首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The general three-state model is formulated first, which is the direct extension of the unified two-state model previously formulated (Kijima & Kijima, 1978). In this model, each protomer in a symmetrically interacting system (oligomers or lattices) can take three states, S, R and Q, where S and R states are the same as in the two-state model and Q state is another state either corresponding to a different open-state of ionophore from R open-state or corresponding to another closed state of ionophore. The model has no restriction on the value of Hill coefficient at the midpoint of the dose-response curves in contrast to two-state models. It is applied on GABA sensitive inhibitory synapse of crayfish muscle to account for anomalous behaviour of the membrane in I? solution.The simplified versions of the above general three-state model are also formulated (simplified three-state model), in which it is assumed that R and Q state are equivalent in regard to the nearest neighbor interaction. By this assumption, R and Q state are collectively treated as state A and mathematical formula obtained on Ising model are applicable on this model. This model is applied on the insect sugar receptor which was shown to be incompatible with the two-state models (Kijima & Kijima, 1980). Further simplification of the above simplified model results in two convenient models: three-state KNF model and three-state MWC model, which have minimum parameters but sufficient to account for most experiments. They give plausible physico-chemical base on the “classical model” in which the existence of both inactive and active ligand-receptor complex is assumed.  相似文献   

2.
The enthalpies of the hexokinase-catalyzed phosphorylation or glucose, mannose, and fructose by ATP to the respective hexose 6-phosphates have been measured calorimetrically in TRIS/TRIS HCl buffer at 25.0, 28.5, and 32.0°C. The effects on the measured enthalpy of the glucose/hexokinase reaction due to variation of pH (over the range 6.7 to 9.0) and ionic strength (over the range 0.02 to 0.25) have been examined. Correction for enthalpy of buffer protonation leads to δHo and δCpo values for the processes: eq-D-hexose + ATP4− = eq-D-hexose 6-phosphate2− + ADP3−+ H+. Results are δHo = −23.8 ± 0.7 kJ · mol−1 and δCpo = −156 ± 280 J·mol−1·K−1 for glucose. δHo = −21.9 ± 0.7 kJ·mol−1 and δCpo = 10 ± 140 J·mol−1·K−1 for mannose, and δHo = −15.0 ± 0.9 kJ·mol−1 and δCpo = −41 ± 160 J·mol−1·K−1 for fructose. Combination of these measured enthalpies with Gibbs energy data for hydrolysis of ATP4− and that for the hexose 6-phosphates lead to δSo values for the above hexokinase-catalyzed reactions.  相似文献   

3.
The interaction of |CnH2n+1N+(CH3)3| · I? (n = 3, 6, 9, 12, 14, 16 or 18) with egg-yolk phosphatidylcholine-water dispersions has been studied by 31P-NMR spectroscopy. It is shown that the effective anisotropy of 31P chemical shift (?Δσeff) of the lamellar phospholipid liquid-crystalline phase Lα increases with increasing concentration and alkyl chain length of the drug. Addition of |C6H13N+(CH3)3| ·I ? or |C9H19N+(CH3)3I? to the phospholipid-water dispersion at a molar ratio ammonium salt:phospholipid > 0.8 induces in the dispersion a structure with an effective isotropic phospholipid motion. This structure is unstable and slowly transforms into the hexagonal phase. These effects have not been observed in phospholipid-water dispersions mixed with the ammonium derivatives with the longer alkyl chains n  12, 14, 16 or 18. It is proposed that these results might explain the effects of the investigated drugs on the nerve, muscle and bacterial cells.  相似文献   

4.
The respiratory properties of the whole blood of the burrowing red band fish Cepola rubescens L. were investigated. Oxygen dissociation curves constructed at 15°C were found to be close to hyperbolic in shape with a mean value for the cooperativity coefficient at half-saturation (n50) of 1.56. Half-saturation oxygen tension (P50) for pH = 7.56 (mean in vivo pH of venous blood) was 27 Torr. The blood showed a marked Bohr effect (Δ log P50ΔpH = ?1.19) and also a Root effect which at the in vivo pH reduced oxygen carrying capacity by 20%. The PvCO2 was 3.2 Torr and the buffering power of the blood was low, the buffer value of true plasma averaging 5.43 mmol · 1?1 · pH?1. It is suggested that the large Bohr effect coupled with the low buffer value confers on the haemoglobin a flexibility, in terms of oxygen affinity, to withstand changes which occur in environmental oxygen tensions.  相似文献   

5.
The physicochemical properties of a high-molecular-weight spin-labeled nucleic acid, (RUGT,U)n, synthesized by enzymatic copolymerization, were evaluated by uv and ESR spectroscopy. It was shown earlier that spin labeling of nucleic acids by chemical modification to an extent which gives a nitroxide-to-nucleotide ratio greater than 0.002 can cause noticeable lattice perturbations (A. M. Bobst, A. Hakam, P. W. Langemeier, and S. Kouidou (1979), Arch. Biochem. Biophys. 187, 339–345). The presence of RUGT, a 5-nitroxide-labeled uridine residue, in a (U)n lattice at a RUGTU ratio of 0.01 is shown here not to affect the complexation with (A)n, since the uv melting temperature (T0OD) of the 2 → 1 transition and the hypochromicity changes were the same for (RUGT,U)n· (A)n and (U)n·(A)n. ESR measurements indicated that the nitroxide radical reflects the transition accurately within the error limit, although a slight destabilization of the spinlabeled segment could not be excluded. Computer simulations showed conclusively that the spin melting temperature (Tmsp) corresponds to the temperature at which half of the spin-labeled segments are no longer complexed, for the ESR spectrum at Tmspcan be simulated with equal contributions from the line shapes of ESR spectra taken before and after the transition. Arrhenius plots obtained by using two different approaches for computing correlation times were qualitatively the same. Computer analysis also revealed that the formation of a (RUGT,U)n·(A)n complex can be described by a two-state model, in contrast to results obtained with chemically spin-labeled (U)n. Thus, using (RUGT,U)n over chemically spin-labeled (U)n can offer distinct advantages.  相似文献   

6.
New complexes of the general formulae Co(o-LH)2X2 (XCl, NCS), Co(o-LH)2Br2·EtOH (EtOHethanol), M(o-LH)(NO3)2 (MCo, Ni), Ni(o-LH)2X2 (XCl, Br, NCS), Cu(o-L)X (XCl, Br), Zn(o-LH)X2 (XCl, Br), Pd(o-L)Cl, Pt(o-LH)2Cl2·H2O, M(m-LH)Cl2·nH2O (MCo, Ni, Pd; n=0, 0.5, 1), Cu(m-LH)Cl2·EtOH, M(m-LH)2Cl2·nH2O (MCo, Zn, Pt; n=0, 1), M(m-LH)Br2 (MCu, Zn), M(m-LH)2Br2 (MCo, Ni), Co(m-LH)(NCS)2 and Co(m-LH)2(NCS)2, where o-LH=N-(2-aminophenyl)quinoline-2′-carboxamide and m-LH=N-(3-aminophenyl)quinoline-2′-carboxamide, have been prepared. The complexes were characterised by elemental analyses, conductivity measurements, X-ray powder patterns, thermogravimetric analyses, magnetic moments and spectral (1H NMR, IR, and electronic) studies. Copper(II) and palladium(II) promote amide deprotonation at nearly acidic pH on coordination with o-LH. A variety of stereochemistries is assigned for the complexes prepared. The deprotonated copper(II) and the nickel(II) and palladium(II) complexes of m-LH appear to be polymeric. The neutral amide group of the ligands is coordinated to the metal ions through oxygen, while N(amide)-coordination is observed for the deprotonated complexes. Coordination of the secondary amide group is not observed for Zn(m-LH)2Cl2, Pd(m-LH)Cl2·0.5H2O and platinum(II) complexes. The neutral ligand o-LH shows bidentate N(ring), O-behaviour, while the anion o-L exhibits tridentate N,N,N-coordination. m-LH acts as a monodentate, bidentate and tridentate ligand depending on the metal ion, the anion and the preparative conditions.  相似文献   

7.
The antibacterial activity and surfactant activity of the compounds trans-[Rh(L)4Cl2]Cl·nH2O increase in the order L = pyridine<4-methylpyridine<4-ethylpyridine<4-n-propylpyridine.As surfactants, the compounds are far more effective at reducing the interfacial tension, n-hexadecaneH2O, than the surface tension, H2Oair.The most effective and efficient surfactant in this series, trans-[Rh(4-n-propylpyridine)4Cl2]Cl·H2O, can cause the leakage of intracellular manganese ions from the gram-positive bacteria, Bacillus brevis ATCC 9999, at a concentration of 130 ppm but there is no observable effect on the retention of intracellular manganese ions at the minimum concentration required to prevent growth of this organism (~0.6 ppm at 23°C in nutrient broth).At 130 ppm, trans-[Rh(4-n-propylpyridine)4Cl2]Cl·H2O does not cause the loss of intracellular manganese ions from the gram-negative bacteria, Escherichia coli JS-1. In this case, a concentration of at least 63 ppm of this rhodium compound is required to prevent the growth of this organism in M9TUH medium at 35°C.On the basis of these results, it is suggested that gross membrane disruption effects caused by the surfactants trans-[Rh(L)4Cl2]Cl·nH2O are not directly responsible for their observed antibacterial action.  相似文献   

8.
The enthalpy change for the oxidative deamination of glutamate by NADP+ catalyzed by bovine liver glutamate dehydrogenase has been determined calorimetrically. The ΔHo values are 64.6 ± 1.2 kJmol and 70.3 ± 1.2 kJmol at 25 and 35°C respectively. The equilibrium constants for the reaction at the two temperatures were determined spectrophotometrically. This enabled the determination of ΔGo and ΔSo of the reaction as well. ΔHovalues were also determined for the reaction using an alternative coenzyme and the deuterated substrate.  相似文献   

9.
A thermodynamic characterization of the Na+-H+ exchange system in Halobacterium halobium was carried out by evaluating the relevant phenomenological parameters derived from potential-jump measurements. The experiments were performed with sub-bacterial particles devoid of the purple membrane, in 1 M NaCl, 2 M KCl, and at pH 6.5–7.0. Jumps in either pH or pNa were brought about in the external medium, at zero electric potential difference across the membrane, and the resulting relaxation kinetics of protons and sodium flows were measured. It was found that the relaxation kinetics of the proton flow caused by a pH-jump follow a single exponential decay, and that the relaxation kinetics of both the proton and the sodium flows caused by a pNa-jump also follow single exponential decay patterns. In addition, it was found that the decay constants for the proton flow caused by a pH-jump and a pNa-jump have the same numerical value. The physical meaning of the decay constants has been elucidated in terms of the phenomenological coefficients (mobilities) and the buffering capacities of the system. The phenomenological coefficients for the Na+-H+ flows were determined as differential quantities. The value obtained for the total proton permeability through the particle membrane via all available channels, LH = (?JH +pH)Δψ,ΔpNa, was in the range of 850–1150 nmol H+·(mg protein)?1·h?1·(pH unit)?1 for four different preparations; for the total Na+ permeability, LNa = (?JNa+pNa)Δψ,ΔpH, it was 1620–2500 nmol Na+·(mg protein)?1·h?1·(pNa unit)?1; and for the proton ‘cross-permeability’, LHNa = (?JH+pNa)Δψ,ΔpH, it was 220–580 nmol H+·(mg protein)?1·h?1·(pNa unit)?1, for different preparations. From the above phenomenological parameters, the following quantities have been calculated: the degree of coupling (q), the maximal efficiency of Na+-H+ exchange (ηmax), the flow and force efficacies (?) of the above exchange, and the admissible range for the values of the molecular stoichiometry parameter (r). We found q ? 0.4; ηmax ? 5%; 0.36 ? r ? 2; ?JNa+ ? 1.3 · 105μmol · (RT unit)?1 at JNa = 1 μmolNa+ · (mgprotein)?1 · h?1; and ?ΔpNa ? 5 · 104 ΔpNa · (mg protein) · h · (RT unit)?1 at ΔpNa = 1 unit, for different preparations.  相似文献   

10.
《Inorganica chimica acta》1986,120(2):177-184
The dark blue dimeric complex di-μ-hydroxo- bis [(1,4,7,10-tetraazacyclododecane)chromium(III)] dithionate tetrahydrate, [Cr(C8H20N4)OH]2(S2O6)2· 4H2O or [Cr(cyclen)OH]2(S2O6)2·4H2O, has been synthesized. The crystal structure of the complex has been determined from threeodimensional counter X-ray data. The complex crystallizes in space group P21/n of the monoclinic system with two dimeric formula units in a cell of dimensions a = 8.837(5), b = 14.472(8), c= 13.943(6) Å andβ=95.83(4)o. The structure has been refined by full-matrix least- squares methods to a final value of the weighted R-factor of 0.059 on the basis of 1774 independent intensities. The geometry of the cyclen macrocycle is unsymmetrical, the observed conformations being λδδλ and its enantiomer. The strained ligand conformation leads to significant deviations from octahedral geometry at the chromium centers, and to a bridged geometry in which the CrOCr angle ø and the Cr···Cr separation of 104.1(1)o and 3.086(2) Å are the largest observed in dimers of this kind. The magnetic susceptibility of the complex indicates antiferromagnetic coupling, with the ground state singlet lying 21.56(6) cm−1 below the lowest lying triplet state. The structural parameters have been used to calculate the triplet energy by means of the Glerup- Hodgson- Pedersen (GHP) model, and the calculated value of 22.3 cme−1 is very similar to the observed value.  相似文献   

11.
Thermo-transient receptor potential channels display outstanding temperature sensitivity and can be directly gated by low or high temperature, giving rise to cold- and heat-activated currents. These constitute the molecular basis for the detection of changes in ambient temperature by sensory neurons in animals. The mechanism that underlies the temperature sensitivity in thermo-transient receptor potential channels remains unknown, but has been associated with large changes in standard-state enthalpy (ΔHo) and entropy (ΔSo) upon channel gating. The magnitude, sign, and temperature dependence of ΔHo and ΔSo, the last given by an associated change in heat capacity (ΔCp), can determine a channel’s temperature sensitivity and whether it is activated by cooling, heating, or both, if ΔCp makes an important contribution. We show that in the presence of allosteric gating, other parameters, besides ΔHo and ΔSo, including the gating equilibrium constant, the strength- and temperature dependence of the coupling between gating and the temperature-sensitive transitions, as well as the ΔHo/ΔSo ratio associated with them, can also determine a channel’s temperature-dependent activity, and even give rise to channels that respond to both cooling and heating in a ΔCp-independent manner.  相似文献   

12.
There are five oxidation-reduction states of horseradish peroxidase which are interconvertible. These states are ferrous, ferric, Compound II (ferryl), Compound I (primary compound of peroxidase and H2O2), and Compound III (oxy-ferrous). The presence of heme-linked ionization groups was confirmed in the ferrous enzyme by spectrophotometric and pH stat titration experiments. The values of pK were 5.87 for isoenzyme A and 7.17 for isoenzymes (B + C). The proton was released when the ferrous enzyme was oxidized to the ferric enzyme while the uptake of the proton occurred when the ferrous enzyme reacted with oxygen to form Compound III. The results could be explained by assuming that the heme-linked ionization group is in the vicinity of the sixth ligand and forms a stable hydrogen bond with the ligand.The measurements of uptake and release of protons in various reactions also yielded the following stoichiometries: Ferric peroxidase + H2O2 → Compound I, Compound I + e? + H+ → Compound II, Compound II + e? + H+ → ferric peroxidase, Compound II + H2O2 → Compound III, Compound III + 3e? + 3H+ → ferric peroxidase.Based on the above stoichiometries and assuming the interaction between the sixth ligand and heme-linked ionization group of the protein, it was possible to picture simple models showing structural relations between five oxidation-reduction states of peroxidase. Tentative formulae are as follows: [Pr·Po·Fe-(II) $?PrH+·Po·Fe(II)] is for the ferrous enzyme, Pr·Po·Fe(III)OH2 for the ferric one, Pr·Po·Fe(IV)OH? for Compound II, Pr(OH?)·Po+·Fe(IV)OH? for Compound I, and PrH+·Po·Fe(III)O2? for Compound III, in which Pr stands for protein and Po for porphyrin. And by Fe(IV)OH?, for instance, is meant that OH? is coordinated at the sixth position of the heme iron and the formal oxidation state of the iron is four.  相似文献   

13.
Proton-gated TASK-3 K+ channel belongs to the K2P family of proteins that underlie the K+ leak setting the membrane potential in all cells. TASK-3 is under cooperative gating control by extracellular [H+]. Use of recently solved K2P structures allows us to explore the molecular mechanism of TASK-3 cooperative pH gating. Tunnel-like side portals define an extracellular ion pathway to the selectivity filter. We use a combination of molecular modeling and functional assays to show that pH-sensing histidine residues and K+ ions mutually interact electrostatically in the confines of the extracellular ion pathway. K+ ions modulate the pKa of sensing histidine side chains whose charge states in turn determine the open/closed transition of the channel pore. Cooperativity, and therefore steep dependence of TASK-3 K+ channel activity on extracellular pH, is dependent on an effect of the permeant ion on the channel pHo sensors.  相似文献   

14.
R.L. Pan  S. Izawa 《BBA》1979,547(2):311-319
NH2OH-treated, non-water-splitting chloroplasts can oxidize H2O2 to O2 through Photosystem II at substantial rates (100–250 μequiv · h?1 · mg?1 chlorophyll with 5 mM H2O2) using 2,5-dimethyl-p-benzoquinone as an electron acceptor in the presence of the plastoquinone antagonist dibromothymoquinone. This H2O2 → Photosystem II → dimethylquinone reaction supports phosphorylation with a Pe2 ratio of 0.25–0.35 and proton uptake with H+e values of 0.67 (pH 8)–0.85 (pH 6). These are close to the Pe2 value of 0.3–0.38 and the H+e values of 0.7–0.93 found in parallel experiments for the H2O → Photosystem II → dimethylquinone reaction in untreated chloroplasts. Semi-quantitative data are also presented which show that the donor → Photosystem II → dibromothymoquinone (→O2) reaction can support phosphorylation when the donor used is a proton-releasing reductant (benzidine, catechol) but not when it is a non-proton carrier (I?, ferrocyanide).  相似文献   

15.
16.
The macromolecular structural transition of Pf1 filamentous bacterial virus detected by X-ray diffraction analysis has been studied in virus solutions by density, circular dichroism, and microcalorimetric measurements. The reversible structural change occurring between 5 °C and 25 °C has a calorimetrically determined transition enthalpy ΔHt,cal of 14·5 ± 1.5 kJ (mol protein)?1. The transition curves resulting from the density, circular dichroism, and calorimetric measurements have been analysed in terms of a two-state process to extract the van't Hoff enthalpy. Comparison of the effective transition enthalpy and the calorimetric ΔHt,cal values gives about 26 protein subunits as the size of the co-operative unit. Parallel heat capacity and density measurements on fd virus show no such transition, in agreement with X-ray diffraction studies.  相似文献   

17.
The micellar properties of gangliosides in water solutions were investigated by quasielastic light scattering measurements. GM1 and GD1a gangliosides were isolated from calf brain, purified to more than 99% and dissolved in 0.025 M Tris—HCI buffer (pH 6.8) at 37°C. The average intensity of scattered light and the intensity correlation function were measured by an apparatus including a 5145 Å argon laser and a real-time digital correlator. The scattered intensity data allowed the derivation of an upper limit to the critical micelle concentration (c0) and the evaluation of the molecular weight (M) of the micelle. The intensity correlation function gave the diffusion coefficient D, and hence the hydrodynamic radius RH, and also contained information on the polydispersity of the sample. We find co < 1 × 10?6 M for both GM1 and GD1a, M = 532 000 ± 50 000 and RH = 63.9 ± 2 A? for GM1, and M = 417 000 ± 40 000 and RH = 59.5 ± 2 A? for GD1a. The mixture 3:1 of the two gangliosides gave intermediate values for all examined parameters. The presence of cations, within the physiological concentration range. and, in particular of Ca2+, did not influence significantly the values of co and the main features of the micelle.  相似文献   

18.
A method was developed for quantitative determination of 5α,7α-dihydroxy-11-ketotetranor-prostane-1,16-dioic acid, the major urinary metabolite of prostaglandins F and F in man. The method was based on the use of the O-methyloxime derivative of [5β-3H; 10,10,12,12-2H4]5α,7α-dihydroxy-11-ketotetranor-prostane-1,16-dioic acid as internal standard and determination of ratios between unlabeled and deuterium-labeled molecules by multiple-ion analysis. Excretion values found for healthy human subjects were: males, 10.8–59.0 μg24 hr (n = 10, mean value, 24.0 ± 17.2 (SD) μg) and females, 7.6–13.6 μg24 hr (n = 10, mean value, 10.5 ± 2.1 (SD) μg).  相似文献   

19.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

20.
Cyclooctaamylose crystallizes from aqueous solution with space-group symmetry P21 and lattice parameters: a = 20.253(8), b = 10.494(5), c = 16.892(6) A and β = 105.32(1)o, Z = 2; the apparent formular per asymmetric unit is C48H80O40·17H2O. The macrocycle is in an open conformation but displays significant deviations from ideal eight fold molecular symmetry. Of the 19 water molecules thus far located, four of which have occupancy factors of one half, 12 may be characterized as being in the torus of the cycloamylose.  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