ was different. The affinity constant was higher and resembled that of a glycerol teichoic acid. It is concluded that Mg2+ forms bridges across phosphate groups in teichoic acid chains lying adjacent to each other in the wall. The effect of alanyl esters was similar to that in solution, but Scatchard plots were not linear at low concentrations of Mg2+ where it was shown that the difference in affinities between walls with and without alanyl ester residues was much greater than it was at higher concentrations of Mg2+. Thus at very low concentrations of Mg2+ effective binding to the wall is markedly improved by loss of alanyl ester residues.  相似文献   

2.
Danel F  Paetzel M  Strynadka NC  Page MG 《Biochemistry》2001,40(31):9412-9420
The factors influencing the oligomerization state of OXA-10 and OXA-14 class D beta-lactamases in solution have been investigated. Both enzymes were found to exist as an equilibrium mixture of a monomer and dimer, with a K(d) close to 40 microM. The dimeric form was stabilized by divalent metal cations. The ability of different metal ions to stabilize the dimer was in the following order: Cd(2+) > Cu(2+) > Zn(2+) > Co(2+) > Ni(2+) > Mn(2+) > Ca(2+) > Mg(2+). The apparent K(d)s describing the binding of Zn(2+) and Cd(2+) cations to the OXA-10 dimer were 7.8 and 5.7 microM, respectively. The metal ions had a profound effect on the thermal stability of the protein complex observed by differential scanning calorimetry. The enzyme showed a sharp transition with a T(m) of 58.7 degrees C in the absence of divalent cations, and an equally sharp transition with a T(m) of 78.4 degrees C in the presence of a saturating concentration of the divalent cation. The thermal transition observed at intermediate concentrations of divalent metal ions was rather broad and lies between these two extremes of temperature. The equilibrium between the monomer and dimer is dependent on pH, and the optimum for the formation of the dimer shifted from pH 6.0 in the absence of divalent cations to pH 7.5 at saturating concentrations. The beta-lactamase activity increased approximately 2-fold in the presence of saturating concentrations of zinc and cadmium ions. Reaction with beta-lactams caused a shift in the equilibrium toward monomer formation, and thus an apparent inactivation, but the divalent cations protected against this effect.  相似文献   

3.
The interaction of aqueous phospholipid dispersions of negatively charged 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol, sodium salt (DMPG) with the divalent cations Mg(2+), Ca(2+) and Sr(2+) at equimolar ratios in 100 mM NaCl at pH 7 was investigated by Fourier transform infrared spectroscopy. The binding of the three cations induces a crystalline-like gel phase with highly ordered and rigid all-trans acyl chains. These features are observed after storage below room temperature for 24 h. When the gel phase is heated after prolonged incubation at low temperature phase transitions into the liquid crystalline phase are observed at 58 degrees C for the DMPG:Sr(2+), 65 degrees C for the DMPG:Mg(2+), and 80 degrees C for the DMPG:Ca(2+) complex. By subsequent cooling from temperatures above T(m) these complexes retain the features of a liquid crystalline phase with disordered acyl chains until a metastable gel phase is formed at temperatures between 38 and 32 degrees C. This phase is characterized by predominantly all-trans acyl chains, arranged in a loosely packed hexagonal or distorted hexagonal subcell lattice. Reheating the DMPG:Sr(2+) samples after a storage time of 2 h at 4 degrees C results in the transition of the metastable gel to the liquid crystalline phase at 35 degrees C. This phase transition into the liquid crystalline state at 35 degrees C is also observed for the Mg(2+) complex. However, for DMPG:Mg(2+) at higher temperatures, a partial recrystallization of the acyl chains occurs and the high temperature phase transition at 65 degrees C is also detected. In contrast, DMPG:Ca(2+) exhibits only the phase transition at 80 degrees C from the crystalline gel into the fluid state upon reheating. Below 20 degrees C, the rate of conversion from the metastable gel to a thermodynamically stable, crystalline-like gel phase decreases in the order Ca(2+)&z. Gt;Mg(2+)>Sr(2+). This conversion into the crystalline gel phase is accompanied by a complete dehydration of the phosphate groups in DMPG:Mg(2+) and by a reorientation of the polar lipid head groups in DMPG:Ca(2+) and in DMPG:Sr(2+). The primary binding sites of the cations are the PO(2)(-) groups of the phosphodiester moiety. Our infrared spectroscopic results suggest a deep penetration of the divalent cations into the polar head group region of DMPG bilayers, whereby the ester carbonyl groups, located in the interfacial region of the bilayers, are indirectly affected by strong hydrogen bonding of immobilized water molecules. In the liquid crystalline phase, the interaction of all three cations with DMPG is weak, but still observable in the infrared spectra of the DMPG:Ca(2+) complex by a slight ordering effect induced in the acyl chains, when compared to pure DMPG liposomes.  相似文献   

4.
We characterized the human Na(+)-ascorbic acid transporter SVCT2 and developed a basic model for the transport cycle that challenges the current view that it functions as a Na(+)-dependent transporter. The properties of SVCT2 are modulated by Ca(2+)/Mg(2+) and a reciprocal functional interaction between Na(+) and ascorbic acid that defines the substrate binding order and the transport stoichiometry. Na(+) increased the ascorbic acid transport rate in a cooperative manner, decreasing the transport K(m) without affecting the V(max), thus converting a low affinity form of the transporter into a high affinity transporter. Inversely, ascorbic acid affected in a bimodal and concentration-dependent manner the Na(+) cooperativity, with absence of cooperativity at low and high ascorbic acid concentrations. Our data are consistent with a transport cycle characterized by a Na(+):ascorbic acid stoichiometry of 2:1 and a substrate binding order of the type Na(+):ascorbic acid:Na(+). However, SVCT2 is not electrogenic. SVCT2 showed an absolute requirement for Ca(2+)/Mg(2+) for function, with both cations switching the transporter from an inactive into an active conformation by increasing the transport V(max) without affecting the transport K(m) or the Na(+) cooperativity. Our data indicate that SVCT2 may switch between a number of states with characteristic properties, including an inactive conformation in the absence of Ca(2+)/Mg(2+). At least three active states can be envisioned, including a low affinity conformation at Na(+) concentrations below 20 mM and two high affinity conformations at elevated Na(+) concentrations whose Na(+) cooperativity is modulated by ascorbic acid. Thus, SVCT2 is a Ca(2+)/Mg(2+)-dependent transporter.  相似文献   

5.
1. The effects of teichoic acids on the Mg(2+)-requirement of some membrane-bound enzymes in cell preparations from Bacillus licheniformis A.T.C.C. 9945 were examined. 2. The biosynthesis of the wall polymers poly(glycerol phosphate glucose) and poly(glycerol phosphate) by membrane-bound enzymes is strongly dependent on Mg(2+), showing maximum activity at 10-15mm-Mg(2+). 3. When the membrane is in close contact with the cell wall and membrane teichoic acid, the enzyme systems are insensitive to added Mg(2+). The membrane appears to interact preferentially with the constant concentration of Mg(2+) that is bound to the phosphate groups of teichoic acid in the wall and on the membrane. When the wall is removed by the action of lysozyme the enzymes again become dependent on an external supply of Mg(2+). 4. A membrane preparation that retained its membrane teichoic acid was still dependent on Mg(2+) in solution, but the dependence was damped so that the enzymes exhibited near-maximal activity over a much greater range of concentrations of added Mg(2+); this preparation contained Mg(2+) bound to the membrane teichoic acid. The behaviour of this preparation could be reproduced by binding membrane teichoic acid to membranes in the presence of Mg(2+). Addition of membrane teichoic acid to reaction mixtures also had a damping effect on the Mg(2+) requirement of the enzymes, since the added polymer interacted rapidly with the membrane. 5. Other phosphate polymers behaved in a qualitatively similar way to membrane teichoic acid on addition to reaction mixtures. 6. It is concluded that in whole cells the ordered array of anionic wall and membrane teichoic acids provides a constant reservoir of bound bivalent cations with which the membrane preferentially interacts. The membrane teichoic acid is the component of the system which mediates the interaction of bound cations with the membrane. The anionic polymers in the wall scavenge cations from the medium and maintain a constant environment for the membrane teichoic acid. Thus a function of wall and membrane teichoic acids is to maintain the correct ionic environment for cation-dependent membrane systems.  相似文献   

6.
It is known that phosphatidylserine (PS(-)) lipids have a very similar affinity for Ca(2+) and Mg(2+) cations, as revealed by electrokinetic and stability experiments. However, despite this similar affinity, experimental evidence shows that the presence of Ca(2+) or Mg(2+) induces very different aggregation behavior for PS(-) liposomes as characterized by their fractal dimensions. Also, turbidity measurements confirm substantial differences in aggregation behavior depending on the presence of Ca(2+) or Mg(2+) cations. These puzzling results suggest that although these two cations have a similar affinity for PS(-) lipids, they induce substantial structural differences in lipid bilayers containing each of these cations. In other words, these cations have strong ion-specific effects on the structure of PS(-) membranes. This interpretation is supported by all-atomic molecular-dynamics simulations showing that Ca(2+) and Mg(2+) cations have different binding sites and induce different membrane hydration. We show that although both ions are incorporated deep into the hydrophilic region of the membrane, they have different positions and configurations at the membrane. Absorbed Ca(2+) cations present a peak at a distance ~2 nm from the center of the lipid bilayer, and their most probable binding configuration involves two oxygen atoms from each of the charged moieties of the PS molecule (phosphate and carboxyl groups). In contrast, the distribution of absorbed Mg(2+) cations has two different peaks, located a few angstroms before and after the Ca(2+) peak. The most probable configurations (corresponding to these two peaks) involve binding to two oxygen atoms from carboxyl groups (the most superficial binding peak) or two oxygen atoms from phosphate groups (the most internal peak). Moreover, simulations also show differences in the hydration structure of the membrane: we obtained a hydration of 7.5 and 9 water molecules per lipid in simulations with Ca(2+) and Mg(2+), respectively.  相似文献   

7.
The ionic structure of lecithin monolayers   总被引:4,自引:0,他引:4  
Surface potentials of mixed monolayers of dicetyl phosphate and eicosanyl trimethylammonium bromide (1:1) were the same on subsolutions of 0.02 M NaCl or 0.01 M CaCl(2), which indicated that ionic phosphate does not interact with Ca(++) in the presence of a neighboring trimethylammonium group. Surface potential-pH plots of dicetyl phosphate, and of dipalmitoyl, egg, and dioleoyl lecithins showed that as the pH of the subsolution is decreased the phosphate groups in the monolayer are neutralized in the order: dicetyl phosphate > dipalmitoyl lecithin > egg lecithin > dioleoyl lecithin. The binding of cations (Na(+), Ca(++)) to the phosphate group of lecithin also showed the same order. The binding of Ca(++)) to egg phosphatidic acid monolayers, as measured by the increase in surface potential, is considerably greater than that to egg lecithin. These results suggest that there is an internal salt linkage between the phosphate and trimethylammonium groups on the same lecithin molecule. An increase in unsaturation of fatty acyl chains increases the intermolecular spacing, which reduces the ionic repulsion between polar groups, and hence strengthens the internal salt linkage. The results support the concept of a vertical rather than coplanar orientation of the phosphoryl choline group with respect to the interface. A position has been proposed for Ca(++) in the dipole lattice of lecithin from a consideration of the surface potential measurements.  相似文献   

8.
1. The interaction of aflatoxin B(1) with different polynucleotides was studied spectrophotometrically. Equations were derived that enable the degree of binding to be determined without first determining the extinction coefficient of the bound form. 2. The interaction with calf thymus DNA obeys first-order relationships with an association constant of 0.40mm(-1), but there is some evidence for a secondary binding process from results obtained at 390nm. 3. The spectral shifts decreased in the order polyadenylic acid+polyuridylic acid>DNA>polyadenylic acid>polyadenylic acid+polyinosinic acid. Polycytidylic acid, polyuridylic acid, polyinosinic acid (both single- and triple-stranded), AMP, CMP, GMP and UMP did not interact with aflatoxin. It was concluded that there is a requirement for the amino group of adenine (or possibly guanine) for binding of aflatoxin to polynucleotides to occur. 4. Binding is reversed by increasing ionic strength, and by Mn(2+) and Mg(2+) in the concentration range studied (0-5mm). The effect of the Mn(2+) or Mg(2+) was far greater than would be expected on the basis of their ionic strength. With both the bivalent cations and sodium chloride the reversal is greatest with double-stranded polynucleotides. 5. Inhibition in vitro of the DNA-dependent RNA polymerase of Escherichia coli by aflatoxin B(1) was detected only in the absence of Mg(2+) and at concentrations of Mn(2+) below the optimum for RNA synthesis in vitro. 6. The degree of inhibition (maximally 30%) was dependent on the concentration of Mn(2+) and decreased during incubation.  相似文献   

9.
Mg(2+) at an optimal concentration of 2mM (ph 6.5) induces large increases (up to 30 percent) in the optical density of bovine heart mitochondria incubated under conditions of low ionic strength (< approx. 0.01). The increases are associated with aggregation (sticking together) of the inner membranes and are little affected by changes in the energy status of the mitochondria. Virtually all of a number of other polyvalent cations tested and Ag(+) induce increases in mitochondrial optical density similar to those induced by Mg(2+), their approximate order of concentration effectiveness in respect to Mg(2+) being: La(3+) > Pb(2+) = Cu(2+) > Cd(2+) > Zn(2+) > Ag(+) > Mn(2+) > Ca(2+) > Mg(2+). With the exception of Mg(2+), all of these cations appear to induce swelling of the mitochondria concomitant with inner membrane aggregation. The inhibitors of the adenine nucleotide transport reaction carboxyatratyloside and bongkrekic acid are capable of preventing and reversing Mg(2+)-induced aggregation at the same low concentration required for complete inhibition of phosphorylating respiration, suggesting that they inhibit the aggregation by binding to the adenine nucleotide carrier. The findings are interpreted to indicate (a) that the inner mitochondrial membrane is normally prevented from aggregating by virtue of its net negative outer surface change, (b) that the cations induce the membrane to aggregate by binding at its outer surface, decreasing the net negative charge, and (c) that carboxyatractyloside and bongkrekic acid inhibit the aggregation by binding to the outer surface of the membrane, increasing the net negative charge.  相似文献   

10.
C Ebel  P Faou  B Kernel  G Zaccai 《Biochemistry》1999,38(28):9039-9047
Halophilic malate dehydrogenase unfolds at low salt, and increasing the salt concentration stabilizes, first, the folded form and then, in some cases, destabilizes it. From inactivation and fluorescence measurements performed on the protein after its incubation in the presence of various salts in a large range of concentrations, the apparent effects of anions and cations were found to superimpose. A large range of ions was examined, including conditions that are in general not of physiological relevance, to explore the physical chemistry driving adaptation to extreme environments. The order of efficiency of cations and anions to maintain the folded form is, for the low-salt transition, Ca(2+) approximately Mg(2+) > Li(+) approximately NH(4)(+) approximately Na(+) > K(+) > Rb(+) > Cs(+), and SO(4)(2)(-) approximately OAc(-) approximately F(-) > Cl(-), and for the high-salt transition, NH(4)(+) approximately Na(+) approximately K(+) approximately Cs(+) > Li(+) > Mg(2+) > Ca(2+), and SO(4)(2)(-) approximately OAc(-) approximately F(-) > Cl(-) > Br(-) > I(-). If a cation or anion is very stabilizing, the effect of the salt ion of opposite charge is limited. Anions of high charge density are always the most efficient to stabilize the folded form, in accordance with the order found in the Hofmeister series, while cations of high charge density are the most efficient only at the lower salt concentrations and tend to denature the protein at higher salt concentrations. The stabilizing efficiency of cations and anions can be related in a minor way to their effect on the surface tension of the solution, but the interaction of ions with sites only present in the folded protein has also to be taken into account. Unfolding at high salt concentrations corresponds to interactions of anions of low charge density and cations of high charge density with the peptide bond, as found for nonhalophilic proteins.  相似文献   

11.
Yamaguchi N  Xu L  Pasek DA  Evans KE  Chen SR  Meissner G 《Biochemistry》2005,44(45):15074-15081
Ryanodine receptors (RyRs) are a family of intracellular Ca(2+) channels that are regulated by calmodulin (CaM). At low Ca(2+) concentrations (<1 microM), CaM activates RyR1 and RyR3 and inhibits RyR2. At elevated Ca(2+) concentrations (>1 microM), CaM inhibits all three RyR isoforms. Here we report that the regulation of recombinant RyR3 by CaM is sensitive to redox regulation. RyR3 in the presence of reduced glutathione binds CaM with 10-15-fold higher affinity, at low and high Ca(2+) concentrations, compared to in the presence of oxidized glutathione. However, compared to RyR1 assayed at low Ca(2+) concentrations under both reducing and oxidizing conditions, CaM binds RyR3 with reduced affinity but activates RyR3 to a greater extent. Under reducing conditions, RyR1 and RyR3 activities are inhibited with a similar affinity at [Ca(2+)] > 1 microM. Mutagenesis studies demonstrate that RyR3 contains a single conserved CaM binding site. Corresponding amino acid substitutions in the CaM binding site differentially affect CaM binding and CaM regulation of RyR3 and those of the two other isoforms. The results support the suggestion that other isoform dependent regions have a major role in the regulation of RyRs by CaM [Yamaguchi et al. (2004) J. Biol. Chem. 279, 36433-36439].  相似文献   

12.
The ionic specificity of IAA-induced acidification by Avena coleoptiles was studied, using zwitterionic, presumably impermeant buffers. The acidification was almost totally dependent on divalent cations with an order of effectiveness of Ca(2+) >/= Sr(2+) > Mn(2+), Mg(2+); whereas other polyvalent cations tested were ineffective. The Ca(2+) response was IAA-dependent. The CaCl(2) concentration was optimal at 0.3 to 1 mm and inhibitory at higher concentrations. Sr(2+) inhibited Ca(2+)-dependent acidification and monovalent cations such as K(+) did not induce additional acidification in the presence of optimal CaCl(2). These data are consistent with a mechanism for IAA-induced acidification involving a Ca(2+) -H(+) exchange.  相似文献   

13.
Ba(2+), a doubly charged analogue of K(+), specifically blocks K(+) channels by virtue of electrostatic stabilization in the permeation pathway. Ba(2+) block is used here as a tool to determine the equilibrium binding affinity for various monovalent cations at specific sites in the selectivity filter of a noninactivating mutant of KcsA. At high concentrations of external K(+), the block-time distribution is double exponential, marking at least two Ba(2+) sites in the selectivity filter, in accord with a Ba(2+)-containing crystal structure of KcsA. By analyzing block as a function of extracellular K(+), we determined the equilibrium dissociation constant of K(+) and of other monovalent cations at an extracellular site, presumably S1, to arrive at a selectivity sequence for binding at this site: Rb(+) (3 μM) > Cs(+) (23 μM) > K(+) (29 μM) > NH(4)(+) (440 μM) > Na(+) and Li(+) (>1 M). This represents an unusually high selectivity for K(+) over Na(+), with |ΔΔG(0)| of at least 7 kcal mol(-1). These results fit well with other kinetic measurements of selectivity as well as with the many crystal structures of KcsA in various ionic conditions.  相似文献   

14.
Magnesium (Mg2+) increases binding of follicle-stimulating hormone (FSH) to membrane-bound receptors and increases adenylyl cyclase activity. We examined the effects of divalent and monovalent cations on FSH binding to receptors in granulosa cells from immature porcine follicles. Divalent and monovalent cations increased binding of [125I]iodo-porcine FSH (125I-pFSH). The divalent cations Mg2+, calcium (Ca2+) and manganese, (Mn2+) increased specific binding a maximum of 4- to 5-fold at added concentrations of 10 mM. Mg2+ caused a half-maximal enhancement of binding at 0.6 mM, whereas Ca2+ and Mn2+ had half-maximal effects at 0.7 mM and 0.8 mM, respectively. The monovalent cation potassium (K+) increased binding a maximum of 1.5-fold at an added concentration of 50 mM, whereas the monovalent cation (Na+) did not increase binding at any concentration tested. The difference between K+ and Na+ suggested that either enhancement of binding was not a simple ionic effect or Na+ has a negative effect that suppresses its positive effect. Ethylenediamine tetraacetic acid, a chelator of Mg2+, prevented binding of 125I-pFSH only in the presence of Mg2+, whereas pregnant mare's serum gonadotropin, a competitor with FSH for the receptor, prevented binding in both the absence and the presence of Mg2+. Guanyl-5-ylimidodiphosphate (Gpp[NH]p) inhibited binding of 125I-pFSH in the absence or presence of Mg2+, but only at Gpp(NH)p concentrations greater than 1 mM. We used Mg2+ to determine if divalent cations enhanced FSH binding by increasing receptor affinity or by increasing the apparent number of binding sites.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
L-ascorbic acid preincubation with rabbit kidney (Na+-K+)-ATPase inhibits the activity of p-nitrophenylphosphatase partial reaction with a pseudo-first order decay. The presence of the pseudosubstrate, p-nitrophenylphosphate, counteracts the inhibiting effect. During the reaction course, the kinetic rate is enhanced at ascorbic acid concentrations below 0.7 mmol/1, but is inhibited above that amount. The intrinsic fluorescence of the enzyme in E1 and E2 conformations is modified suggesting the occurrence of ascorbate-induced intermediate forms, distinct from those provoked by the addition of cations, magnesium and phosphate. These destabilized forms appear easier to be converted into catalytically active or increasingly inhibited states.  相似文献   

16.
1. Microsomes from guinea-pig brain grey matter were incubated with [(32)P]ATP at 3mm concentration and the phosphate bound to the acid-washed, lipid-free residue was determined. 2. The binding process was Mg(2+)-dependent and resulted in the transfer of about 1-2 mmumoles of phosphate/mg. of protein/min. Under the conditions used univalent cations (Na(+),K(+) and Li(+)) inhibited the binding. 3. An unspecified proportion of this bound phosphate could be recovered in protein-derived phosphorylserine. The yield of labelled phosphorylserine was also decreased by univalent cations. 4. The bound phosphate formed with 3mm-MgATP was stable; addition of Na(+) or K(+) ions to the already labelled preparation had no effect on the bound phosphate level. 5. Bound phosphate was also formed when a solubilized fraction of the microsomes was incubated with ATP; univalent cations also inhibited this process. 6. p-Chloromercuribenzoate reduced the binding by about 25%; the inhibition was restored by cysteine.  相似文献   

17.
In this study, we performed all-atom long-timescale molecular dynamics simulations of phospholipid bilayers incorporating three different proportions of negatively charged lipids in the presence of K(+), Mg(2+), and Ca(2+) ions to systemically determine how membrane properties are affected by cations and lipid compositions. Our simulations revealed that the binding affinity of Ca(2+) ions with lipids is significantly stronger than that of K(+) and Mg(2+) ions, regardless of the composition of the lipid bilayer. The binding of Ca(2+) ions to the lipids resulted in bilayers having smaller lateral areas, greater thicknesses, greater order, and slower rotation of their lipid head groups, relative to those of corresponding K(+)- and Mg(2+)-containing systems. The Ca(2+) ions bind preferentially to the phosphate groups of the lipids. The complexes formed between the cations and the lipids further assembled to form various multiple-cation-centered clusters in the presence of anionic lipids and at higher ionic strength-most notably for Ca(2+). The formation of cation-lipid complexes and clusters dehydrated and neutralized the anionic lipids, creating a more-hydrophobic environment suitable for membrane aggregation. We propose that the formation of Ca(2+)-phospholipid clusters across apposed lipid bilayers can work as a "cation glue" to adhere apposed membranes together, providing an adequate configuration for stalk formation during membrane fusion.  相似文献   

18.
Rod photoreceptors contain three different glutamic acid-rich proteins (GARPs) that have been proposed to control the propagation of Ca(2+) from the site of its entry at the cyclic nucleotide-gated channel to the cytosol of the outer segment. We tested this hypothesis by measuring the binding of Ca(2+) to the following five constructs related to GARPs of rod photoreceptors: a 32-mer peptide containing 22 carboxylate groups, polyglutamic acid, a recombinant segment comprising 73 carboxylate groups (GLU), GARP1, and GARP2. Ca(2+) binding was investigated by means of a Ca(2+)-sensitive electrode. In all cases, Ca(2+) binds with low affinity; the half-maximum binding constant K(1/2) ranges from 6 to 16 mM. The binding stoichiometry between Ca(2+) ions and carboxylic groups is approximately 1:1; an exception is GARP2, where a binding stoichiometry of approximately 1:2 was found. Hydrodynamic radii of 1.6, 2.8, 3.3, 5.7, and 6.7 nm were determined by dynamic light scattering for the 32-mer, polyglutamic acid, GLU, GARP2, and GARP1 constructs, respectively. These results suggest that the peptides as well as GARP1 and GARP2 do not adopt compact globular structures. We conclude that the structures should be regarded as loose coils with low-affinity, high-capacity Ca(2+) binding.  相似文献   

19.
Brooks KM  Hampel KJ 《Biochemistry》2011,50(13):2424-2433
The glmS ribozyme is a conserved riboswitch found in numerous Gram-positive bacteria and responds to the cellular concentrations of glucosamine 6-phosphate (GlcN6P). GlcN6P binding promotes site-specific self-cleavage in the 5' UTR of the glmS mRNA, resulting in downregulation of gene expression. The glmS ribozyme has previously been shown to lack strong cation specificity when the rate-limiting folding step of the cleavage reaction pathway is measured. This does not provide data regarding cation and ligand specificities of the glmS ribozyme during the rapid ligand binding chemical catalysis events. Prefolding of the ribozyme in Mg(2+)-containing buffers effectively isolates the rapid ligand binding and catalytic events (k(obs) > 60 min(-1)) from rate-limiting folding (k(obs) < 4 min(-1)). Here we employ this experimental design to assay the cations and ligand requirements for rapid ligand binding and catalysis. We show that molar concentrations of monovalent cations are also capable of inducing the formation of the native GlcN6P binding structure but are unable to promote ligand binding and catalysis rates of >4 min(-1). Our data show that the sole obligatory role for divalent cations, for which there is crystallographic evidence, is coordination of the phosphate moiety of GlcN6P in the ligand-binding pocket. In further support of this hypothesis, our data show that a nonphosphorylated analogue of GlcN6P, glucosamine, is unable to promote rapid ligand binding and catalysis in the presence of divalent cations. Folding of the ribozyme is, therefore, relatively independent of cation identity, but the rapid initiation of catalysis upon the addition of ligand is stricter.  相似文献   

20.
Inhibition by Zn(2+) of iron uptake by apoferritin at very low substrate concentrations is shown to be competitive. It is proposed that Zn(2+) competes with Fe(2+) for sites on the protein at which the oxidation of Fe(2+) is catalysed. Interpretation of titration data suggests there are two independent classes of binding site for Zn(2+) and several other cations. Sites in one such class are probably on the external surface of the apoferritin molecule. The catalytic binding sites are presumed to be internal and may involve histidine or possibly cysteine as ligands.  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