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.
Hemoglobins of three baboons, Theropithecus gelada, Papio hamadryas- and Papio anubis, were purified and their oxygen equilibrium characteristics were studied. (a) Oxygen affinity, as expressed by P50, oxygen partial pressure for 50% oxygen binding, was in the order of gelada hemoglobin > anubis hemoglobin > hamadryas hemoglobin although the differences were small. (b) The presence of 2,3-diphosphoglycerate reduced their oxygen affinity in a similar manner. The effect on baboon hemoglobins was greater than that on human and Japanese monkey hemoglobins. (c) The intensity of the Bohr effect, as expressed by ?ΔlogP50ΔpH, at pH 7·4 agreed well with each other and the value was 0·62 in the presence of 2 mm diphosphoglycerate and 0·52 in its absence. These results indicate that phenotypic adaptation (acclimatory) may play an important role in the adaptation of gelada baboon to high altitudes.  相似文献   

2.
The oxygen equilibria of the hemoglobins of one holostean fish, the spotted gar (Lepisosteus osculatus), and of four teleost fish, the carpsucker (Carpiodes carpio), the small mouth buffalo fish (Ictiobus bubalus), the Rio Grande cichlid (Cichlasoma cyanoguttatum), and the redear sunfish (Lepomis microlophus), have been measured as a function of pH in the presence and absence of ATP. The oxygen equilibria of the teleost hemoglobins in the presence of 200 μm ATP can be superimposed within experimental error upon the data obtained in the absence of ATP by a simple downward shift of the pH scale by 0.5 unit. Thus the effects of proton and ATP binding appear equivalent: Both can be interpreted in terms of a two-state allosteric model in which binding occurs preferentially to the low-affinity T-state. The oxygen affinities of each of the teleost hemoglobins approach asymptotically a maximal value at high pH. Although these high affinities are associated with decreased cooperativity of oxygen binding, as reflected by the Hill coefficient n, the asymptotic value of n never appears lower than 1.2 to 1.4. This indicates that the data cannot be completely described in terms of a single high-affinity R-state in alkaline solution: At least two different conformations are required. The oxygen affinity of the spotted gar hemoglobin, like that of each of the teleost hemoglobins, reaches a maximal value (minimal value of log PO2 for half-saturation) above pH 8, but unlike teleost hemoglobins, the Hill coefficient reaches maximal values of 2.6 to 2.7 at high pH. The data in the absence of ATP are superimposable on the data in its presence by a downward shift of the pH scale by 0.25 unit. The measurement of the Bohr effect (ΔlogP30ΔpH) in the presence and absence of ATP shows that the Bohr effect in each of the hemoglobins is substantially enhanced by organic phosphates, as it is in mammalian hemoglobins. The extent of the enhancement of the Bohr effect by 200 μm ATP for each of the hemoglobins is approximately the same in the range pH 6.7 to 7.3 (increase in ΔlogP50ΔpH ~ 0.3). This is a direct consequence of the equivalence of the linked-function relationship to the effects of ATP and proton binding on oxygenation.  相似文献   

3.
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   

4.
Hemoglobin Cranston has an elongated β subunit owing to a frame shift mutation. Oxygen equilibrium measurements of stripped Hb Cranston3 at 20 °C in the absence of phosphate revealed a high affinity (P50 = 0·2 mm Hg at pH 7), non-co-operative hemoglobin variant with markedly reduced Böhr effect (logP50Δ pH7–8 = 0·2). The addition of inositol hexaphosphate resulted in an overall decrease in oxygen affinity (P50 = 0·7 mm Hg at pH 7), as well as an increase in co-operativity and Böhr effect (logP50Δ pH7–8 = 0·2). Rapid mixing and flash photolysis experiments reflected the equilibrium results. Over a pH range from 6 to 9 in the absence of phosphate, the rate of combination of carbon monoxide with Hb Cranston measured by a stopped-flow technique and following full or partial flash photolysis was extremely rapid (l′, l4, of ~ 6 × 106m?1s?1). In rapid kinetic experiments the addition of inositol hexaphosphate lowered the value of l′ to ~ 0·5 × 106m?1s?1 only after prior incubation with the deoxygenated protein. Inositol hexaphosphate had no effect on the rate of recombination of carbon monoxide following either full or partial flash photolysis. Overall oxygen dissociation and oxygen dissociation with carbon monoxide replacement, were measured and found to be slow (k, k4~ 11 s?1), consistent with a high affinity hemoglobin. Sedimentation equilibrium experiments revealed that Hb Cranston, at concentrations used in the functional studies, is somewhat less tetrameric than Hb A but nonetheless does not exist solely as a non-co-operative dimer. These kinetic and centrifugational findings in conjunction with X-ray diffraction evidence suggested that a high affinity tetramer of Hb Cranston exists which may equilibrate slowly with inositol hexaphosphate. Oxygen equilibrium measurements, ligand binding kinetics and X-ray diffraction studies on equivalent mixtures of Hb Cranston and Hb A revealed an interaction between these two hemoglobins in vitro that most probably exists in vivo. The presence of asymmetric hybrid molecules, α2βAβCranston, in the difference Fourier maps indicated that the hydrophobic tail of Hb Cranston is accommodated in the central cavity of the hybrid molecule between the two β chains and is relatively protected from the water environment, thus aiding in the stability of Hb Cranston in the red cell.  相似文献   

5.
6.
The observed equilibrium constants (Kobs) for the l-phosphoserine phosphatase reaction [EC 3.1.3.3] have been determined under physiological conditions of temperature (38 °C) and ionic strength (0.25 m) and physiological ranges of pH and free [Mg2+]. Using Σ and square brackets to indicate total concentrations Kobs = Σ L-serine][Σ Pi]Σ L-phosphoserine]H2O], K = L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O]. The value of Kobs has been found to be relatively sensitive to pH. At 38 °C, K+] = 0.2 m and free [Mg2+] = 0; Kobs = 80.6 m at pH 6.5, 52.7 m at pH 7.0 [ΔGobs0 = ?10.2 kJ/mol (?2.45 kcal/mol)], and 44.0 m at pH 8.0 ([H2O] = 1). The effect of the free [Mg2+] on Kobs was relatively slight; at pH 7.0 ([K+] = 0.2 m) Kobs = 52.0 m at free [Mg2+] = 10?3, m and 47.8 m at free [Mg2+] = 10?2, m. Kobs was insignificantly affected by variations in ionic strength (0.12–1.0 m) or temperature (4–43 °C) at pH 7.0. The value of K at 38 °C and I = 0.25 m has been calculated to be 34.2 ± 0.5 m [ΔGobs0 = ?9.12 kJ/mol (?2.18 kcal/ mol)]([H2O] = 1). The K for the phosphoserine phosphatase reaction has been combined with the K for the reaction of inorganic pyrophosphatase [EC 3.6.1.1] previously estimated under the same physiological conditions to calculate a value of 2.04 × 104, m [ΔGobs0 = ?28.0 kJ/mol (?6.69 kcal/mol)] for the K of the pyrophosphate:l-serine phosphotransferase [EC 2.7.1.80] reaction. Kobs = [Σ L-serine][Σ Pi][Σ L-phosphoserine][H2O], K = [L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O. Values of Kobs for this reaction at 38 °C, pH 7.0, and I = 0.25 m are very sensitive to the free [Mg2+], being calculated to be 668 [ΔGobs0 = ?16.8 kJ/mol (?4.02 kcal/mol)] at free [Mg2+] = 0; 111 [ΔGobs0 = ?12.2 kJ/mol (?2.91 kcal/mol)] at free [Mg2+] = 10?3, m; and 9.1 [ΔGobs0 = ?5.7 kJ/mol (?1.4 kcal/mol) at free [Mg2+] = 10?2, m). Kobs for this reaction is also sensitive to pH. At pH 8.0 the corresponding values of Kobs are 4000 [ΔGobs0 = ?21.4 kJ/mol (?5.12 kcal/mol)] at free [Mg2+] = 0; and 97.4 [ΔGobs0 = ?11.8 kJ/ mol (?2.83 kcal/mol)] at free [Mg2+] = 10?3, m. Combining Kobs for the l-phosphoserine phosphatase reaction with Kobs for the reactions of d-3-phosphoglycerate dehydrogenase [EC 1.1.1.95] and l-phosphoserine aminotransferase [EC 2.6.1.52] previously determined under the same physiological conditions has allowed the calculation of Kobs for the overall biosynthesis of l-serine from d-3-phosphoglycerate. Kobs = [Σ L-serine][Σ NADH][Σ Pi][Σ α-ketoglutarate][Σ d-3-phosphoglycerate][Σ NAD+][Σ L-glutamat0] The value of Kobs for these combined reactions at 38 °C, pH 7.0, and I = 0.25 m (K+ as the monovalent cation) is 1.34 × 10?2, m at free [Mg2+] = 0 and 1.27 × 10?2, m at free [Mg2+] = 10?3, m.  相似文献   

7.
Isolation and characterization of isocitrate lyase of castor endosperm   总被引:1,自引:0,他引:1  
Isocitrate lyase (threo-DS-isocitrate glyoxylate-lyase, EC 4.1.3.1) has been purified to homogeneity from castor endosperm. The enzyme is a tetrameric protein (molecular weight about 140,000; gel filtration) made up of apparently identical monomers (subunit molecular weight about 35,000; gel electrophoresis in the presence of sodium dodecyl sulfate). Thermal inactivation of purified enzyme at 40 and 45 °C shows a fast and a slow phase, each accounting for half of the intitial activity, consistent with the equation: At = A02 · e?k1t + A02 · e?k2t, where A0 and At are activities at time zero at t, and k1 and k2 are first-order rate constants for the fast and slow phases, respectively. The enzyme shows optimum activity at pH 7.2–7.3. Effect of [S]on enzyme activity at different pH values (6.0–7.5) suggests that the proton behaves formally as an “uncompetitive inhibitor.” A basic group of the enzyme (site) is protonated in this pH range in the presence of substrate only, with a pKa equal to 6.9. Successive dialysis against EDTA and phosphate buffer, pH 7.0, at 0 °C gives an enzymatically inactive protein. This protein shows kinetics of thermal inactivation identical to the untreated (native) enzyme. Full activity is restored on adding Mg2+ (5.0 mm) to a solution of this protein. Addition of Ba2+ or Mn2+ brings about partial recovery. Other metal ions are not effective.  相似文献   

8.
Presteady-state kinetic studies of α-chymotrypsin-catalyzed hydrolysis of a specific chromophoric substrate, N-(2-furyl)acryloyl-l-tryptophan methyl ester, were performed by using a stopped-flow apparatus both under [E]0 ? [S]0 and [S]0 ? [E]0 conditions in the pH range of 5–9, at 25 °C. The results were accounted for in terms of the three-step mechanism involving enzyme-substrate complex (E · S) and acylated enzyme (ES′); no other intermediate was observed. This substrate was shown to react very efficiently, i.e., the maximum of the second-order acylation rate constant (k2Ks)max = 4.2 × 107 M?1 s?1. The limiting values of Ks′ (dissociation constant of E · S), K2 (acylation rate) and k3 (deacylation rate) were obtained from the pH profiles of these parameters to be 0.6 ± 0.2 × 10?5 m, 360 ± 15 s?1 and 29.3 ± 0.8 s?1, respectively. Likewise small values were observed for Ki of N-(2-furyl)-acryloyl-l-tryptophan and N-(2-furyl)acryloyl-d-tryptophan methyl ester and Km of N-(2-furyl)acryloyl-l-tryptophan amide. The strong affinities observed may be due to intense interaction of β-(2-furyl)acryloyl group with a secondary binding site of the enzyme. This interaction led to a k?1k2 value lower than unity, i.e., the rate-limiting process of the acylation was the association, even with the relatively low k2 value of this methyl ester substrate, compared to those proposed for labile p-nitrophenyl esters.  相似文献   

9.
A method for calculating the rate constant (KA1A2) for the oxidation of the primary electron acceptor (A1) by the secondary one (A2) in the photosynthetic electron transport chain of purple bacteria is proposed.The method is based on the analysis of the dark recovery kinetics of reaction centre bacteriochlorophyll (P) following its oxidation by a short single laser pulse at a high oxidation-reduction potential of the medium. It is shown that in Ectothiorhodospira shaposhnikovii there is little difference in the value of KA1A2 obtained by this method from that measured by the method of Parson ((1969) Biochim. Biophys. Acta 189, 384–396), namely: (4.5±1.4) · 103s?1 and (6.9±1.2) · 103 s?1, respectively.The proposed method has also been used for the estimation of the KA1A2 value in chromatophores of Rhodospirillum rubrum deprived of constitutive electron donors which are capable of reducing P+ at a rate exceeding this for the transfer of electron from A1 to A2. The method of Parson cannot be used in this case. The value of KA1A2 has been found to be (2.7±0.8) · 103 s?1.The activation energies for the A1 to A2 electron transfer have also been determined. They are 12.4 kcal/mol and 9.9 kcal/mol for E. shaposhnikovii and R. rubrum, respectively.  相似文献   

10.
A maximal rate of the ouabain-sensitive 204Tl influx in human erythrocytes can be attained at trace concentrations of Tl+ in Mg2+ isotonic media free of K+ and Na+. The maximal influx of Tl+ from isotonic Mg(NO3)2 at 20°C and pH 7.4 was 0.45 mM · 1?1 · h?1 with a Km of 0.025 mM. In contrast to the active influx of Tl+, the passive Tl+ fluxes were neither saturated nor influenced by external cations in the range of concentrations of Tl+ and K+ studied. The rate constants of Tl+ passive fluxes in human and cat erythrocytes can be related to pH by the equation log kin(out) = –A + B · pH, where A and B are empirical constants for particular conditions. The apparent activation energy was 16 and 11 kcal/mol in sulphate and nitrate media, respectively. Tl+ and the alkali metal cations seem to overcome a common barrier in the erythrocyte membrane. Nevertheless, the rate of the passive penetration of Tl+ is about two orders of magnitude faster than those of K+ or Rb+. An extra non-Coulombic interaction between Tl+ and membrane ligands appears to be involved providing an accumulation of Tl+ somewhere in the vicinity of the membrane barrier and increasing the diffusion fluxes of Tl+ in both directions.  相似文献   

11.
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   

12.
Hemoglobin Wayne (Hb Wayne) is a frame-shift, elongated α-chain variant that exists in two forms, with either asparagine or aspartic acid as residue 139. Oxygen equilibrium studies showed that stripped Hb Wayne Asn and Hb Wayne Asp possessed high oxygen affinity (P12 = 0.60 and 0.23 mmHg at pH 7, respectively), were non-co-operative and have a markedly reduced Bohr effect (log P12/pH (7 to 8) = 0.34 and 0.10, respectively). Adding organic phosphate results in a decreased oxygen affinity and increased Bohr effect for both Hbs Wayne. The overall rate of carbon monoxide binding at pH 7 (l′ = 5.6 × 106m?1s?1) was similar for both stripped Hbs Wayne and was 25-fold more rapid than that of stripped Hb A. When organic phosphate was added, Hb Wayne Asn exhibited a homogeneous slower rate of carbon monoxide binding (l′ = 2.6 × 106m?1s?1), whereas Hb Wayne Asp showed heterogeneous binding (l′ = 6.1 × 106 and 2.6 × 106m?1s?1 for fast and slow phases, respectively). The rates of overall oxygen dissociation and oxygen dissociation with carbon monoxide replacement for both Hbs Wayne were found to be slow compared to Hb A and uniquely different from each other. Similarly, sedimentation velocity experiments indicated that, although Hb Wayne Asn and Hb Wayne Asp were both less tetrameric than Hb A, each hemoglobin exhibited a distinct degree of oxygen-linked subunit dissociation. These observed differences in the allosteric properties of Hb Wayne Asn and Hb Wayne Asp appeared to be directly attributable to residue 139. The equilibrium and kinetic data are consistent with the X-ray diffraction analysis of Hb Wayne Asp, which shows that the C terminus of the deoxytetramers are severely disordered, a condition that results in major destabilization of the T conformation and disruption of normal hemoglobin function.  相似文献   

13.
14.
N-Phenylhydroxylamine is oxidized in aqueous phosphate buffer to nitrosobenzene, nitrobenzene, and azoxybenzene. Degradation is O2 dependent and shows general catalysis by H2PO4? (k1 = 2.3 M?2 sec?1) and PO4?3 (k2 = 2.3 × 105M?2 sec?1) or kinetically equivalent terms. Evidence is presented suggesting the intermediacy of a highly reactive species leading to these products.  相似文献   

15.
Dispersed acini from dog pancreas were used to examine the ability of dopamine to increase cyclic AMP cellular content and the binding of [3H]dopamine. Cyclic AMP accumulation caused by dopamine was detected at 1·10?8 M and was half-maximal at 7.9±3.4·10?7M. The increase at 1·10?5 M, (7.5-fold) was equal to the half-maximal increase caused by secretin at 1·10?9 M. Haloperidol, a dopaminergic receptor antagonist inhibited cyclic AMP accumulation caused by dopamine. The IC50 value for haloperidol, calculated from the inhibition of cyclic AMP increase caused by 1·10?5 M dopamine was 2.3±0.9·10?6M. Haloperidol did not alter basal or secretin-stimulated cyclic AMP content. [3H]Dopamine binding was studied on the same batch of cells as cyclic AMP accumulation. At 37°C, it was rapid, reversible, saturable and stereospecific. The Kd value for high affinity binding sites was 0.43±0.1·10?7M and 4.7±1.6·10?7M for low affinity binding sites. The concentration of drugs necessary to inhibit specific binding of dopamine by 50% was 1.2±0.4·10/t-7M noradrenaline, 2·10/t-7 M epinine, 4.1±1.8·10/t-6M fluphenazine, 8.0±1.6·10/t-6M haloperidol, 4.2±1.2·10?6Mcis-flupenthixol, 2.7±0.4·10?5Mtrans-flupenthixol, >1·10?5M apomorphine, sulpiride, naloxone and isoproterenol.  相似文献   

16.
17.
The electrical potential (Δψ) of intact cholinergic synaptic vesicles was measured in the presence and absence of the proton translocator carbonyl cyanide p-trifluoromethoxy-phenylhydrazone (FCCP), and the results were utilized to calculate the vesicular proton chemical gradient (ΔpH) and proton electrochemical potential μH+). At external pH = 7.4 the vesicles maintain a proton electrochemical gradient of ?+20 mV (positive inside) which is composed of Δψ??80 mV (negative inside) and ΔpH?1.6 (acidic inside). The proton chemical gradient (ΔpH) increases as a function of pHout whereas the vesicular electrical potential (Δψ) is only slightly affected by the external pH. Consequently, ΔμH+ is larger at basic external pH values (?+40 mV at pHout = 9.0) and smaller at acidic external pH values (ΔμH+?0 at (pHout = 5.6). The possible physiological role of the electrochemical potentials in maintaining high concentrations of acetylcholine within the cholinergic synaptic vesicle is discussed.  相似文献   

18.
A. Telfer  J. Barber 《BBA》1978,501(1):94-102
1. Ionophore A23187 induces uncoupling of potassium ferricyanide-dependent O2 evolution by envelope-free chloroplasts and oxaloacetate-dependent O2 evolution by intact chloroplasts. The half maximal concentration (C12) for stimulation of oxygen evolution in both cases is approximately 4 μM · 100 μg chlorophyll · ml?1.2. Ionophore A23187 also induces inhibition of CO2 and 3-phosphoglycerate-dependent O2 evolution by intact chloroplasts in the presence of 3 mM MgCl2. The half maximal concentrations (C12) for inhibition of O2 evolution are 3 μM and 5 μM respectively · 100 μg?1 chlorophyll · ml?1.3. A very high concentration of ionophore A23187 (10 μM · 20 μg?1 chlorophyll · ml?1) plus 0.1 mM EDTA lowers the fluorescence yield of intact chloroplasts suspended in a cation-free medium in the presence of 3-(3,4-dichlorophenyl)-1,1-dimethylurea, indicating loss of divalent cation from the diffuse double layers of the thylakoid membranes.4. These results are discussed in relation to ionophore A23187-induced divalent cation/proton exchange at both the thylakoid and the envelope membranes of intact chloroplasts.  相似文献   

19.
Binding of the chromogenic ligand p-nitrophenyl α-d-mannopyranoside to concanavalin A was studied in a stopped-flow spectrometer. Formation of the protein-ligand complex could be represented as a simple one-step process. No kinetic evidence could be obtained for a ligand-induced change in the conformation of concanavalin A, although the existence of such a conformational change was not excluded. The entire change in absorbance produced on ligand binding occurred in the monophasic process monitored in the stopped-flow spectrometer. The value of the apparent second-order rate constant (ka) for complex formation (ka = 54,000 s?1m? at 25 °C, pH 5.0, Γ/2 0.5) was independent of the protein concentration when the protein was in the range of 233–831 μm in combining sites and in excess of the ligand. The apparent first-order rate constant (k?a) for dissociation of the complex was obtained from the rate constant for the decomposition of the complex upon the addition of excess methyl α-d-mannopyranoside (k?a = 6.2 s?1 at 25 °C, pH 5.0, Γ/2 0.5). The ratio ka?a (0.9 × 104m?1) was in reasonable agreement with value of 1.1 ± 0.1 × 104m?1 determined for the equilibrium constant for complex formation by ultraviolet difference spectrometry. Plots of ln(kaT) and ln(kaT) vs 1T were linear (T is temperature) and were used to evaluate activation parameters. The enthalpies of activation for formation and dissociation of the complex are 9.5 ± 0.3 and 16.8 ± 0.2 kcal/mol, respectively. The unitary entropies of activation for formation and dissociation of the complex are 2.8 ± 1.1 and 1.3 ± 0.7 entropy units, respectively. These entropy changes are much less than those usually associated with substantial changes in the conformation of proteins.  相似文献   

20.
A wide range of concentrated random coil polysaccharide solutions have been assessed for textural attributes by a trained sensory panel. The only textural terms invoked to describe these model systems were ‘thickness’ and ‘stickiness’, which were shown to be highly correlated, and essentially identical numerically, using a ratio scaling technique. Viscosity (η) measurements over a wide range of shear rates (γ) for all these samples gave flow curves (log η versus log γ) of the same form. Differences in flow behaviour between samples could then be characterised completely by two parameters, the maximum viscosity at low shear rates (η0), and the shear rate (γ?0·1) at which η = solη010. A simple linear relationship was demonstrated between these two parameters and perceived thickness (T) or stickiness (S), irrespective of polysaccharide type. For Newtonian liquids, log T (or log S) varied linearly with log η. Hence the effective ‘in-mouth’ thickness of random coil polysaccharide solutions, in normal viscosity units, may be predicted directly from η0 and γ?0·1 by the simple relationship: log ηN = 1·13 log η0 + 0·45 logγ?0·1 ? 1·72 where ηN is the viscosity of a Newtonian solution which would be perceived as identical in thickness (and stickiness) to the polysaccharide solution.  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