首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The kinetics of fructose uptake was determined in perfused rat liver during steady-state fructose elimination. On the basis of the corresponding values of fructose concentration in the affluent and in the effluent medium, and the fructose and ATP concentration in biopsies, the kinetics of membrane transport and intracellular phosphorylation in the intact organ was calculated according to a model system. Carrier-mediated fructose transport has a high Km (67 mM) and V (30 μmoles · min?1 ·g?1). The calculated kinetic constants of the intracellular phosphorylation were compared with values obtained with an acid-treated rat liver high speed supernatant (values given in parentheses). Km with fructose 1.0 mM (0.7 mM), Km with ATP 0.54 mM (0.37 mM), V 10.3 μmoles · min?1 · g?1 (10.1 μmoles · min?1 · g?1, calculated on the basis of the highest measured rate of fructose uptake correcting the ATP concentration to saturating values). The kinetics of fructose uptake reveals that at Physiological fructose concentrations the membrane transport limits the rate of fructose uptake, thus protecting the liver from severe depletion of adenine nucleotides.  相似文献   

2.
The transport of sucrose by selected mutant and wild-type cells of Streptococcus mutans was studied using washed cocci harvested at appropriate phases of growth, incubated in the presence of fluoride and appropriately labelled substrates. The rapid sucrose uptake observed cannot be ascribed to possible extracellular formation of hexoses from sucrose and their subsequent transport, formation of intracellular glycogen-like polysaccharide, or binding of sucrose or extracellular glucans to the cocci. Rather, there are at least three discrete transport systems for sucrose, two of which are phosphoenolpyruvate-dependent phosphotransferases with relatively low apparent Km values and the other a non-phosphotransferase (non-PTS) third transport system (termed TTS) with a relatively high apparent Km. For strain 6715-13 mutant 33, the Km values are 6.25·10?5 M, 2.4·10?4 M, and 3.0·10?3 M, respectively; for strain NCTC-10449, the Km values are 7.1·10?5 M, 2.5·10?4 M and 3.3·10?3 M, respectively. The two lower Km systems could not be demonstrated in mid-log phase glucose-adapted cocci, a condition known to repress sucrose-specific phosphotransferase activity, but under these conditions the highest Km system persists. Also, a mutant devoid of sucrose-specific phosphotransferase activity fails to evidence the two high affinity (low apparent Km) systems, but still has the lowest affinity (highest Km) system. There was essentially no uptake at 4°C indicating these processes are energy dependent. The third transport system, whose nature is unknown, appears to function under conditions of sucrose abundance and rapid growth which are known to repress phosphoenolpyruvate-dependent sucrose-specific phosphotransferase activity in S. mutans. These multiple transport systems seem well-adapted to S. mutans which is faced with fluctuating supplies of sucrose in its natural habitat on the surfaces of teeth.  相似文献   

3.
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   

4.
The uptake of radiolabeled carnitine and butyrobetaine has been studied in human heart cells (CCL 27). The uptake of carnitine is 3–10-fold higher in heart cells than in fibroblasts (pmol · μg DNA?1). The uptake of carnitine increases with temperature coefficient KT of 1.6 in the interval 10–20° C and with a negligible uptake at 4 and 10° C. The uptake of carnitine follows Michaelis-Menten kinetics with a KM of 4.8 ± 2.2 μM and V = 8.7 ± 3.2 pmol · μg DNA?1 · h?1. Carnitine uptake is suppressed 90% by NaF (24 mM). Butyrobetaine is taken up into heart cells to the same extent as carnitine with a KM of 5.7–17.3 μM and V = 8.7–9.3 pmol · μg DNA?1 · h?1. Butyrobetaine inhibits competitively the uptake of carnitine and carnitine inhibits the uptake of butyrobetaine to the same extent. No conversion of radiolabeled butyrobetaine to carnitine, or carnitine to methyl choline was observed intra- or extracellulary during incubation. These data are compatible with a selective transport mechanism for carnitine which is also responsible for the uptake of butyrobetaine.  相似文献   

5.
10?5 M cyclic AMP has high permeability in human erythrocyte ghosts (p = 0.061 · 10?6cm · s?1). Saturation of influx and efflux occurs. Kztoi = 4.43 mM. Vztoi = 259.6 μM · min?1. Kztio = 0.475 μM. Vztio = 28.3 μM · min?1 at 30°C. Equilibrium exchange entry of cyclic AMP has similar kinetics to zero trans influx, though the system does show counterflow. Cythochalasin B is an apparent competitive inhibitor of cyclic AMP exit. (Ki = 3.9 · 10?7M).Control experiments indicated that cyclic AMP remains intact during incubation with red blood cell ghosts and is contained within the intravesicular space during the transport experiments.  相似文献   

6.
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   

7.
Flufenamate, a non-steroidal anti-inflammatory drug, is a powerful inhibitor of anion transport in the human erythrocyte (I50 = 6·10?7M). The concentration dependence of the binding to ghosts reveals two saturable components. [14C]Flufenamate binds with high affinity (Kd1 = 1.2·10?7M) to 8.5·105 sites per cell (the same value as the number of band 3 protein per cell); it also binds, with lower affinity (Kd2 = 10?4M) to a second set of sites (4.6·107 per cell). Pretreatment of cells with 4-acetamido-4′-isothiocyanostilbene-2,2′-disulfonic acid (SITS), a specific inhibitor of anion transport, prevents [14C]flufenamate binding only to high affinity sites. These results suggest that high affinity sites are located on the band 3 protein involved in anion transport. Extracellular chymotrypsin and pronase at low concentration cleave the 95 kDa band 3 into 60 kDa and 35 kDa fragments without affecting either anion transport or [14C]flufenamate binding. Splitting by trypsin at the inner membrane surface of the 60 kDa chymotryptic fragment into 17 kDa transmembrane fragment and 40 kDa water-soluble fragment does not affect [14C]flufenamate binding. In contrast degradation at the outer membrane surface of the 35 kDa fragment by high concentration of pronase or papain decreases both anion transport capacity and number of high affinity binding sites for [14C]flufenamate. Thus it appears that 35 kDa peptide is necessary for both anion transport and binding of the inhibitors and that the binding site is located in the membrane-associated domain of the band 3 protein.  相似文献   

8.
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   

9.
The rotational correlation time of melittin, obtained from the nanosecond anisotropy of the emission from its single tryptophan residue, has been found to increase considerably in phosphate solution relative to that in aqueous solution, consistent with protein aggregation. The steady-state fluorescence spectra as well as the absorption spectra in phosphate solution exhibit a very good degree of similarity with those of the protein bound to egg phosphatidylcholine (PC) and distearoylphosphatidylcholine (DSPC) bilayer liposomes. The value of the second-order rate constant for dynamic quenching, kq = 1.4·109M?1·s?1, by acrylamide in 0.5 M phosphate solution is comparable to those for the protein-phospholipids complexes (1·109 and 0.7·109 M?1·s?1 for egg PC and DSPC, respectively). Similarities are also found in the nanosecond properties. There is a much stronger and quite similar dependence of the fluorescence spectra on time in the nanosecond range and of the fluorescence decay times on the emission wavelength in both cases as compared to the case in aqueous solution. These observations support the notion that melittin binds to the phospholipids in an aggregated form. The results suggest that the reduction in the kq values of bound melittin relative to that in aqueous solution and the blue shift of the fluorescence spectrum (from 352 to 337 nm) are brought about by shielding of the tryptophan residue from the solvent through a combination of protein aggregation and enhancement of its α-helical content (suggested by published CD data). The magnitude of the kq values for bound melittin, however, is still relatively high implying the occurrence of rather frequent encounters between the tryptophan residue and the hydrophilic acrylamide molecules. Thus, the residue is found not to penetrate deep into the phospholipid bilayer.  相似文献   

10.
In an accompanying publication by Duckwitz-Peterlein, Eilenberger and Overath ((1977) Biochim. Biophys. Acta 469, 311–325) it is shown that the exchange of lipid molecules between negatively charged vesicles consisting of total phospholipid extracts from Escherichia coli occurs by the transfer of single lipid monomers or small micelles through the water. Here a kinetic interpretation is presented in terms of a rate constant, k?, for the escape of lipid molecules from the vesicle bilayer into the water. The evaluated rate constants are k?P = (0.86 ± 0.05) · 10?5s?1 and k?E = (1.09 ± 0.13) · 10?6s?1 for phospholipid molecules with trans-Δ9-hexadecenoate and trans-Δ9-octadecenoate, respectively, as the predominant acyl chain component. The rate constants are discussed in terms of the acyl chain and polar head group composition of the lipids.  相似文献   

11.
12.
In order to test the question if a pool of lipophilic ions may exist in black lipid membranes which cannot be detected by electrical relaxation measurements we have performed simultaneously measurements of the optical absorption of a lipophilic ion. The absorbance of membrane-bound dipicrylamine at 410 nm was measured with a sensitive spectrophotometer which can detect absorbance changes ? 4 · 10?5. A minimal concentration of about 6 · 1011 dipicrylamine ions per cm2 of the membrane could be detected with this instrument. The dipicrylamine concentration in the membrane obtained with the optical method Ntopt is compared with the concentrations Ntel obtained from simultaneous electrical relaxation measurements. Ntopt and Ntel agreed at low dipicrylamine concentrations (10?8–10?7 M in the aqueous phase) and showed saturation at higher concentrations (up to 5 · 10?6 M). In the saturation range Ntopt was maximally four times higher than Ntel. The significance of this difference is discussed together with general aspects of the saturation phenomenon.  相似文献   

13.
Two new crystal forms of oxidized uteroglobin have been obtained. An orthorhombic one (P21212, Z = 2, a = 44.48 (5) A?, b = 36.93 (5) A?, c = 32·34 (5) A?) and a monoclinic one (P21, Z = 2, a = 44.56 (5) A?, b = 46.06 (5) A?, c = 37.43 (4) A?, β = 120.92 ° (5)). Both were grown at pH ~7.0 and diffract to a resolution of 2·1 to 2·2 Å. Data collections for native crystals have been recorded with an automatic four-circle diffractometer.  相似文献   

14.
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   

15.
The electronic structure of 19 established and potential biological oxidants has been studied by semiempirical all-valence-electron quantum-chemical methods. Electronic ground and excited states of O2, HO2, HO, H2O2, H3O, H4O2 and their (radical) ions have been investigated in order to get information on the geometry, vertical ionization potentials, vertical electron affinities and low-lying electronic excited states. The actual aim has been (i) to arrange the studied species according to their oxidizing power as given by gas-phase electron affinity.
9·HO·OH2O12>(1?+g).·OH>O12(1δ+g) >HO12(2A′)>O12(2A′)>O2(3?-g>HO·2)
and (ii) to contribute to the thermodynamics of early changes of the O2 molecule
O2+e→O?2·;O?2·+H+→HO·2
. Moreover, it has been found theoretically that the hydrated form of the hydroxyl radical (·HO.OH2) should be a relatively stable species with very high electron affinity (2·4 eV, INDO method). This circumstance and the theoretically predicted, extraordinarily low-lying, excited doublet state of the peroxyl radical (about 6000 cm?1) could be of biological significance.  相似文献   

16.
(1) Analysis of the data from steady-state kinetic studies shows that two reactions between cytochrome c and cytochrome c oxidase sufficed to describe the concave Eadie-Hofstee plots (Km ? 1 · 10?8M and Km ? 2 · 10?5M). It is not necessary to postulate a third reaction of Km ? 10?6M. (2) Change of temperature, type of detergent and type of cytochrome c affected both reactions to the same extent. The presence of only a single catalytic cytochrome c interaction site on the oxidase could explain the kinetic data. (3) Our experiments support the notion that, at least under our conditions (pH 7.8, low-ionic strength), the dissociation of ferricytochrome c from cytochrome c oxidase is the rate-limiting step in the steady-state kinetics. (4) A series of models, proposed to describe the observed steady-state kinetics, is discussed.  相似文献   

17.
The rhabdomeres of cephalopod photoreceptors, which are built up mainly of rhodopsin and phospholipid molecules, show a very high alkaline phosphatase activity. The enzyme has been partially characterized in purified rhodopsin vesicle fractions of the rhabdomeres by the following kinetic data: pH optimum 8.7; activation energy 9100 cal·m?1; Vmax = 2.5 μmol·min?1·mg?1; Km = 1.5·10?4M; its activity depends on Mg2+. There is good evidence that the alkaline phosphatase is a membrane-bound enzyme with receptor sites presumably located on the inside of the membrane. This enzyme has not been purified but its high activity compared to that of other known alkalin phosphatases (see Table I) indicates that each mirovillus, the structural unit of the rhabdomere, contains 1–20 enzyme molecules. This finding supports the hypothesis that the alkaline phosphatase is involved in the biochemical amplification process of excitation, or adaptation.  相似文献   

18.
125I-labelled α2-macroglobulin complexed with thrombin or plasmin bound to hepatocytes in a concentration-and time-dependent manner. The apparent Kd values calculated from displacement experiments were 7.9 · 10?8 M for α2-macroglobulin-thrombin and 8.5 · 10?8 M for α2-macroglobulin-plasmin. Association of these complexes was only partially reversible; after a 180 min incubation period, 50–60% of the bound radioactivity was internalized by the cells. α2-Macroglobulin itself bound also to hepatocytes, but the affinity of the α2-macroglobulin complexes was higher than that of the inhibitor alone, and α2-macroglobulin was not internalized, either. 125I-labelled thrombin or plasmin bound to hepatocytes as well. These bindings were also concentration-dependent and could be decreased with an excess of unlabelled ligands. Binding rates and amounts of the bound proteinases were higher than those of their α2-macroglobulin complexes. The α2-macroglobulin-thrombin complex competed with the α2-macroglobulin-plasmin complex in binding to hepatocytes, whereas there was no competition between these complexes and the antithrombin III-thrombin complex. These results suggest that the binding sites of hepatocytes for α2-macroglobulin-proteinase and antithrombin III-proteinase complexes are different.  相似文献   

19.
(1) Treatment of (Na+ + K+)-ATPase from rabbit kidney outer medulla with the γ-35S labeled thio-analogue of ATP in the presence of Na+ + Mg2+ and the absence of K+ leads to thiophosphorylation of the enzyme. The Km value for [γ-S]ATP is 2.2 μM and for Na+ 4.2 mM at 22°C. Thiophosphorylation is a sigmoidal function of the Na+ concentration, yielding a Hill coefficient nH = 2.6. (2) The thio-analogue (Km = 35 μM) can also support overall (Na+ + K+)-ATPase activity, but Vmax at 37°C is only 1.3 γmol · (mg protein)? · h?1 or 0.09% of the specific activity for ATP (Km = 0.43 mM). (3) The thiophosphoenzyme intermediate, like the natural phosphoenzyme, is sensitive to hydroxylamine, indicating that it also is an acylphosphate. However, the thiophosphoenzyme, unlike the phosphoenzyme, is acid labile at temperatures as low as 0°C. The acid-denatured thiophosphoenzyme has optimal stability at pH 5–6. (4) The thiophosphorylation capacity of the enzyme is equal to its phosphorylation capacity, indicating the same number of sites. Phosphorylation by ATP excludes thiophosphorylation, suggesting that the two substrates compete for the same phosphorylation site. (5) The (apparent) rate constants of thiophosphorylation (0.4 s?1 vs. 180 s?1), spontaneous dethiophosphorylation (0.04 s?1 vs. 0.5 s?1) and K+-stimulated dethiophosphorylation (0.54 s?1 vs. 230 s?1) are much lower than those for the corresponding reactions based on ATP. (6) In contrast to the phosphoenzyme, the thiophosphoenzyme is ADP-sensitive (with an apparent rate constant in ADP-induced dethiophosphorylation of 0.35 s?1, KmADP = 48 μM at 0.1 mM ATP) and is relatively K+-insensitve. The Km for K+ in dethiophosphorylation is 0.9 mM and in dephosphorylation 0.09 mM. The thiophosphoenzyme appears to be for 75–90% in the ADP-sensitive E1-conformation.  相似文献   

20.
Consistent with a recent literature report (Repine, J. E. etal. (1981) Proc.Nat.Acad.Sci.USA7?8?, 1001–1003), the release of [3H]-thymine from PM-2 DNA by Fe(II)-H2O2-generated ·OH was suppressed by dimethyl sulfoxide. In contrast, DMSO did not affect [3H]-thymine release mediated by Fe(II)-bleomycin. Under aerobic conditions in the presence of t-butyl phenylnitrone, Fe(II)-BLM produces an epr signal that has been presumed to arise by transfer of ·OH or O2? from the “active complex” of bleomycin to the spin trap. Remarkably, high concentrations (80 mM) of PBN had no effect on the ability of Fe(II)-BLM to solubilize [3H]-thymine, although the ability of authentic ·OH to degrade DNA was completely suppressed under these condition. The suproxide dismutase catalyst tetrakis(4-N-methylpyridyl)porphineiron(III) also failed to suppress BLM-mediated DNA degradation. Moreover, the epr signal observed with 1.6 mM Fe(II)-BLM in the presence of 80 mM PBN was found to be much less intense than that produced by 1.6 mM Fe(II) and 290 mM H2O2, but equivalent in intensity to that obtained with 45 mM Fe(II) and exoess H2O2. We conclude that the fragmentation of DNA produced by Fe(II)-BLM can be due neither to free ·OH nor to O2?. We suggest that DNA degradation is initiated by an “active complex” consisting of BLM, metal and oxygen that functions by abstracting H· from susceptible sites on DNA.  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