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.
The binding of carrier ampholytes to polyanions is markedly pH-dependent: it is very strong at pH 3, rather weak at pH 5 and abolished at pH 7. Binding is affected by the type of negative charge, its density and spatial orientation on the polyanion. On the basis of the type of negative charge, the binding strength decreases in the following order: polyphosphate greater than polysulphate greater than polycarboxylate. Given the same type of negative charge, the binding is dependent on charge density and its space orientation: thus polyglutamic acid forms stronger complexes than polygalacturonic acid. The minimum length of the polyanion eliciting a measurable binding appears to be of the order of about six negative charges, as demonstrated with hexametaphosphate.  相似文献   

2.
Evidence is given for a high density of negative surface charge near the sodium channel of myelinated nerve fibres. The voltage dependence of peak sodium permeability is measured in a voltage clamp. The object is to measure voltage shifts in sodium activation as the following external variables are varied: divalent cation concentration and type, monovalent concentration, and pH. With equimolar substitution of divalent ions the order of effectiveness for giving a positive shift is: Ba equals Sr less than Mg less than Ca less than Co approximately equal to Mn less than Ni less than Zn. A tenfold increase of concentration of any of these ions gives a shift of +20 to +25 mV. At low pH, the shift with a tenfold increase in Ca-2+ is much less than at normal pH, and conversely for high pH. Soulutions with no added divalent ions give a shift of minus 18 mV relative to 2 mM Ca-2+. Removal of 7/8 of the cations from the calcium-free solution gives a further shift of minue 35 mV. All shifts are explained quantitatively by assuming that changes in an external surface potential set up by fixed charges near the sodium channel produce the shifts. The model involves a diffuse double layer of counterions at the nerve surface and some binding of H+ions and divalent ions to the fixed charges. Three types of surface groups are postulated: (1) an acid pKa equals 2.88 charge density minus 0.9 nm- minus 2; (i) an acid pKa equals 4.58, charge density minus 0.58 nm- minus 2; (3) a base pKa equals 6.28, charge density +0.33 nm- minus 2. The two acid groups also bind Ca-2+ ions with a dissociation constant K equals 28 M. Reasonable agreement can also be obtained with a lower net surface charge density and stronger binding of divalent ions and H+ ions.  相似文献   

3.
A. Telfer  J. Barber  A.T. Jagendorf 《BBA》1980,591(2):331-345
1. Increase in electron transport rate and the decay rate of the 518 nm absorption change, induced by EDTA treatment, is prevented by cations. The order of effectiveness is C3+ > C2+ > C+.2. In this respect methyl viologen is an effective divalent cation in addition to its action as an electron acceptor.3. Complete cation irreversible EDTA-induced uncoupling occurs in the dark in 2 min. Light greatly stimulates the rate of uncoupling by EDTA. It is concluded that the uncoupling is due to release of coupling factor I from the thylakoid membrane.4. Binding of purified coupling factor I to coupling factor I-depleted thylakoids can be achieved with any cation. The order of effectiveness is C3+ > C2+ > C+, reconstituted thylakoids are active in photophosphorylation regardless of the cation used for coupling factor I binding.5. The marked difference in the concentration requirements for cation effects on 9-aminoacridine fluorescence yield and for prevention of uncoupling by EDTA indicate that coupling factor I and its binding site have a lower surface charge density than the net surface charge density of the thylakoid membrane.6. It is concluded that coupling factor I binding only occurs when negative charges on coupling factor I and its binding site are electrostatically screened by cations.7. Previously reported examples of uncoupling by low ionic conditions are discussed in relation to the basic concepts of diffuse electrical layer theory.  相似文献   

4.
A pH decrease in chloroplast suspension in media of low salt concentration was observed when a salt was added at pH values higher than 4.4, while at lower pH values a pH increase was observed. The salt-induced pH changes depended on the valence and concentration of cations of added salts at neutral pH values (higher than 4.4) and on those of anions at acidic pH values (lower than 4.4). The order of effectiveness was trivalent > divalent > monovalent. The pH value change by salt addition was affected by the presence of ionic detergents depending on the sign of their charges. These characteristics agreed with those expected from the Gouy-Chapman theory on diffuse electrical double layers. The results were interpreted in terms of the changes in surface potential, surface pH and the ionization of surface groups which result in the release (or binding) of H+ to (or from) the outer medium.The analysis of the data of KCl-induced pH change suggests that the change in the surface charge density of thylakoid membranes depends mainly on the ionization of carboxyl groups, which is determined by the surface pH. When the carboxyl groups are fully dissociated, the surface charge density reaches ?1.0 ± 0.1 · 10?3 elementary charge/square Å.Dependence of the estimated surface potential on the bulk pH was similar to that of electrophoretic mobility of thylakoid membrane vesicles.  相似文献   

5.
A synthetic polyanion (a copolymer of methacrylate, maleate, and styrene in 1:2:3 proportion with an average molecular weight of 10,000 dalton) inhibits the tricarboxylate, oxoglutarate, dicarboxylate, and adenine nucleotide translocators of rat liver mitochondria. The activity versus inhibitor concentration curves are sigmoidal. The inhibition of the oxoglutarate and tricarboxylate translocators by the polyanion is competitive, while that of the adenine nucleotide translocator is of mixed-type. TheK 1 values of the polyanion are the following: for oxoglutarate translocator 4.0 µM, tricarboxylate translocator 1.2 µM, and adenine nucleotide translocator 1.3 µM with ADP and 0.8 µM with ATP. It is suggested that the polyanion acts primarily by increasing the negative charge of the inner membrane at the outer surface, and the sensitivity of the translocators toward the polyanion depends on the number of negative charges of their substrates.  相似文献   

6.
This work investigates statistical prevalence and overall physical origins of changes in charge states of receptor proteins upon ligand binding. These changes are explored as a function of the ligand type (small molecule, protein, and nucleic acid), and distance from the binding region. Standard continuum solvent methodology is used to compute, on an equal footing, pK changes upon ligand binding for a total of 5899 ionizable residues in 20 protein-protein, 20 protein-small molecule, and 20 protein-nucleic acid high-resolution complexes. The size of the data set combined with an extensive error and sensitivity analysis allows us to make statistically justified and conservative conclusions: in 60% of all protein-small molecule, 90% of all protein-protein, and 85% of all protein-nucleic acid complexes there exists at least one ionizable residue that changes its charge state upon ligand binding at physiological conditions (pH = 6.5). Considering the most biologically relevant pH range of 4-8, the number of ionizable residues that experience substantial pK changes (ΔpK > 1.0) due to ligand binding is appreciable: on average, 6% of all ionizable residues in protein-small molecule complexes, 9% in protein-protein, and 12% in protein-nucleic acid complexes experience a substantial pK change upon ligand binding. These changes are safely above the statistical false-positive noise level. Most of the changes occur in the immediate binding interface region, where approximately one out of five ionizable residues experiences substantial pK change regardless of the ligand type. However, the physical origins of the change differ between the types: in protein-nucleic acid complexes, the pK values of interface residues are predominantly affected by electrostatic effects, whereas in protein-protein and protein-small molecule complexes, structural changes due to the induced-fit effect play an equally important role. In protein-protein and protein-nucleic acid complexes, there is a statistically significant number of substantial pK perturbations, mostly due to the induced-fit structural changes, in regions far from the binding interface.  相似文献   

7.
Bajpayee NS  McGrath WJ  Mangel WF 《Biochemistry》2005,44(24):8721-8729
The interactions of the human adenovirus proteinase (AVP) with polymers with high negative charge densities were characterized. AVP utilizes two viral cofactors for maximal enzyme activity (k(cat)/K(m)), the 11-amino acid peptide from the C-terminus of virion precursor protein pVI (pVIc) and the viral DNA. The viral DNA stimulates covalent AVP-pVIc complexes (AVP-pVIc) as a polyanion with a high negative charge density. Here, the interactions of AVP-pVIc with different polymers with high negative charge densities, polymers of glutamic acid (polyE), were characterized. The rate of substrate hydrolysis by AVP-pVIc increased with increasing concentrations of polyE. At higher concentrations of polyE, the increase in the rate of substrate hydrolysis approached saturation. Although glutamic acid did not stimulate enzyme activity, glutamic acid and NaCl could displace DNA from AVP-pVIc.(DNA) complexes; the K(i) values were 230 and 329 nM, respectively. PolyE binds to the DNA binding site on AVP-pVIc as polyE and DNA compete for binding to AVP-pVIc. The equilibrium dissociation constant for 1.3 kDa polyE binding to AVP-pVIc was 56 nM. On average, one molecule of AVP-pVIc binds to 12 residues in polyE. Comparison of polyE and 12-mer single-stranded DNA interacting with AVP-pVIc revealed the binding constants are similar, as are the Michaelis-Menten constants for substrate hydrolysis. The number of ion pairs formed upon the binding of 1.3 kDa polyE to AVP-pVIc was 2, and the nonelectrostatic change in free energy upon binding was -6.5 kcal. These observations may be physiologically relevant as they infer that AVP may bind to proteins that have regions of negative charge density. This would restrict activation of the enzyme to the locus of the cofactor within the cell.  相似文献   

8.
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   

9.
The net charge of a folded protein is hypothesized to influence myriad biochemical processes (e.g., protein misfolding, electron transfer, molecular recognition); however, few tools exist for measuring net charge and this elusive property remains undetermined—at any pH—for nearly all proteins. This study used lysine-acetyl “protein charge ladders” and capillary electrophoresis to measure the net charge of superoxide dismutase-1 (SOD1)—whose aggregation causes amyotrophic lateral sclerosis (ALS)—as a function of coordinated metal ions and pH. The net negative charge of apo-SOD1 was similar to predicted values; however, the binding of a single Zn2 + or Cu2 + ion reduced the net negative charge by a greater magnitude than predicted (i.e., ~ 4 units, instead of 2), whereas the SOD1 protein underwent charge regulation upon binding 2–4 metal ions. From pH5 to pH8 (i.e., a range consistent with the multiple subcellular loci of SOD1), the holo-SOD1 protein underwent smaller fluctuations in net negative charge than predicted (i.e., ~ 3 units, instead of ~ 14) and did not undergo charge inversion at its isoelectric point (pI = 5.3) but remained anionic. The regulation of SOD1 net charge along its pathways of metal binding, and across solvent pH, provides insight into its metal-induced maturation and enzymatic activity (which remains diffusion-limited across pH5–8). The anionic nature of holo-SOD1 across subcellular pH suggests that ~ 45 different ALS-linked mutations to SOD1 will reduce its net negative charge regardless of subcellular localization.  相似文献   

10.
The surface charge of epithelial cells isolated from the toad bladder has been determined by the microscope method of cell electrophoresis. The cells possess a net negative charge, and a net surface charge density of 3.6 x 104 electronic charges per square micron at pH 7.3. Estimates of net surface charge over the alkaline pH range indicate (a) that an average distance of the order of 40 A separates the negatively charged groups, and (b) that amino as well as acid groups are present at the electrophoretic surface of shear. A significant increase in mobility following cyanate treatment of the cells suggests that a large proportion of the amino groups are the ε-amino groups of lysine. In view of the known effects of calcium and other divalent ions on cell permeability and cell adhesion, the extent of binding of calcium and magnesium to the cell surface was determined by the electrophoretic technique. Mobility was significantly decreased in the presence of calcium or magnesium, indicating that these ions are bound by surface groups. When the pH was lowered from 7.3 to 5.2, calcium binding was markedly decreased, an observation consistent with competition between calcium and hydrogen ions for a common receptor site.  相似文献   

11.
The binding constant (K(obs)) for the beta-lactoglobulin-poly(vinylsulfate) (BLG-PVS) complex was measured by frontal analysis continuous capillary electrophoresis at pH values above the isoelectric point of BLG, and the persistence length (L(p)) of PVS was measured by small angle neutron scattering, to examine the effect of polyelectrolyte chain stiffness on its binding efficiency to proteins. The values of K(obs) and L(p) were compared with those of BLG-PSS and BLG-PAMPS (poly(2-acrylamido-2-methylpropanesulfonate)) reported previously. The relationship between K(obs) and L(p) was reciprocal, indicating that protein binding is enhanced by the flexibility of the polyanion, at least in the case where the net protein charge is negative. In addition, at a fixed pH, the polymer systems displayed a similar ionic strength dependence of K(obs). This similarity was consistent with the proposal that the binding properties of PVS and PAMPS polyanions are governed purely by electrostatic interactions and are independent of their molecular structure.  相似文献   

12.
The interaction of two commercially available cellulose-based polycations, Polymer JR-400 and Celquat L200, with polyacrylates (PAA-Na) and polyvinylsulphonates (PVS-Na) of various molecular weights was investigated by covalently labelling the polycations with dansyl hydrazine and then studying the fluorescence polarisation of the dansyl group. Celquat L200 was shown to form complexes that were more stoichiometric than the complexes formed by Polymer JR-400 at pH 6·1. This was attributed to the higher charge density and lower average molecular weight of the Celquat L200. At pH 3·5, no complex formation was observed with any of the samples of PAA-Na; PVS-Na samples did form complexes with the polycations and the ones with Polymer JR-400 were more stoichiometric than the complexes formed at pH 6·1.The critical electrolyte concentration (C.E.C.) of each of the complexes was studied using sodium chloride as the electrolyte. The C.E.C. values for Polymer JR-400 complexes were in the order: PAA-Na (230 000)>PVS-Na (15 000)>PAA-Na (90 000)>PVS-Na (4300)>PAA-Na (5000). For complexes of Celquat L200, the order was: PAA-Na (230 000)>PVS-Na (15 000)=PVS-Na (4300)>PAA-Na (90 000)>PAA-Na (5000). The numerical values of the C.E.C. for complexes of Celquat L200 were found to be greater than the values for complexes of Polymer JR-400, thus implying that Celquat L200 binds more strongly to the polyanions. The order of binding indicated that, for a given molecular weight, the complexes formed by the polyvinylsulphonates are stronger than those formed by the polyacrylates.  相似文献   

13.
The interaction of a set of monovalent (Na+, K+) and divalent (Mg2+, Ca2+) metal cations with single-chain polyguluronate (periodic chain based on a dodecameric repeat unit, 21-helical conformation) is investigated using explicit-solvent molecular dynamics simulations (at 300 K and 1 bar). A total of 14 (neutralising) combinations of the different ions are considered (single type of cation or simultaneous presence of two types of cation, either in the presence or absence of chloride anions). The main observations are: (1) the chain conformation and intramolecular hydrogen bonding is insensitive to the counter-ion environment; (2) the binding of the cations is essentially non-specific for all ions considered (counter-ion atmosphere confined within a cylinder of high ionic density, but no well-defined binding sites); (3) the density and tightness of the distributions of the different cations within the counter-ion atmosphere follow the approximate sequence Ca2+>Mg2+>K+~Na+; (4) the solvent-separated binding of the cations to the carboxylate groups of the chain is frequent, and its occurrence follows the approximate sequence K+>Na+>Ca2+>Mg2+ (contact-binding events as well as the binding of a cation to multiple carboxylate groups are very infrequent); and (5) the counter-ion atmosphere typically leads to a complete screening of the chain charge within 1.0–1.2 nm of the chain axis and, for most systems, to a charge reversal at about 1.5 nm (i.e. the effective chain charge becomes positive at this distance and as high in magnitude as one-quarter of the bare chain charge, before slowly decreasing to zero). These findings agree well (in a qualitative sense) with available experimental data and predictions from simple analytical models, and provide further insight concerning the nature of alginate–cation interactions in aqueous solution.  相似文献   

14.
This report describes the uptake of high molecular weight RNA by Ehrlich ascites tumor cells treated with enzymes and polycations which reduce cell net negative surface charge density. Enzyme treatment had little effect on RNA uptake, but treatment with poly- -lysine resulted in increased binding and uptake of RNA. Present data indicate that decreased cell surface charge, increased availability of positive surface sites, and cell death, all contribute to increased RNA uptake. The individual contributions of these factors has been partially resolved. A possible mechanism for polyanion uptake by cells is proposed.  相似文献   

15.
The mechanism of complexation of pI range 3.5--5 Ampholine to heparin in isoelectric focusing has been explored by the dye-binding technique at different pH values in solution. There is no significant interaction between heparin and Ampholine at pH 6.7. Weak, or selective, binding occurs at pH 5.1, and very strong interaction at pH 3.5. In the latter system, the Ampholine components appear to behave as polycations due to their ordered sequence of positive charges, each two methylene groups apart, which favors a strong binding to polyanions. In addition, there appear to be variable stoichiometries for the strong binding between heparin and Ampholine, depending on their relative amounts. It is proposed that at a low ratio of heparin to Ampholine (Ampholine excess), aggregation is perpendicular to the heparin chain, with the end ammonium charge of each Ampholine molecule neutralizing one negative charge along the heparin molecule; at higher ratios (heparin excess), the bound Ampholine segment is aligned parallel to the heparin molecule, so that on the average one Ampholine component neutralizes approx. three negative charges. The banding of heparin in isoelectric focusing in the pH range 3.0--4.5 can be explained by aggregation of the various components on heparin in amounts dependent upon the net charge on the Ampholine species at the given pH, and upon the changing stoichiometries as a function of the variation in ratio of heparin to Ampholine along the pH gradient. Binding of Ampholine to polygalacturonate was also demonstrated in excess Ampholine in a pH range dependent on the degree of protonation of the carboxyl groups of this acidic polysaccharide as well as on the net positive charge of the Ampholine. The aggregation seen at pH 4.2--4.5 led to the prediction and subsequent demonstration that polygalacturonate would also exhibit binding upon isoelectric focusing. This supports the hypothesis that aggregation of Ampholine on polyanions having sufficient charge density is a general phenomenon which can lead to spurious banding of certain polymers at appropriate pH ranges in isoelectric focusing. On the basis of their behavior in isoelectric focusing at pH 3.0--4.5, strength of aggregation of the polyanions studied appears to be heparin A = heparin B greather than polyglutamate greater than carboxyl-reduced heparin B greater than polygalacturonic acid.  相似文献   

16.
17β-Hydroxysteroid dehydrogenase type 3 (17β-HSD-3) is a member of the short-chain dehydrogenase/reductase (SDR) family and is essential for the reductive conversion of inactive C19-steroid, androstenedione, to the biologically active androgen, testosterone, which plays a central role in the development of the male phenotype. Mutations that inactivate this enzyme give rise to a rare form of male pseudohermaphroditism, referred to as 17β-HSD-3 deficiency. One such mutation is the replacement of arginine at position 80 with glutamine, compromising enzyme activity by increasing the cofactor binding constant 60-fold. In the absence of a 17β-HSD-3 crystal structure, we have grafted its amino acid sequence for the NADPH binding site on the X-ray crystal structures of glutathione reductase (Protein Data Bank code 1gra) and 17β-HSD type 1 (Protein Data Bank codes 1fdv and 1fdu) where we find the trunk of the arginine 80 side chain forms part of the hydrophobic pocket for the purine ring of adenosine while its guanidinium moiety interacts with the 2′-phosphate to both stabilize cofactor binding and neutralize its intrinsic negative charge through two hydrogen bonds. To qualitatively assess the role arginine 80 plays in both selecting and stabilizing NADPH binding, it was replaced with each amino acid and the mutant enzymes subjected to enzymatic analysis. There are only seven enzymes exhibiting any measurable enzymatic activity with arginine~lysine>leucine>glutamine>methionine>tyrosine>isoleucine. With an aspartic acid at position 58 in 17β-HSD-3 occupying the equivalent space in the cofactor binding pocket as arginine 224 in glutathione reductase or serine 12 in 17β-HSD-1, there was an expectation that some of the mutants might use NADH as a cofactor. In no case was NADH found to substitute for NADPH.  相似文献   

17.
Conformaitons of poly(L -arginine)/polyanion complexes were studies by CD measurements. The polyanions were the homoplolypeptides poly(L -glutamic acid) and poly(L -aspartic acid); the synthetic polyelectrolytes and polyethylenesulfonate; and the polynucleotides were native DNA, denatured DNA, and poly(U). It was found that poly(L -arginine) forms the α-helical conformation by interacting with the acidic homopolypeptides and the synthetic anionic polyelectrolytes. In each complex, poly(L -glutamic acid) is in the α-helical conformation, whereas poly(L -aspartic acid) is mostly in the random structure. The poly(L -glutamic acid) complex changed into the β-sheet structure at the transition temperature about 65°C in 0.01M cacodylate buffer (pH 7). Even in the presence of 5M urea, this complex remained in the α-helical conformation at room temperature. The existence of the stable complex of α-helical poly(L -arginine) and α-helical poly(L -glutamic acid) was successfully supported by the model building study of the complex. The α-helix of poly(L -arginine) induced by binding with polyacrylate was the most stable of the poly(L -arginine)-polyanion complexes examined as evidenced by thermal and urea effects. The lower helical content of the polyethylenesulfonate-complexed poly(L -aginine) was explained in terms of the higher charge density of the polyanion. On the other hand, native DNA, denatured DNA, and poly(U) were not effective in stabilizing the helical structure of poly(L -arginine). This may be due to the rigidity of polyanions and to the steric hindrance of bases. Furthermore, the distinitive structual behavior of poly(L -arginine) and poly(L -lysine) regarding polyanion interaction has been noticed throughout the study.  相似文献   

18.
Barnase, the guanine specific ribonuclease of Bacillus amyloliquefaciens, was subjected to mutations in order to alter the electrostatic properties of the enzyme. Ser-85 was mutated into Glu with the goal to introduce an extra charge in the neighborhood of His-102. A double mutation (Ser-85-Glu and Asp-86-Asn) was introduced with the same purpose but without altering the global charge of the enzyme. A similar set of mutations was made using Asp at position 85. For all mutants the pI was determined using the technique of isoelectric focusing and calculated on the basis of the Tanford-Kirkwood theory. When Glu was used to replace Ser-85, the correlation between the experimental and the calculated values was perfect. However, in the Ser-85-Asp mutant, the experimental pI drop is bigger than the calculated one, and in the double mutant (Ser-85-Asp and Asp-86-Asn) the compensation is not achieved. The effect of the mutations on the pKa of His-102 can be determined from the pH dependence of the kcat/KM for the hydrolysis of dinucleotides, e.g., GpC. The effect can also be calculated using the method of Honig. In this case the agreement is very good for the Glu-mutants and the single Asp-mutant, but less for the double Asp-mutant. The global stability of the Asp-mutants is, however, the same as the wild type, as shown by stability studies using urea denaturation. Molecular dynamics calculations, however, show that in the double Asp-mutant His-102 (H+) swings out of its pocket to make a hydrogen bridge with Gln-104 which should cause an additional pKa rise. The effect of the Glu-mutations was also tested on all the kinetic parameters for GpC and the cyclic intermediate G > p at pH 6.5, for RNA at pH 8.0, and for poly(A) at pH 6.2. The effect of the mutations is rather limited for the dinucleotide and the cyclic intermediate, but a strong increase of the KM is observed in the case of the single mutant (extra negative charge) with polymeric substrates. These results indicate that the extra negative charge has a strong destabilizing effect on the binding of the polymeric substrates in the ground state and the transition state complex. A comparison with the structure of bound tetranucleotides (Buckle, A.M. and Fersht, A.R., Biochemistry 33:1644–1653, 1994) shows that the extra negative charge points towards the P2 site.  相似文献   

19.
The effects of different anionic polymers on the kinetic properties of ionic channels formed by neutral gramicidin A (gA) and its positively charged analogs gramicidin-tris(2-aminoethyl)amine (gram-TAEA) and gramicidin-ethylenediamine (gram-EDA) in a bilayer lipid membrane were studied using a method of sensitized photoinactivation. The addition of Konig's polyanion caused substantial deceleration of the photoinactivation kinetics of gram-TAEA channels, which expose three positive charges to the aqueous phase at both sides of the membrane. In contrast, channels formed of gram-EDA, which exposes one positive charge, and neutral gA channels were insensitive to Konig's polyanion. The effect strongly depended on the nature of the polyanion added, namely: DNA, RNA, polyacrylic acid, and polyglutamic acid were inactive, whereas modified polyacrylic acid induced deceleration of the channel kinetics at high concentrations. In addition, DNA was able to prevent the action of Konig's polyanion. In single-channel experiments, the addition of Konig's polyanion resulted in the appearance of long-lived gram-TAEA channels. The deceleration of the gram-TAEA channel kinetics was ascribed to electrostatic interaction of the polyanion with gram-TAEA that reduces the mobility of gram-TAEA monomers and dimers in the membrane via clustering of channels.  相似文献   

20.
The adhesion of Escherchia coli (E. coli) to the colloids of three variable charge soils and its effect on surface charge properties and potassium adsorption of these soil colloids were investigated. The adhesion isotherms of E. coli by soil colloids can be described using the Langmuir equation. The amount of E. coli adhered by the soil colloids varied with soil type and followed the order: Ultisol from Guangxi > Oxisol from Yunnan > Ultisol from Jiangxi. The iron and aluminum oxide contents and CECs of the soils are the important factors affecting the adhesion of E. coli to soil colloids. The relatively lower iron and aluminum oxide contents and higher CEC of the Ultisol from Jiangxi led to the lower adhesion of E. coli to the soil colloids compared to the Ultisol from Guangxi and the Oxisol from Yunnan. The amount of E. coli adhered to the soil colloids decreased with increasing pH, which was consistent with the results predicted from the DLVO theory. E. coli adhesion made the zeta potential of the soil colloids more negative and reduced the isoelectric point of the soil colloids, suggesting that E. coli decreased the surface positive charge and increased negative charge of the soil colloids. In addition, E. coli adhesion increased K+ adsorption by the soil colloids. Therefore, bacterial adhesion improves the fertility of variable charge soils by increasing soil CEC because the CECs of variable charge soils are usually low.  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