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.
Mitochondria were isolated from the hepatopancreas of the Florida spiny lobster Panulirus argus using a high osmolarity medium containing 600 mm mannitol, 83 mm sucrose, 5 mm 4-morpholinepropanesulfonic acid, pH 7.6, 0.5% bovine serum albumin (BSA), and 1 mm EDTA. O2 uptake and Ca2+ transport were measured by electrode methods in similar media (plus 4 mm KPi, 3.3 mm MgCl2, and 0.67 mg/ml BSA, with 80 mm KCl replacing a portion of the osmotic support). Substrate-supported respiration was observed to be coupled to phosphorylation of ADP or uptake of Ca2+ ions. State 3 rates (nanogram atoms O × minute?1 × milligram protein?1 ± SEM (N)) were: 49.2 ± 3.9 (19), succinate; 30.9 ± 3.9 (6), dl-palmitoyl carnitine; 29.0 ± 2.7 (9), l-malate; 40.0 ± 2.3 (3), l-glutamate; 27.7 ± 2.2 (5), d-3-hydroxybutyrate; and 26.4 ± 2.4 (18), l-proline ± pyruvate. α-Glycerol phosphate was not oxidized. Ca2+ uptake driven by succinate oxidation proceeded with Ca:O ratios of 4.0 ± 0.2 (SEM). Hepatopancreas mitochondria were not uncoupled by Ca2+ uptake in excess of 1100 ng atoms × mg protein?1. Ca2+ efflux could be induced by ruthenium red, indicating the presence of an active Ca2+ cycle. These mitochondria may provide a favorable model system in which to study regulation of the Ca2+ cycle.  相似文献   

2.
Abstract

Cordil-LND796 is a new cardiotonic glycoside under development. In rat brain microsomes where three isoforms of the Na, K-ATPase with differential affinities for cardiac glycosides have been identified, Cordil had higher affinity for the α3 (IC50 = 0.02 μM) than for the α2 (IC50 = 0.6 μM) and the α1 (IC50 = 30 μM) isozymes. Cordil is potentially a selective inhibitor for both α2 and α3 Na, K-ATPase isoforms. Using inside out vesicles we have shown that Cordil binds to and inhibits Na, K-ATPase at an extracellular site. The dissociation kinetic rates (k?1) from the ATPase and the phosphatase activity (K-dependent dephosphorylation) of the Na, K-ATPase were similar for Cordil. Despite these similarities to ouabain comparison of the kinetics of the Na, K-ATPase inhibition by ouabain and Cordil revealed marked differences in their association rates (k+1 = 0.7 1 mol1 min?1 and k+1 = 6 × 10?3 1 mol?1 min?1 respectively) and their dissociation rates (k?1 = 1.3 ± 0.2 × 10?4 S?1 and k?1 = 69 ± 7 × 10?4 s?1 respectively). Both binding association and dissociation rates were enhanced for Cordil. These data are compatible with a stabilizing effect of Cordil on the E2P conformational state of Na, K-ATPase.  相似文献   

3.
Abstract: The carotid injection technique, used previously to quantitate the kinetics of blood-brain barrier transport of metabolic substrates, may be modified to analyze the rate of cerebral glucose utilization. A 0.2-ml solution of [14C]glucose (GF) and [3H]methylglucose (M), an internal reference, is rapidly injected into the carotid artery, followed by microwave fixation of brain at various times up to 4 min after injection. The brain radioactivity is separated into a fraction containing neutral hexoses (GF and M) and a fraction containing metabolites of glucose. The GF/M ratio is related to the rate constant (k3) of brain glucose utilization by the simple, linear equation: In(GF/M) = In(GF°/M°) –k3t, where GF°/M°= the brain uptake index of glucose, relative to methylglucose, at 5-15 s after injection, and t= the time after carotid injection, e.g., 1–4 min. It is assumed that (a) the rate of influx due to recirculation of label is minimal during the 4-min circulation period; and (b) the rate constants of glucose efflux (k2) and methylglucose efflux (k2*) are identical. Independent estimates of k2 and k2* showed these parameters to be identical: k2= 0.14 + 0.08 min-I; k2*= 0.14 ± 0.02 min-I. A logarithmic plot of GF/M ratios versus time was linear (r = 0.99), and was described by the slope k2= 0.21 ± 0.02 min?1. Assuming glucose is uniformly distributed in brain, then the glycolytic rate = k3× brain glucose = (0.21 min?1) (2.6 μmol g?1) = 0.55 μmol min?1 g?1 for the cortex of the barbiturate-anesthetized rat. These studies provide the basis for a simple method of measurement of regional brain glycolysis that does not require either the use of correction factors, e.g., the lumped constant, or the use of differentially labeled glucose.  相似文献   

4.
The steady-state kinetics of the NADPH + FAD-dependent reduction of nitrate by nitrate reductase from Penicilliumchrysogenum was studied at pH 6.18. At this sub-optimum pH, Vmax was about 83 units × mg protein?1 compared with 225 units × mg protein?1 at pH 7.20. All initial velocity reciprocal plot patterns at pH 6.18 as well as the NADP+/nitrate product inhibition pattern were intersecting. In contrast, the NADP(H)/nitrate plots at pH 7.20 were parallel (Renosto, F. etal. J. Biol. Chem. 256, 8616, 1981). A major effect of lowering the assay pH was to change the Km for FAD from 0.17 μM at pH 7.20 to 4 μM at pH 6.18. The results suggest that nitrate reductase has a steady-state random kinetic mechanism in which kcat in the forward direction at pH 7.20 (ca. 375 sec?1) is greater that koff for the dissociation of one or more substrates. Several observations suggest that koff for FAD is extremely small at pH 7.20.  相似文献   

5.
《Free radical research》2013,47(1):205-210
Using the direct method of pulse radiolysis to determine the superoxide dismutase like activity of copper(II) cimetidine complexes, it was found that the reaction rate constant with O?2, kcat, was (8.5 ± 0.5) × 108 M?1s?1 independent of the cimetidine concentrations present in excess of 50–200 μM over the metal. The results suggest that either the 1:1 ligand to metal complex does not catalyze O?2 dismutation at a comparable rate to that of the 2:1 complex, or that the stability constant of the last species is much higher than that determined earlier by Kimura el al.,1 and only the 2:1 species is present in the solutions. With the indirect methods of cytochrome c and NBT for determining the ability of these complexes to catalyze O?2 dismutation, these compounds exhibited a much lower SOD activity. and kcat was determined to be (5.0 ± 0.3) × 106 and (7.± 0.4) × 101 M?1s?1. respectively using the two assays.  相似文献   

6.
Tumor microenvironment has a high concentration of inorganic phosphate (Pi), which is actually a marker for tumor progression. Regarding Pi another class of transporter has been recently studied, an H+-dependent Pi transporter, that is stimulated at acidic pH in Caco2BBE human intestinal cells. In this study, we characterized the H+-dependent Pi transport in breast cancer cell (MDA-MB-231) and around the cancer tissue. MDA-MB-231 cell line presented higher levels of H+-dependent Pi transport as compared to other breast cell lines, such as MCF-10A, MCF-7 and T47-D. The Pi transport was linear as a function of time and exhibited a Michaelis-Menten kinetic of Km = 1.387 ± 0.1674 mM Pi and Vmax = 198.6 ± 10.23 Pi × h?1 × mg protein?1 hence reflecting a low affinity Pi transport. H+-dependent Pi uptake was higher at acidic pH. FCCP, Bafilomycin A1 and SCH28080, which deregulate the intracellular levels of protons, inhibited the H+-dependent Pi transport. No effect on pHi was observed in the absence of inorganic phosphate. PAA, an H+-dependent Pi transport inhibitor, reduced the Pi transport activity, cell proliferation, adhesion, and migration. Arsenate, a structural analog of Pi, inhibited the Pi transport. At high Pi conditions, the H+-dependent Pi transport was five-fold higher than the Na+-dependent Pi transport, thus reflecting a low affinity Pi transport. The occurrence of an H+-dependent Pi transporter in tumor cells may endow them with an alternative path for Pi uptake in situations in which Na+-dependent Pi transport is saturated within the tumor microenvironment, thus regulating the energetically expensive tumor processes.  相似文献   

7.
The kinetics and thermodynamics of Ga(III) exchange between gallium mononitrilotriacetate and human serum transferrin as well as those of the interaction between gallium-loaded transferrin and the transferrin receptor 1 were investigated in neutral media. Gallium is exchanged between the chelate and the C-site of human serum apotransferrin in interaction with bicarbonate in about 50 s to yield an intermediate complex with an equilibrium constant K 1 = (3.9 ± 1.2) × 10−2, a direct second-order rate constant k 1 = 425 ± 50 M−1 s−1 and a reverse second-order rate constant k −1 = (1.1 ± 3) × 104 M−1 s−1. The intermediate complex loses a single proton with proton dissociation constant K 1a = 80 ± 40 nM to yield a first kinetic product. This product then undergoes a modification in its conformation which lasts about 500 s to produce a second kinetic intermediate, which in turn undergoes a final extremely slow (several hours) modification in its conformation to yield the gallium-saturated transferrin in its final state. The mechanism of gallium uptake differs from that of iron and does not involve the same transitions in conformation reported during iron uptake. The interaction of gallium-loaded transferrin with the transferrin receptor occurs in a single very fast kinetic step with a dissociation constant K d = 1.10 ± 0.12 μM and a second-order rate constant k d = (1.15 ± 0.3) × 1010 M−1 s−1. This mechanism is different from that observed with the ferric holotransferrin and suggests that the interaction between the receptor and gallium-loaded transferrin probably takes place on the helical domain of the receptor which is specific for the C-site of transferrin and HFE. The relevance of gallium incorporation by the transferrin receptor-mediated iron-acquisition pathway is discussed.  相似文献   

8.
Anthroylcholine was utilized as an extrinsic fluorescent probe in rapid kinetic studies of calcium dissociation from calmodulin (koff = 10 S?1) and the calmodulin-troponin I complex (koff = 6 S?1). At concentrations lower than 70 μM, the mechanism of dye binding agreed with the simple kinetic scheme in which the dye binds exclusively to the respective calcium complexes of calmodulin and calmodulin-troponin I. The sensitivity of anthroylcholine also made possible the estimation of values for the association (1.0 ± 0.8) × 108M?1 S?1) and dissociation rate constants (2 ± 170 S?1) for troponin I binding to the calcium4-calmodulin complex.  相似文献   

9.
Kinetic parameters for high affinity [HA] uptake in vitro in synaptosomes from different mouse brain regions were investigated. Vmax was highest in the striatum [200 pmol.· mg protein?1 · 4 min?1], followed by the cortex [111 pmol · mg protein?1 · 4 min?1], hippocampus [63 pmol · mg protein?1 · 4 min?1], midbrain [21 pmol · mg protein?1 · 4 min?1] and, lowest, medulla oblongata [5 pmol · mg protein?1 · 4 min?1]. Km was about the same in all brain regions [0.9–1.4 μM]. No sign of HA uptake was detected in synaptosomes from the cerebellum. A clear relationship between Vmax for synaptosomal HA uptake of Ch in vitro and apparent turnover of ACh in vivo was found between the brain regions. Administration of oxotremorine [1 mg·kg?1 i.p.] decreased Vmax for HA uptake of Ch by 60% in the cortex and hippocampus, by 50% in the striatum and by 20% in the midbrain. This effect is in accordance with the previously observed marked decrease in turnover of ACh in these brain regions following oxotremorine treatment.  相似文献   

10.
Properties of the fully developed phosphate transport system in the fertilized egg of the sea urchin, Strongylocentrotus purpuratus, were investigated. The rates of phosphate transport at concentrations of external phosphate of 1 to 44 μM, both in the absence and in the presence of 100 μM arsenate, exhibit typical saturation kinetics. At sea water concentrations of 2 μM phosphate, the rate of uptake is about 2 × 10?9 μm/egg/minute at 15°C. Arsenate is a competitive inhibitor of phosphate transport, fully and immediately reversible in its effects, yielding Ki values ranging from 10.5 to 14.1 × 10?6 M in comparison to the corresponding apparent KM (Michaelis-Menten) constants for phosphate of 5.6 to 7.5 × 10?6 M (pH 8.0, 15°C). The rate of arsenate uptake in a phosphate deficient medium amounts to 2.8 to 2.9 × 10?10 μm arsenate/egg/minute at an arsenate concentration of 2.9 to 10.2 μM arsenate (HAsO4??), which is 9.5 and 5.6% of the rate of phosphate uptake at corresponding phosphate concentrations. Arsenate has essentially the same developmental effects at initial concentrations of 5–10 μM and 100 μM arsenate, namely no observable effects for exposure periods of 7.5 hours, although longer periods result in blockage of development at the early blastula stage. Outward flux of phosphate ions cannot be demonstrated by washing prelabelled eggs with sea water containing low or high concentrations of phosphate, even when phosphorylation has been blocked by exposing the eggs to a metabolic inhibitor. Phosphate uptake rates measured in the pH range from 5.0 to 10.0 reveal a sharp optimum at pH 8.8–8.9. Reference to the apparent pK' values of the phosphoric acid system indicate that the entering species is the HPO4?? ion. The effects on rates of phosphate uptake of exposure to sea water at pH values between 7 and 10 for 30 minute periods are fully reversible, but at lower pH values, reversal is delayed, and is only partial. Sodium molybdate (0.01 M), sodium pyrophosphate (1.5 × 10?4 M), and adenosine triphosphate (1–5 × 10?4 M) for exposure periods ranging from 40 to 180 minutes did not significantly affect phosphate uptake. Omission of Ca++ ion from artificial sea water is without effect on phosphate uptake but the absence of both Ca++ and Mg++ results in profound and irreversible depression of both phosphate uptake and development. The data of this and the following paper are consistent with the conclusion that the transport of phosphate involves a surface located carrier. The apparent secondary and tertiary ionization constants of phosphoric acid in sea water (ionic strength = 0.6885) were measured, resulting in a value for pK′2 = 6.14 and for pK′3 = 10.99, at 15°C and phosphate at infinite dilution.  相似文献   

11.
Nongelling solutions of structurally regular chain segments of agarose sulphate show disorder–order and order–disorder transitions (as monitored by the temperature dependence of optical rotation) that are closely similar to the conformational changes that accompany the sol–gel and gel–sol transitions of the unsegmented polymer. The transition midpoint temperature (Tm) for formation of the ordered structure on cooling is ~25 K lower than Tm for melting. Salt-induced conformational ordering, monitored by polarimetric stopped-flow, occurs on a millisecond time scale, and follows the dynamics expected for the process 2 coil ? helix. The equilibrium constant for helix growth (s) was calculated as a function of temperature from the calorimetric enthalpy change for helix formation (ΔHcal = ?3.0 ± 0.3 kJ per mole of disaccharide pairs in the ordered state), measured by differential scanning calorimetry. The temperature dependence of the nucleation rate constant (knuc), calculated from the observed second-order rate constant (kobs) by the relationship kobs = knuc(1 ? 1/s) gave the following activation parameters for nucleation of the ordered structure of agarose sulphate (1 mg mL?1; 0.5M Me4NCl or KCl): ΔH* = 112 ± 5 kJ mol?1; ΔS* = 262 ± 20 J mol?1 K?1; ΔG*298 = 34 ± 6 kJ mol?1; (knuc)298 = (7.5 ± 0.5) × 106 dm3 mol?1 s?1. The endpoint of the fast relaxation process corresponds to the metastable optical rotation values observed on cooling from the fully disordered form. Subsequent slow relaxation to the true equilibrium values (i.e., coincident with those observed on heating from the fully ordered state) was monitored by conventional optical rotation measurements over several weeks and follows second-order kinetics, with rate constants of (2.25 ± 0.07) × 10?4 and (3.10 ± 0.10) × 10?4 dm3 mol?1 s?1 at 293.7 and 296.2 K, respectively. This relaxation is attributed to the sequential aggregation processes helix + helix → dimer, helix + dimer → trimer, etc., with depletion of isolated helix driving the much faster coil–helix equilibrium to completion. Light-scattering measurements above and below the temperature range of the conformational transitions indicate an average aggregate size of 2–3 helices.  相似文献   

12.
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   

13.
Abstract: The kinetics of transport across the blood-brain barrier and metabolism in brain (hemisphere) of [14C]2-fluoro-2-deoxy-d -glucose (FDG) were compared to that of [3H]2-deoxy-d -glucose (DG) and d -glucose in the pentobarbital-anesthetized adult rat. Saturation kinetics of transport were measured with the brain uptake index (BUI) method. The BUI for FDG was 54.3 ± 5.6. Nonlinear regression analysis gave a Km of 6.9 ± 1.1 mM and a Vmax of 1.70 ± 0.32 μmol/min/g. The K1 for glucose inhibition of FDG transport was 10.7 ± 4.4 mM. The kinetic constants of influx (k1) and efflux (K2) for FDG were calculated from the Km, Vmax, and glucose concentrations of the hemisphere and plasma (2.3 ± 0.2 μmol/g and 9.9 ± 0.4 mM, respectively). The transport coefficient (k1 FDG/k1glucose) was 1.67 ± 0.07 and the phosphorylation constant was 0.55 ± 0.16. The predicted lumped constant for FDG was 0.89, whereas the measured hexose utilization index for FDG was 0.85 ± 0.16. Conclusion: The value for the lumped constant can be predicted on the basis of the known kinetic constants of FDG and glucose transport and metabolism, as well as brain and plasma glucose levels. Knowledge of the lumped constant is crucial in interpreting data obtained from 18FDG analysis of regional glucose utilization in human brain in pathological states. We propose that the lumped constant will rise to a maximum equal to the transport coefficient for FDG under conditions of transport limitation (hypoglycemia) or elevated glycolysis (ischemia, seizures), and will fall to a minimum equal to the phosphorylation coefficient during phosphorylation limitation (extreme hyperglycemia).  相似文献   

14.
The fluorescent analog of adenosine triphosphate (ATP)1 1,N6-ethenoadenosine triphosphate, (εATP), has been utilized as a substitute for ATP in the myosin and heavy meromyosin ATPase systems. For myosin, the analog εATP replaced ATP with a somewhat larger Km (2.6 × 10?4 mole ??1 for εATP as opposed to 8.8 × 10?5 mole ??1 for ATP), indicating that the apparent affinity of the enzyme for εATP is less than for ATP. Perhaps of more interest, further comparison yielded a Vmax for εATP about two and one half times the value for ATP (20 μmole PO4 sec?1 g protein?1 as opposed to 8.1 μmole sec?1 g protein?1). Results for the HMM-εATPase system were similar, yielding a Km value of 1.47 × 10?4 mole ??1 and a Vmax of 54.2 μmole PO4 sec?1 g protein?1, as opposed to corresponding Km and Vmax values of 1.23 × 10?4 mole ??1 and 20.4 μmole PO4 sec?1 g protein?1, respectively for the HMM-ATP interaction. The pH dependence of εATPase for both systems was comparable to ATP, suggesting a similarity in the mechanism of hydrolysis of the two nucleotides. Activation of εATPase by Ca2+ in the presence of 0.5 M KCl was comparable to ATPase for both systems, but inhibition by Mg2+ seemed to be more effective for εATPase. These results indicate that εATP is an excellent substitute for ATP in the myosin and heavy meromyosin systems and because of its insertion into the active site of these muscle proteins, it promises to be a very useful probe for conformation studies at this level.  相似文献   

15.
Static and dynamic light-scattering measurements are reported on zinc-insulin at room temperature (21 ± l°C) and pH = 6.88 in 0.1M NaCl aqueous solution. The experiments were performed at very low concentration, in the range 0.12 × 10?4 to 0.90 × 10?4 g cm?3. Within experimental error, we find no evidence for a critical micellar concentration in this system. The aggregation phenomenon starts immediately after preparation of the solutions, and takes several days to come to stable equilibrium. The concentration dependence of the diffusion coefficients, D z, = Do (1 — kDC), is negative, and kD was observed to decrease as a function of time, while the aggregate size was found to increase. The equivalent concentration coefficient, ?2BM W, obtained from static light scattering, showed a similar behavior, and, within experimental error, was found to be numerically equal to kD. From the relation found between the diffusion coefficient at infinite dilution and the molecular weight of the aggregates, log D0 = ?0.240 log M w ? 5.077, we deduce that the insulin aggregates are compact structures with a characteristic radius of 0.71 Å/(dalton)1/3, surrounded by a hydration layer of a thickness of 8.0 Å. The equilibrium aggregation number is approximately 10.  相似文献   

16.
It is demonstrated that cyanobacteria (both azotrophic and non‐azotrophic) contain heme b oxidoreductases that can convert chlorite to chloride and molecular oxygen (incorrectly denominated chlorite ‘dismutase’, Cld). Beside the water‐splitting manganese complex of photosystem II, this metalloenzyme is the second known enzyme that catalyses the formation of a covalent oxygen–oxygen bond. All cyanobacterial Clds have a truncated N‐terminus and are dimeric (i.e. clade 2) proteins. As model protein, Cld from Cyanothece sp. PCC7425 (CCld) was recombinantly produced in Escherichia coli and shown to efficiently degrade chlorite with an activity optimum at pH 5.0 [kcat 1144 ± 23.8 s?1, KM 162 ± 10.0 μM, catalytic efficiency (7.1 ± 0.6) × 106 M?1 s?1]. The resting ferric high‐spin axially symmetric heme enzyme has a standard reduction potential of the Fe(III)/Fe(II) couple of ?126 ± 1.9 mV at pH 7.0. Cyanide mediates the formation of a low‐spin complex with kon = (1.6 ± 0.1) × 105 M?1 s?1 and koff = 1.4 ± 2.9 s?1 (KD ~ 8.6 μM). Both, thermal and chemical unfolding follows a non‐two‐state unfolding pathway with the first transition being related to the release of the prosthetic group. The obtained data are discussed with respect to known structure–function relationships of Clds. We ask for the physiological substrate and putative function of these O2‐producing proteins in (nitrogen‐fixing) cyanobacteria.  相似文献   

17.
The translational and rotational dynamics of tobacco mosaic virus in sodium phosphate buffer (pH =7.5) solutions has been investigated by polarized and depolarized light scattering Rayleigh linewidth studies. For concentrations ranging from 1.75 × 10?4 g ml?1 to 0.25 × 10?4 g ml?1 the translational diffusion coefficient (DT) has been found to be slightly concentration dependent and extrapolation to zero concentration gives D020°C = 0.34 ± 0.01 × 10?7 cm2S?1. A full analysis of the polarized spectra obtained at high and low scattering angles and the depolarized spectra at near zero scattering angles has enabled these techniques to be compared and the rotational diffusion constant DR to be determined. At a solution concentration of 1.75 × 10?4 g ml?1 a mean value is found to be DR20°C = 350 ± 30s?1. These values of DT and DR are in approximate agreement with calculations based on models of the tobacco mosaic virus molecule as a cylindrical rod.  相似文献   

18.
The interactions between cytochrome P450 2B4 (d-2B4), NADPH:cytochrome P450 reductase and cytochrome b5 have been investigated in the monomeric reconstituted P450 2B4-containing monooxygenase system in the presence of a substrate (7-pentoxyresorufin) and an electron donor, NADPH. Each partner was immobilized via its amino groups on the carboxymethyldextran biochip surface of the optical biosensor IAsys+. Such mode immobilization was not accompanied by any loss of activities of the immobilized proteins. The formation of binary d-Fp/d-2B4 complexes was registered. The association/dissociation rate constants (kon/koff) were (0.013 ± 0.005) × 106 M?1 s?1/0.05 ± 0.02 s?1, and dissociation constant (KD) was (0.26 ± 0.13) × 10?6 M. Comparison of kon, koff and KD values for d-Fp/d-2B4 complexes formed under hydroxylation (O-dealkylation) with corresponding constants obtained for the oxidized proteins of (0.10 ± 0.03) × 106 M?1 s?1/(0.14 ± 0.06) s?1, and (0.71 ± 0.37) × 10?6 M, respectively shows that the decrease in kon and an insignificant decrease in KD are associated with the increase of complex lifetime during transition from the oxidized to hydroxylation conditions. Complex formation between d-Fp and d-b5 was not registered in both hydroxylation conditions and in the case of oxidized forms of these proteins. In both cases formation of the ternary d-Fp/d-2B4/d-b5 complexes occurred.  相似文献   

19.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

20.
Florida's red tide organism, Gymnodinium breve, utilized exogenous glucose in the light for the synthesis of cellular components. Glucose was not taken up in the dark. Kinetic parameters for glucose uptake include a KFD of 11 μM and a Vmax of 1 × 10?10 mol of glucose taken up/mg cellular protein/hr. Glucose uptake was competitively inhibited by phloridzin (Ki = 40 μM), mannose (Ki = 12O μM), and 2-deoxy-d-glucose (Ki = 190 μM) and non-competitively inhibited by galactose (Ki = 125 μM). Kinetics and inhibition of glucose uptake are consistent with a facilitated diffusion transport system.  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