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.
Acid dissociation constants of aqueous cyclohexaamylose (6-Cy) and cycloheptaamylose (7-Cy) have been determined at 10–47 and 25–55°C, respectively, by pH potentiometry. Standard enthalpies and entropies of dissociation derived from the temperature dependences of these pKa's are ΔH0 = 8.4 ± 0.3 kcal mol?1, ΔS0 = ?28. ± 1 cal mol?10K?1 for 6-Cy and ΔH0 = 10.0 ± 0.1 kcal mol?1, ΔS0 = ?22.4 ±0.3 cal mol?10K?1 for 7-Cy. Intrinsic 13C nmr resonance displacements of anionic 6- and 7-Cy were measured at 30°C in 5% D2O (vv). These results indicate that the dissociation of 6- and 7-Cy involves both C2 and C3 20-hydroxyl groups. The thermodynamic and nmr parameters are discussed in terms of interglucosyl hydrogen bonding.  相似文献   

2.
The hexahistidine (His6)/nickel(II)-nitrilotriacetic acid (Ni2+-NTA) system is widely used for affinity purification of recombinant proteins. The NTA group has many other applications, including the attachment of chromophores, fluorophores, or nanogold to His6 proteins. Here we explore several applications of the NTA derivative, (Ni2+-NTA)2-Cy3. This molecule binds our two model His6 proteins, N-ethylmaleimide sensitive factor (NSF) and O6-alklyguanine-DNA alkyltransferase (AGT), with moderate affinity (K ∼ 1.5 × 106 M−1) and no effect on their activity. Its high specificity makes (Ni2+-NTA)2-Cy3 ideal for detecting His6 proteins in complex mixtures of other proteins, allowing (Ni2+-NTA)2-Cy3 to be used as a probe in crude cell extracts and as a His6-specific gel stain. (Ni2+-NTA)2-Cy3 binding is reversible in 10 mM ethylenediaminetetraacetic acid (EDTA) or 500 mM imidazole, but in their absence it exchanges slowly (kexchange ∼ 5 × 10−6 s−1 with 0.2 μM labeled protein in the presence of 1 μM His6 peptide). Labeling with (Ni2+-NTA)2-Cy3 allows characterization of hydrodynamic properties by fluorescence anisotropy or analytical ultracentrifugation under conditions that prevent direct detection of protein (e.g., high ADP absorbance). In addition, fluorescence resonance energy transfer (FRET) between (Ni2+-NTA)2-Cy3-labeled proteins and suitable donors/acceptors provides a convenient assay for binding interactions and for measurements of donor-acceptor distances.  相似文献   

3.
Histone deacetylase 8 (HDAC8) was originally classified as a Zn(II)-dependent deacetylase on the basis of Zn(II)-dependent HDAC8 activity in vitro and illumination of a Zn(II) bound to the active site. However, in vitro measurements demonstrated that HDAC8 has higher activity with a bound Fe(II) than Zn(II), although Fe(II)-HDAC8 rapidly loses activity under aerobic conditions. These data suggest that in the cell HDAC8 could be activated by either Zn(II) or Fe(II). Here we detail the kinetics, thermodynamics, and selectivity of Zn(II) and Fe(II) binding to HDAC8. To this end, we have developed a fluorescence anisotropy assay using fluorescein-labeled suberoylanilide hydroxamic acid (fl-SAHA). fl-SAHA binds specifically to metal-bound HDAC8 with affinities comparable to SAHA. To measure the metal affinity of HDAC, metal binding was coupled to fl-SAHA and assayed from the observed change in anisotropy. The metal KD values for HDAC8 are significantly different, ranging from picomolar to micromolar for Zn(II) and Fe(II), respectively. Unexpectedly, the Fe(II) and Zn(II) dissociation rate constants from HDAC8 are comparable, koff ∼0.0006 s−1, suggesting that the apparent association rate constant for Fe(II) is slow (∼3 × 103 M−1 s−1). Furthermore, monovalent cations (K+ or Na+) that bind to HDAC8 decrease the dissociation rate constant of Zn(II) by ≥100-fold for K+ and ≥10-fold for Na+, suggesting a possible mechanism for regulating metal exchange in vivo. The HDAC8 metal affinities are comparable to the readily exchangeable Zn(II) and Fe(II) concentrations in cells, consistent with either or both metal cofactors activating HDAC8.  相似文献   

4.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

5.
The binding of Mn2+ to the anthranilate synthetase-phosphoribosyltransferase enzyme complex from Salmonella typhimurium was examined by electron paramagnetic resonance studies. Two types of binding sites were observed: one to two tight sites with a dissociation constant of 3–5 μm and five to six weaker sites with a dissociation constant of 40–70 μm. The activator constant for Mn2+ was found to be 9 μm for the glutamine-linked anthranilate synthetase activity and 4 μm for the phosphoribosyltransferase activity. These values are both in the range of the dissociation constant for the tight sites. Water proton relaxation rate measurements showed that the binary enhancement values for both classes of sites were equivalent, ?b = 10.7 ± 2.0. The addition of chorismate to the Mn2+-enzyme complexes when predominantly the tight Mn2+ sites were occupied resulted in a large decrease in the observed enhancement (?T = 2.0). Addition of 5-phosphoribosyl-1-pyrophosphate to the enzyme-Mn2+ complexes caused large decreases in the water proton relaxation rate (?T = 1.5) when tight or tight plus weaker Mn2+ sites were occupied. No changes in the water proton relaxation rate were observed when glutamine, pyruvate, or anthranilate were added; a small decrease was observed when enzyme-Mn2+ was titrated with tryptophan. Tryptophan significantly altered the effect of the binding of chorismate but not of 5-phosphoribosyl-1-pyrophosphate. The effect of tryptophan on the water proton relaxation rate of a Mn2+-enzyme-chorismate complex using a variant enzyme complex which is tryptophan hypersensitive (P. D. Robison, and H. R. Levy, 1976, Biochim. Biophys. Acta. 445, 475–485) occurred at lower concentrations than for the normal enzyme complex. The uncomplexed anthranilate synthetase subunit was titrated with Mn2+ and found to have one to two binding sites with a dissociation constant of 300 ± 100 μm. This dissociation constant is much larger than the activator constant for Mn2+ for uncomplexed anthranilate synthetase which was determined to be 4 μm. These results indicate that the Mn2+-binding sites on anthranilate synthetase are altered when the enzyme complex is formed and that both chorismate and 5-phosphoribosyl-1-pyrophosphate interact closely with enzyme-bound Mn2+ or cause a large effect upon its environment.  相似文献   

6.
The dissociation of a series of bovine catalases, in which a proportion of the carboxylic acid groups of glutamic and aspartic acids have been chemically modified by coupling with glycine methyl ester (GME) or ethylenediamine (ED), has been investigated by sedimentation rate and equilibrium methods. Sedimentation equilibrium measurements on GME derivatives have been analysed in terms of a monomer-dimer-trimer- tetramer model. The results show that the association of monomeric (M1) catalase subunits is consistent with the equilibria 4M1?2M2?M4. The Gibbs energies of association at 284K of the monomeric subunit to dimes (M2) and tetramers (M4) were found to be in the range ? 28 to ? 30 kJ mol?1 and ? 91 to ? 97 kJ mol ?1, respectively. The Gibbs energy for association of dimer to tetramer is in the range ? 32 to ? 34 kJ mol?1. Chemical modification of bovine catalase markedly increases its susceptibility to dissociation by sodium n-dodecyl sulphate (SDS) and sedimentation rate measurements suggest that the initial event on addition of SDS is the dissociation of the whole molecule to half-molecules  相似文献   

7.
Maize (Zea mays L.) leaf phosphoenopyruvate (PEP) carboxylase activity at subsaturating levels of PEP was increased by the inclusion of glycerol (20%, v/v) in the assay medium. The extent of activation was dependent on H+ concentration, being more marked at pH 7 (with activities 100% higher than in aqueous medium) than at pH 8 (20% activation). The determination of the substrate concentration necessary to achieve half-maximal enzyme activity (S0.5) (PEP) and maximal velocity (V) between pH 6.9 and 8.2 showed a uniform decrease in S0.5 in the presence of glycerol over the entire pH range tested, and only a slight decrease in V at pH values near 8. Including NaCl (100 millimolar) in the glycerol containing assay medium resulted in additional activation, mainly due to an increase in V over the entire range of pH. Glucose-6-phosphate (5 millimolar) activated both the native and the glycerol-treated enzyme almost to the same extent, at pH 7 and 1 millimolar PEP. Inhibition by 5 millimolar malate at pH 7 and subsaturating PEP was considerably lower in the presence of glycerol than in an aqueous medium (8% against 25%, respectively). Size-exclusion high performance liquid chromatography in aqueous buffer revealed the existence of an equilibrium between the tetrameric and dimeric enzyme forms, which is displaced to the tetramer as the pH was increased from 7 to 8. In the presence of glycerol, only the 400 kilodalton tetrameric form was observed at pH 7 or 8. However, dissociation into dimers by NaCl could not be prevented by the polyol. We conclude that the control of the aggregation state by the metabolic status of the cell could be one regulatory mechanism of PEP carboxylase.  相似文献   

8.
When subjected to alternating electric fields in the frequency range 102-108 Hz, giant lipid vesicles attain oblate, prolate, and spherical shapes and undergo morphological transitions between these shapes as one varies the field frequency and/or the conductivities λin and λex of the aqueous solution inside and outside the vesicles. Four different transitions are observed with characteristic frequencies that depend primarily on the conductivity ratio λin/λex. The theoretical models that have been described in the literature are not able to describe all of these morphological transitions.  相似文献   

9.
Squid giant axons were treated with tetrodotoxin (TTX) in concentrations ranging from 1 nM to 25 nM and the resulting decrease in sodium current was followed in time using the voltage clamp technique. The removal of TTX from the bathing solution produced only partial recovery of the sodium current. This suggests that the over-all interaction is more complex than just a reversible reaction. By correcting for the partial irreversibility of the decrease in sodium current, a dissociation constant of 3.31 x 10-9 M was calculated for the reaction between TTX and the reactive site of the membrane. The data obtained fit a dose-response curve modified to incorporate the correction for partial irreversibility when calculated for a one-to-one stoichiometry. The fit disagreed with that calculated for a reaction between two molecules of TTX with a single membrane-reactive site, but neither supported nor disproved the possibility of a complex formed by two reactive sites with one molecule of TTX. Values of the rate constants for the formation and dissociation of the TTX-membrane complex, k 1 and k 2, respectively, were obtained from the kinetic data. The values are: k 1 = 0.202 x 108 M -1, and k 2 = 0.116 min-1. The magnitude of the dissociation constant derived from these values is 5.74 x 10-9 M, which has the same order of magnitude as that obtained from equilibrium measurements. Arrhenius plots of the rate constants gave values for the thermodynamic quantities of activation.  相似文献   

10.
Using single-molecule force spectroscopy to probe ICAM-1 interactions with recombinant αLβ2 immobilized on microspheres and β2 integrin on neutrophils, we quantified an impressive hierarchy of long-lived, high-strength states of the integrin bond, which start from basal levels with integrin activation in solutions of divalent cations and shift dramatically upward to hyperactivated states with cell signaling in leukocytes. Taking advantage of very rare events, we used repeated measurements of bond lifetimes under steady ramps of force to achieve a direct assay for the off-rates of ICAM-1 from β2 integrin in each experiment. Of fundamental importance, the assay for off-rates does not depend on how the force is applied over time, and remains valid when the rates of dissociation change with different levels of force. In this first article, we present results from tests of a monovalent ICAM-1 probe against immobilized αLβ2 in environments of divalent cations (Ca2+, Mg2+, and Mn2+) and demonstrate in detail the method for assay of off-rates. When extrapolated to zero force, the force-free values for the off-rates are found to be consistent with published solution-based assays of soluble ICAM-1 dissociation from immobilized LFA-1, i.e., ∼10−2/s in Mg2+ or Mn2+ and ∼1/s in Ca2+. At the same time, as expected for adhesive function, we find that the β2 integrin bonds activated in Mn2+ or Mg2+ possess significant and persistent mechanical strength (e.g., >20 pN for >1 s) even when subjected to slow force ramps (<10 pN/s). As discussed in our companion article, using the same assay, we find that although the rates of dissociation for diICAM-1fc bonds to LFA-1 on neutrophils in Mn2+ are similar to those for mICAM-1 bonds to recombinant αLβ2 on microspheres, they appear to represent a dimeric attachment to a pair of tightly clustered integrin heterodimers. The mechanical strengths and lifetimes of the dimeric interactions increase dramatically when the neutrophils are stimulated by the chemokine IL-8 or are bound with an allosterically activating (anti-CD18) monoclonal antibody, demonstrating the major impact of cell signaling on LFA-1.  相似文献   

11.
Electronic absorption and emission spectra, along with lifetime measurements and vibrational spectra, are used to investigate the interaction between nitrate and trivalent europium ions in dilute solutions in anhydrous and aqueous acetonitrile. Upon addition of increasing quantities of nitrate, the complexes [Eu(NO3)n](3?n)+, with n = 1–5, form quantitatively in anhydrous acetonitrile. In solution, the pentanitrato species is not further solvated and its spectroscopic properties are similar to those of solid samples, indicating a similar structure with five bidentate nitrates bonded to the 10-coordinate Eu(III) ion. The lifetimes of the 5D0 level are 1.35(5) and 1.25(5) ms for Eu(NO3)3 and (Me4N)2Eu(NO3)5 0.05 M in CH3CN. The quantum yield of Eu(NO3)3 in CH3CN is 27.4%.The addition of small quantities of water to Eu(NO3)3 solutions does not result in the dissociation of the nitrate ions, provided Rw = [H2O]t/[Eu3+]t is smaller than 8; the apparent equilibrium rations for [Eu(NO3)3(H2O)n] are K3 = 40 ± 15 M?1 and K4 = 9 ± 3 M?1; K1 and K2 are too large to be determined. The formation of nitrato complexes is studied in mixtures containing increasing amounts of water and nitrate. Deconvolution of the different components of the 5D07F0 transition allows a semi-quantitative estimate of the relative concentration of the nitrato complexes. The total number of coordinated nitrate ions per europium ion can be determined on the basis of fluorescence lifetime measurements. The apparent equilibrium ratios for the formation of the mono- and dinitrato species amount to K1 = 23 ± 3, 15 ± 5 and 5 ± 1 for Rw = 44, 94 and 304, respectively, and to K2 = 17 ± 8 for Rw = 44 and 94.  相似文献   

12.
We have previously shown that methionine–heme iron coordination is perturbed in domain-swapped dimeric horse cytochrome c. To gain insight into the effect of methionine dissociation in dimeric cytochrome c, we investigated its interaction with cyanide ion. We found that the Soret and Q bands of oxidized dimeric cytochrome c at 406.5 and 529 nm redshift to 413 and 536 nm, respectively, on addition of 1 mM cyanide ion. The binding constant of dimeric cytochrome c and cyanide ion was obtained as 2.5 × 104 M?1. The Fe–CN and C–N stretching (ν Fe–CN and ν CN) resonance Raman bands of CN?-bound dimeric cytochrome c were observed at 443 and 2,126 cm?1, respectively. The ν Fe–CN frequency of dimeric cytochrome c was relatively low compared with that of other CN?-bound heme proteins, and a relatively strong coupling between the Fe–C–N bending and porphyrin vibrations was observed in the 350–450-cm?1 region. The low ν Fe–CN frequency suggests weaker binding of the cyanide ion to dimeric cytochrome c compared with other heme proteins possessing a distal heme cavity. Although the secondary structure of dimeric cytochrome c did not change on addition of cyanide ion according to circular dichroism measurements, the dimer dissociation rate at 45 °C increased from (8.9 ± 0.7) × 10?6 to (3.8 ± 0.2) × 10?5 s?1, with a decrease of about 2 °C in its dissociation temperature obtained with differential scanning calorimetry. The results show that diatomic ligands may bind to the heme iron of dimeric cytochrome c and affect its stability.  相似文献   

13.
Jay Sung 《Inorganica chimica acta》2006,359(12):3888-3894
The kinetic studies of the addition reactions of pentacyanonitrosylferrate(2−) (NP) with acidic methylene ligands of the form CH2LL′ (L, L′ = -CN, -CONH2; -CN, -CN; CH3CO-, CH3CO-) have been carried out in basic solutions. The increase in the second order rate constants for the formation of Fe(CN)5N(O)CLL′4− complexes with the increase of OH concentration indicated that ?CHLL′ is the reactive species of the reactions. In addition to the pKa, the lability of the methylene protons of the ligands in aqueous solution also governs the reactivity of the reactions. The nitrosation products were rather unstable with respect to the dissociation into and oxime compounds. The kinetic measurements showed that the rates of dissociation were first order and were rather insensitive to the hydroxide concentration for the ligands under study.  相似文献   

14.
15.
Human growth hormone binding sites from female rabbit kidney microsomes were solubilized by treatment with the nonionic detergent Triton X-100. The binding of 125I-labelled human growth hormone to the solubilized sites retains many of the properties observed in the particulate fraction, such as saturability, reversibility, high affinity and structural specificity. The association and the dissociation process are time- and temperature-dependent. The association rate constant, k1, is 1.6·107 mol?1·l·min?1 at 25°C, and the dissociation rate constant, k?1, is 2.8·10?4 min?1 at 25°C. Solubilization causes an increase in affinity as well as in binding capacity. Scatchard plots from saturation curves suggest the presence of a single class of binding site with a dissociation equilibrium constant, Kd, of 1.3·10?11 M and a binding capacity of 133 fmol/mg of protein. Similar results were obtained from competition experiments. Specificity studies revealed the lactogenic characteristics of the solubilized sites. The Stokes radii of the free binding sites and of the 125I-labelled human growth hormone-binding site complex, determined on a Sepharose CL-6B column, are 57 and 53 Å, respectively.  相似文献   

16.
B Attardi  S Ono 《Cell》1974,2(4):205-212
Steroid binding has been studied in cytoplasmic extracts of normal mouse kidney, an androgen sensitive organ, and of kidney from mice affected with testicular feminization (Tfm mutant) that have inherited androgen resistance. Macromolecules that bind 3H-5α-dihydrotestosterone (DHT, the presumed active androgen in most testosterone target organs) and sediment in glycerol gradient at 8–9S can be observed in cytosol from kidney of mice of different sex, age, and hormonal history. The 8–9S component from normal females is heat labile, pronase sensitive, and dissociated by high salt to a lower molecular weight entity. The apparent equilibrium dissociation constant (Kd) for the DHT-receptor complex is 1.4 × 10?9M, and there are about 1500 binding sites per testosterone-sensitive kidney proximal tubule cell. Cytosol from Tfm/Y animals also shows a sharp peak of 3H-DHT-binding activity at 8–9S. The Tfm protein, however, has reduced affinity for DHT and binds only 10–25% as much 3H-DHT as wild-type receptor at 3H-DHT concentrations from 5 × 10?11M to 1.2 × 10?8M. Scatchard analysis, and studies involving competition with unlabeled steroids, relative binding of various androgens, and dissociation of the 3H-DHT-binding protein complex after extensive dialysis have led to the conclusion that Tfm kidney contains very little, if any, androgen receptor with properties similar to that found in normal kidney.  相似文献   

17.
Using single-molecule force spectroscopy to probe ICAM-1 interactions with recombinant αLβ2 immobilized on microspheres and β2 integrin on neutrophils, we quantified an impressive hierarchy of long-lived, high-strength states of the integrin bond, which start from basal levels with activation in solutions of divalent cations and shift dramatically upward to hyperactivated states with cell signaling. Taking advantage of very rare events, we used repeated measurements of bond lifetimes under steady ramps of force to achieve a direct assay for the off-rates of ICAM-1 from β2 integrin throughout the course of each experiment. In our companion article I, we demonstrate the assay using results from tests of a monovalent ICAM-1 probe against recombinant αLβ2 on microspheres in millimolar solutions of divalent cations (Ca2+, Mg2+, Mn2+). In this article, we examine the impact of inside-out and outside-in signaling in neutrophils on the lifetimes and mechanical strengths of ICAM-1 bonds to β2 integrin on the cell surface. Even though ICAM-1 bonds to recombinant αLβ2 on microspheres in Mg2+ or Mn2+ can live for long periods of time under slow pulling, here we show that stimulation of neutrophils in Mg2+ plus the chemokine IL-8 (i.e., inside-out signaling) induces several-hundred-fold longer lifetimes for ICAM-1 attachments to LFA-1, creating strong bonds at very slow pulling speeds where none are perceived in Mg2+ or Mn2+ alone. Similar changes are observed with outside-in signaling, i.e., long lifetimes and increased bond strength also occur when neutrophils are bound with the activating (anti-CD18) monoclonal 240Q. Limiting our investigation to rare events using very dilute ICAM-1 probes, we show that although the prolonged lifetimes of cell surface attachments for both inside-out and outside-in signaling exhibit single-bond-like statistics for dissociation under force, they are consistent with a tightly coupled dimeric ICAM-1 interaction with a pair of LFA-1 heterodimers.  相似文献   

18.
The unidirectional uptake of sodium across the outer surface of the isolated frog skin (J12Na) was measured in the presence of transepithelial potential difference (Δψ) ranging from +100 to -100 mV. With a sodium concentration of 115 mM in the bathing solutions J12Na increases significantly when the spontaneous Δψ is reduced to zero by short-circuiting the skin. With an Na concentration of 6 mM a progressive increase J12Na can be observed when Δψ is decreased in several steps from +100 to -100 mV (serosal side positive and negative, respectively). The observed change J12Na amounts to a fraction only of that predicted from the shift in Δψ. The results suggest that under open circuit conditions the potential step across the outside surface is at most one half of Δψ and that the resistance across the outside and inside barrier of the skin is ohmic. This is in agreement with measurements of intracellular potentials in the frog skin and with resistance measurements carried out in the toad skin. The data strongly support the view that the saturating component of Jψ proceeds via a charged carrier system. Exposure to negative values of Δψ of 50 mV or more for times of 24 min or more result in a marked reduction of J12Na which shows only partial or no reversibility.  相似文献   

19.
Ultrathin (black) lipid membranes were made from sheep red cell lipids dissolved in n-decane. The presence of aliphatic alcohols in the aqueous solutions bathing these membranes produced reversible changes in the ionic permeability, but not the osomotic permeability. Heptanol (8 mM), for example, caused the membrane resistance (Rm) to decrease from >108 to about 105 ohm-cm2 and caused a marked increase in the permeability to cations, especially potassium. In terms of ionic transference numbers, deduced from measurements of the membrane potential at zero current, T cat/T Cl increased from about 6 to 21 and T K/T Na increased from about 3 to 21. The addition of long-chain (C8ndash;C10) alcohols to the lipid solutions from which membranes were made produced similar effects on the ionic permeability. A plot of log Rm vs. log alcohol concentration was linear over the range of maximum change in Rm, and the slope was -3 to -5 for C2 through C7 alcohols, suggesting that a complex of several alcohol molecules is responsible for the increase in ionic permeability. Membrane permselectivity changed from cationic to anionic when thorium or ferric iron (10-4 M) was present in the aqueous phase or when a secondary amine (Amberlite LA-2) was added to the lipid solutions from which membranes were made. When membranes containing the secondary amine were exposed to heptanol, Rm became very low (103–104 ohm-cm2) and the membranes became perfectly anion-selective, developing chloride diffusion potentials up to 150 mv.  相似文献   

20.
Low concentrations of Mg2+ (concn < 10 mm) generate structural changes in delipidated spinach chloroplast lamellae, that appear as changes in the fluorescence yield of native tryptophyl residues and of the externally added polarity probe magnesium 1-anilinonaphthalene-8-sulfonate.The delipidated lamellae, consisting essentially of structural protein monomers and aggregates, bind magnesium 1-anilinonaphthalene-8-sulfonate to the extent of 126 ± 13 nmol/mg protein, and with a dissociation constant KD = 167 μM. Bound ANS fluoresces at 458 nm with a quantum yield Φ = 0.121. Tryptophyls sensitize the fluorescence of bound ANS with a maximal efficiency Tmax = 0.85. Assuming completely random orientation of the interacting chromophores, an interchromophore separation R = 17.3 A? is calculated. Only two-thirds of the membrane tryptophyls have ANS-binding sites in their vicinity.Mg2+ binds to the delipidated membranes with a dissociation constant KD = 2 mM. The binding is attended by enhancement of magnesium 1-anilinonaphthalene-8-sulfonate fluorescence, and deenhancement of tryptophyl fluorescence, while the efficiency of interchromophore excitation transfer increases only slightly. These effects suggest that Mg2+ generates a structural change which lowers the polarity of the membrane region where tryptophyl and magnesium 1-anilinonaphthalene-8-sulfonate are situated, but which has a minor effect only on the interchromophore separation.  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