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.
A comparative study using laser flash photolysis of the kinetics of reduction and intramolecular electron transfer among the redox centers of chicken liver xanthine dehydrogenase and of bovine milk xanthine oxidase is described. The photogenerated reductant, 5-deazariboflavin semiquinone, reacts with the dehydrogenase (presumably at the Mo center) in a second-order manner, with a rate constant (k = 6 x 10(7) M-1 s-1) similar to that observed with the oxidase [k = 3 x 10(7) M-1 s-1; Bhattacharyya et al. (1983) Biochemistry 22, 5270-5279]. In the case of the dehydrogenase, neutral FAD radical formation is found to occur by intramolecular electron transfer (kobs = 1600 s-1), presumably from the Mo center, whereas with the oxidase the flavin radical forms via a bimolecular process involving direct reduction by the deazaflavin semiquinone (k = 2 x 10(8) M-1 s-1). Biphasic rates of Fe/S center reduction are observed with both enzymes, which are due to intramolecular electron transfer (kobs approximately 100 s-1 and kobs = 8-11 s-1). Intramolecular oxidation of the FAD radical in each enzyme occurs with a rate constant comparable to that of the rapid phase of Fe/S center reduction. The methylviologen radical, generated by the reaction of the oxidized viologen with 5-deazariboflavin semiquinone, reacts with both the dehydrogenase and the oxidase in a second-order manner (k = 7 x 10(5) M-1 s-1 and 4 x 10(6) M-1 s-1, respectively). Alkylation of the FAD centers results in substantial alterations in the kinetics of the reaction of the viologen radical with the oxidase but not with the dehydrogenase. These results suggest that the viologen radical reacts directly with the FAD center in the oxidase but not in the dehydrogenase, as is the case with the deazaflavin radical. The data support the conclusion that the environments of the FAD centers differ in the two enzymes, which is in accord with other studies addressing this problem from a different perspective [Massey et al. (1989) J. Biol. Chem. 264, 10567-10573]. In contrast, the rate constants for intramolecular electron transfer among the Mo, FAD, and Fe/S centers in the two enzymes (where they can be determined) are quite similar.  相似文献   

2.
We have examined the binding of [3H]bradykinin to bovine myometrial membranes and assessed its sensitivity to guanine nucleotides. Total binding displayed a typical B2 kinin receptor specificity. However, saturation binding isotherms were resolved into at least two components with KD values of 8 pM (45%) and 378 pM (55%). Low affinity binding exhibited relatively rapid rates of association (kobs = 1.40 x 10(-2) s-1) and dissociation (k-1 = 3.82 x 10(-3) s-1), while high affinity binding exhibited considerably slower rates (kobs = 9.52 x 10(-4) s-1 and k-1 = 4.43 x 10(-5) s-1). Pre-equilibrium dissociation kinetics revealed that formation of high affinity binding was characterized as a time-dependent accumulation of the slow dissociation rate at the expense of at least one other more rapid dissociation rate. In the presence of 10 microM guanyl-5'-yl imidodiphosphate (Gpp(NH)p), at least two binding components were resolved with KD values of 37 pM (12%) and 444 pM (88%). Gpp(NH)p apparently specifically perturbed high affinity binding by completely preventing the accumulation of the slow dissociation phase. Instead, two more rapid dissociation rates (k-1 = 8.53 x 10(-3) s-1 and 4.43 x 10(-4) s-1) were observed. These results suggest that [3H]bradykinin interacts with at least two B2 kinin receptor-like binding sites in bovine myometrial membranes. A three-state model for the guanine nucleotide-sensitive agonist interaction with the high affinity binding sites is proposed.  相似文献   

3.
The association process of FAD and apo-electron-transferring flavoprotein (apoETF) from hog kidney was investigated. The reaction schemes which involve the association-dissociation of the protein species could be excluded by the light scattering data, which indicated that the molecular weights of apoETF and holoETF are identical. The binding reaction between FAD and a large excess of apoETF was monophasic and obeyed pseudo-first order kinetics. On the other hand, the reaction between apoETF and a large excess of FAD was biphasic: the fast phase obeyed a pseudo-first order reaction, and the rate of the slow phase was almost independent of FAD concentration. These results suggest the existence of two different forms of apoETF, as represented in the following reaction scheme: [formula: see text] where "F" is FAD, "H" is holoETF, and "A" and "A" are the different forms of apoETF. The kinetic parameters were determined as k-1 = 3.9 x 10(4) M-1.s-1, k-1 approximately 10(-5) s-1, k+2 = 1.0 x 10(-3) s-1, and k-2 = 3.1 x 10(-3) s-1, in 50 mM potassium phosphate buffer, pH 7.6, containing 0.3 mM EDTA, and 5% v/v glycerol, at 7 degrees C. The elution patterns of apoETF on molecular sieve chromatography were very different from that of holoETF although the true molecular weights were identical. This result suggests that the structure of apoETF differs greatly from that of holoETF.  相似文献   

4.
The molecular mass of destabilase isolated from the medicinae leech Hirudo medicinalis was found to be equal to 12.3 kDa. A kinetic analysis of the sole presently known synthetic substrate, L-gamma-Glu-pNA, showed that the enzyme is relatively stable to heating (5 min, 70 degrees C); the pH optimum lies at 7.0-8.5. The enzyme has a specific activity of 0.15 x 10(-9) mol.s-1.mg-1; Km = 2.2 x 10(-4) M, kcat is 3.53 x 10(-3) s-1 (pH 8.0, 37 degrees C).  相似文献   

5.
Acryloyl-CoA reductase from Clostridium propionicum catalyses the irreversible NADH-dependent formation of propionyl-CoA from acryloyl-CoA. Purification yielded a heterohexadecameric yellow-greenish enzyme complex [(alpha2betagamma)4; molecular mass 600 +/- 50 kDa] composed of a propionyl-CoA dehydrogenase (alpha2, 2 x 40 kDa) and an electron-transferring flavoprotein (ETF; beta, 38 kDa; gamma, 29 kDa). A flavin content (90% FAD and 10% FMN) of 2.4 mol per alpha2betagamma subcomplex (149 kDa) was determined. A substrate alternative to acryloyl-CoA (Km = 2 +/- 1 microm; kcat = 4.5 s-1 at 100 microm NADH) is 3-buten-2-one (methyl vinyl ketone; Km = 1800 microm; kcat = 29 s-1 at 300 microm NADH). The enzyme complex exhibits acyl-CoA dehydrogenase activity with propionyl-CoA (Km = 50 microm; kcat = 2.0 s-1) or butyryl-CoA (Km = 100 microm; kcat = 3.5 s-1) as electron donor and 200 microm ferricenium hexafluorophosphate as acceptor. The enzyme also catalysed the oxidation of NADH by iodonitrosotetrazolium chloride (diaphorase activity) or by air, which led to the formation of H2O2 (NADH oxidase activity). The N-terminus of the dimeric propionyl-CoA dehydrogenase subunit is similar to those of butyryl-CoA dehydrogenases from several clostridia and related anaerobes (up to 55% sequence identity). The N-termini of the beta and gamma subunits share 40% and 35% sequence identities with those of the A and B subunits of the ETF from Megasphaera elsdenii, respectively, and up to 60% with those of putative ETFs from other anaerobes. Acryloyl-CoA reductase from C. propionicum has been characterized as a soluble enzyme, with kinetic properties perfectly adapted to the requirements of the organism. The enzyme appears not to be involved in anaerobic respiration with NADH or reduced ferredoxin as electron donors. There is no relationship to the trans-2-enoyl-CoA reductases from various organisms or the recently described acryloyl-CoA reductase activity of propionyl-CoA synthase from Chloroflexus aurantiacus.  相似文献   

6.
Rate constants for the hydration of bilirubin bound to unilamellar bilayers of dioleoylphosphatidylcholine and albumin were measured by stopped-flow methods. Rate constants for association of bilirubin with these vesicles and albumin were calculated from measured rate constants for dissociation and the equilibrium binding constants of bilirubin and lipids or albumin. Rate constants for hydration (dissociation) for bilirubin bound to dioleoylphosphatidylcholine and albumin were 71 s-1 and 1.8 s-1 respectively. Rate constants for association were 4.0 10(7) s-1 and 1.1 10(9) M-1 s-1, respectively. Both rates for interactions of bilirubin with bilayers were essentially independent of temperature in the range 0-40 degrees C, indicating that barriers to entry and exit of bilirubin from bilayers were entropic. Rates of transbilayer movement of bilirubin in dioleoylphosphatidylcholine were too fast to resolve by measuring rates of hydration of bilirubin. Rate constants for hydration of bilirubin bound to bilayers with less avidity for bilirubin as compared with dioleoylphosphatidylcholine also were too fast to measure with stopped-flow methods. In addition to providing details of the energetic basis for interactions between bilirubin and membranes, the data allow for calculating the maximal rates at which bilirubin could transfer spontaneously from sites on albumin in blood to the interior of cells. The data show, in this regard, that this rate is 10-50 fold faster than measured rates of uptake of bilirubin by intact liver.  相似文献   

7.
Conformational kinetics of triligated hemoglobin.   总被引:1,自引:0,他引:1       下载免费PDF全文
We have used the method of modulated excitation (Ferrone, F.A., and J.J. Hopfield, 1976, Proc. Natl. Acad. Sci. USA. 73:4497-4501), with an improved apparatus and a revised analytical procedure, to measure the rate of conformational change between the oxy (R) and deoxy (T) conformations of triligated carboxy-hemoglobin A at pH 6.5 and 7.0. We have found the rates to be kRT = 1.2 X 10(3) s-1 and kTR = 3.5 X 10(3) s-1 for pH 6.5, while for pH 7.0, kRT = 1.0 X 10(3) s-1, and kTR = 3.0 X 10(3) s-1. The value for L3, the equilibrium constant between conformations, was virtually unchanged between pH 6.5 and 7.0. While the rates measured here differ from those obtained in the original use of this method, these new rates are fully consistent with the original data when analyzed by the revised procedures presented here. When taken with other kinetic and equilibrium data, our measurements suggest that the transition state between structures is dominated by the behavior of the T quaternary structure. Finally, a spectral feature near the HbCO Soret peak has been observed that we ascribe to an allosteric perturbation of the spectra of the liganded hemes.  相似文献   

8.
B Durham  L P Pan  J E Long  F Millett 《Biochemistry》1989,28(21):8659-8665
Cytochrome c derivatives labeled at specific lysine amino groups with ruthenium bis(bipyridine) dicarboxybipyridine [RuII(bpy)2(dcbpy)] were prepared by using the procedure described previously [Pan, L. P., Durham, B., Wolinska, J., & Millett, F. (1988) Biochemistry 27, 7180-7184]. Four additional singly labeled derivatives were purified, bringing the total number to 10. These derivatives have a strong luminescence emission centered at 662 nm arising from the excited state, RuII*. Transient absorption spectroscopy was used to directly measure the rate constants for the photoinduced electron-transfer reaction from RuII* to the ferric heme group (k1) and for the thermal back-reaction from the ferrous heme group to RuIII (k2). The rate constants were found to be k1 = 14 X 10(6) s-1 and k2 = 24 X 10(6) s-1 for the derivative modified at lysine 72, which has a distance of 8-16 A between the ruthenium and heme groups. Similar rate constants were found for the derivatives modified at lysines 13 and 27, which have distances of 6-12 A separating the ruthenium and heme groups. The rate constants were significantly slower for the derivatives modified at lysine 25 (k1 = 1 X 10(6) s-1, k2 = 1.5 X 10(6) s-1) and lysine 7 (k1 = 0.3 X 10(6) s-1, k2 = 0.5 X 10(6) s-1), which have distances of 9-16 A. Transients due to photoinduced electron transfer could not be detected for the remaining derivatives, which have larger distances between the ruthenium and heme groups.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Quasielastic incoherent neutron scattering has been used to investigate the rate of local translational diffusion of lipid molecules in phospholipid bilayers of dipalmitoyl-phosphatidylcholine. The measured translational diffusion constants (4 x 10(-10) m2 s-1 at 63 degrees C and 1.4 x 10(-11) m2 s-1 at 30 degrees C) are considerably faster than those deduced using other less direct methods, but are in agreement with those measured in soap-water lyotropic liquid crystals, and with calculated values. This disagreement is attributed to differences in the time and distance scales characterising the various measurements. Quasielastic neutron scattering experiments observe fast motions over molecular distances, whereas other methods tend to measure a rate of diffusion which is averaged over macroscopic distances, and may thus contain contributions from long distance slow diffusive motions such as diffusion between the bilayers.  相似文献   

10.
A quantitative fluorescence assay has been developed to measure Cl flux across liposomal membranes for use in chloride transporter reconstitution studies. A Cl-sensitive fluorophore [6-methoxy-N-(3-sulfopropyl)quinolinium; SPQ] was entrapped into phospholipid/cholesterol liposomes formed by bath sonication, high-pressure extrusion, and detergent dialysis. Liposomes containing entrapped SPQ were separated from external SPQ by passage down a Sephadex G25 column. There was less than 10% leakage of SPQ from liposomes in 8 h at 4 degrees C and in 2 h at 23 degrees C. Cl influx (JCl in millimolar per second or nanomoles per second per centimeter squared) was determined from the time course of SPQ fluorescence, measured by cuvette or stopped-flow fluorometry, in response to inward Cl gradients. In 90% phosphatidylcholine (10% cholesterol liposomes at 23 degrees C, JCl in response to a 50 mM inward Cl gradient was 0.06 +/- 0.01 mM.s-1 (SD, n = 3) in the absence and 0.27 +/- 0.02 mM.s-1 in the presence of a K/valinomycin voltage clamp (0 mV), showing that the basal Cl "leak" is conductive; JCl increased (1.7 +/- 0.1)-fold in the presence of a 60-mV inside-positive diffusion potential. Accuracy of chloride influx rates determined by the SPQ method was confirmed by measurement of 36Cl uptake. In liposomes voltage-clamped to 0 mV, JCl was linear with external [Cl] (0-100 mM), independent of pH gradients, and strongly dependent on temperature (activation energy 18 +/- 1 kcal/mol, 12-42 degrees C) as predicted for channel-independent Cl diffusion.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
D Y Cai  M Tien 《Biochemistry》1990,29(8):2085-2091
The oxycomplexes (compound III, oxyperoxidase) of two lignin peroxidase isozymes, H1 (pI = 4.7) and H8 (pI = 3.5), were characterized in the present study. After generation of the ferroperoxidase by photochemical reduction with deazoflavin in the presence of EDTA, the oxycomplex is formed by mixing ferroperoxidase with O2. The oxycomplex of isozyme H8 is very stable, with an autoxidation rate at 25 degrees C too slow to measure at pH 3.5 or 7.0. In contrast, the oxycomplex of isozyme H1 has a half-life of 52 min at pH 4.5 and 29 min at pH 7.5 at 25 degrees C. The decay of isozyme H1 oxycomplex follows a single exponential. The half-lives of lignin peroxidase oxycomplexes are much longer than those observed with other peroxidases. The binding of O2 to ferroperoxidase to form the oxycomplex was studied by stopped-flow methods. At 20 degrees C, the second-order rate constants for O2 binding are 2.3 X 10(5) and 8.9 X 10(5) M-1 s-1 for isozyme H1 and 6.2 X 10(4) and 3.5 X 10(5) M-1 s-1 for isozyme H8 at pH 3.6 and pH 6.8, respectively. The dissociation rate constants for the oxycomplex of isozyme H1 (3.8 Z 10(-3) s-1) and isozyme H8 (1.0 X 10(-3) s-1) were measured at pH 3.6 by CO trapping. Thus, the equilibrium constants (K, calculated from kon/koff) for both isozymes H1 (7.0 X 10(7) M-1) and H8 (6.2 X 10(7) M-1) are higher than that of myoglobin (1.9 Z 10(6) M-1).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Distinct semimetmyohemerythrin species are produced by one-electron oxidation of deoxymyohemerythrin and one-electron reduction of metmyohemerythrin. The former, (semimetmyo)o, changes (greater than or equal to 90%) to the latter, (semimetmyo)R, with k = 1.0 x 10(-2) s-1, delta H = 15.1 kcal mol-1 and delta S = -17 eu. Oxidation of (semimetmyo)o by Fe(CN)6(3)- rapidly produces an unstable metmyohemerythrin form which converts to the final metmyohemerythrin with k = 4.6 x 10(-3) s-1, delta H = 16.8 kcal mol-1, and delta S = -13 eu. The two met forms react at the same rate with N3-, but the unstable form reacts very rapidly with S2O4(2-) in contrast to stable metmyohemerythrin. (Semimetmyo)R or a mixture of metmyohemerythrin and deoxymyohemerythrin equilibrate very slowly to a mixture containing all three species. The rate constants for disproportionation and comproportionation are 0.89 M-1 s-1 and 9.4 M-1 s-1, respectively. EPR spectra near liquid He temperatures and optical absorption spectra have been used to characterize and measure the rates at 25 degrees C, pH 8.2, and I = 0.15 M. The comparative behavior of octameric and monomeric protein is discussed.  相似文献   

13.
Recently, we have obtained evidence in favor of a structural interaction between the epidermal growth factor (EGF) receptor and the Triton X-100-insoluble cytoskeleton of epidermoid carcinoma A431 cells. Here we present a further analysis of the properties of EGF receptors attached to the cytoskeleton. Steady-state EGF binding studies, analyzed according to the Scatchard method, showed that A431 cells contain two classes of EGF-binding sites: a high-affinity site with an apparent dissociation constant (KD) of 0.7 nM (7.5 x 10(4) sites per cell) and a low-affinity site with a KD of 8.5 nM (1.9 x 10(6) sites per cell). Non-equilibrium binding studies revealed the existence of two kinetically distinguishable sites: a fast-dissociating site, with a dissociation rate constant (k-1) of 1.1 x 10(-3) s-1 (1.0-1.3 x 10(6) sites per cell) and a slow-dissociating site, with a k-1 of 3.5 x 10(-5) s-1 (0.6-0.7 x 10(6) sites per cell). The cytoskeleton of A431 cells was isolated by Triton X-100 extraction. Scatchard analysis revealed that approximately 5% of the original number of receptors were associated with the cytoskeleton predominantly via high-affinity sites (KD = 1.5 nM). This class of receptors is further characterized by the presence of a fast-dissociating component (k-1 = 2.0 x 10(-3) s-1) and a slow-dissociating component (k-1 = 9.1 x 10(-5) s-1). The distribution between fast and slow sites of the cytoskeleton was similar to that of intact cells (65% fast and 35% slow sites). Incubation of A431 cells for 2 h at 4 degrees C in the presence of EGF resulted in a dramatic increase in the number of EGF receptors associated to the cytoskeleton. These newly cytoskeleton-associated receptors appeared to represent low-affinity binding sites (KD = 7 nM). Dissociation kinetics also revealed an increase of fast-dissociating sites. These results indicate that at 4 degrees C EGF induces the binding of low-affinity, fast-dissociating sites to the cytoskeleton of A431 cells.  相似文献   

14.
Hydrofluoric and nitric acid transport through lipid bilayer membranes   总被引:5,自引:0,他引:5  
Hydrofluoric and nitric acid transport through lipid bilayer membranes were studied by a combination of electrical conductance and pH electrode techniques. Transport occurs primarily by nonionic diffusion of molecular HF and HNO3. Membrane permeabilities to HF and HNO3 ranged from 10(-4) to 10(-3) cm . s-1, five to seven orders of magnitude higher than the permeabilities to NO-3, F- and H+. Our results are consistent with the hypothesis that F- transport through biological membranes occurs mainly by nonionic diffusion of HF. Our results also suggest that of the two principal components of 'acid rain', HNO3 may be more toxic than H2SO4.  相似文献   

15.
Simultaneous studies were carried out of isotope and electric parameters of spheric bilayer membranes modified with gramicidin A and its analog O-pyromellithylgramicidin (PG) having three negative charges on the N-end of the molecule. The relationship between the electric coefficients of permeability and the isotope ones PG/P* = n was determined by two independent methods. It has been found that for the membranes modified with gramicidin A in RbCl concentrations from 2.2 x 10(-3) to 10(-1) M the value n is constant and approximates 2 and with RbCl concentration 1 M, n = 1.6. For the membranes modified with PG in 0.1 M solutions of PbCl n = 2. The results obtained in terms of the model of unilinear ion diffusion in a narrow pore indicate that in a gramicidin channel there are two sites of cation binding which are located near the channel mouth.  相似文献   

16.
The temperature dependences of fluorescence and phosphorescence spectra maxima of chromophor labels--endogenic (tryptophan) and exogenic (eosinisothiocyanate)--were measured for the preparations of photosynthetic membranes and reaction centers from Rhodospirillum rubrum. It was found that the dipole mobility of protein-lipid matrix in the vicinity of the chromophores intensified markedly with a temperature rise from 150 to 300K resulting in the corresponding relaxation time tau r decrease from 10(0) to 10(-8) s. The efficiency of direct transfer of the photomobilized electron in the system of quinone acceptors (A1- leads to A2) of reaction centers (characteristic half-times of the process being 10(-3) divided by 10(-4) s) was shown also to increase sharply at temperatures higher than 200K parallel to the enhancement of molecular motions with tau r approximately 10(-8) s. Meanwhile, changes observed in the rate of recombination of primary photoproducts, i.e. an oxidized bacteriochlorophyll dimer, P+ and a reduced acceptor, A1- (characteristic half-time of 10(-1) divided by 10(-2) s) and the activization of low-frequency motions with tau r approximately 10(-3) s in the external layers and tau r less than 1 s in the internal parts of the reaction centers protein develop over the same range of low temperatures (150-220 K). The nature of interactions which determine the dependence of the photosynthetic electron transport on the molecular mobility of the membrane proteins is discussed.  相似文献   

17.
Lecithin monolayer liposomes (1000 A in diameter) loaded with cytochrome c were placed into the external solution, in which O2 superoxide radicals were regenerated by the xanthine-xanthine oxidase system. The penetration of superoxide radicals across the liposomal membranes was followed by cytochrome c reduction in the interval volume of the liposomes. The effects of lipid membrane modifiers and temperature on this process were investigated. The results obtained were used for calculation of the permeability coefficients of bilayer lipid membranes for O(2) (P'O(2) = (7.6 +/- 0.3) . 10(-8) cm . s-1) or HO . 2(P'HO(2) = 4.9 x 10(-4) cm . s-1). The effect of the transmembrane electric potential (concentration gradient of H+, valinomycin) on the permeability of liposomal membranes for the superoxide radical was studied. The superoxide radical was down to penetrate across the bilayer lipid membranes in an unloaded state. Using an intramolecular cholesterol-amphotericin B-complex, the superoxide radicals were shown to penetrate across the bilayer lipid membranes, predominantly via the anionic channels.  相似文献   

18.
Glycolate oxidase from spinach has been expressed in Saccharomyces cerevisiae. The active enzyme was purified to near-homogeneity (purification factor approximately 1400-fold) by means of hydroxyapatite and anion-exchange chromatography. The purified glycolate oxidase is nonfluorescent and has absorbance peaks at 448 (epsilon = 9200 M-1 cm-1) and 346 nm in 0.1 M phosphate buffer, pH 8.3. The large bathochromic shift of the near-UV band indicates that the N(3) position is deprotonated at pH 8.3. A pH titration revealed that the pK of the N(3) is shifted from 10.3 in free flavin to 6.4 in glycolate oxidase. Glycolate oxidase is competitively inhibited by oxalate with a Kd of 0.24 mM at 4 degrees C in 0.1 M phosphate buffer, pH 8.3. Three pieces of evidence demonstrate that glycolate oxidase stabilizes a negative charge at the N(1)-C(2 = O) locus: the enzyme forms a tight sulfite complex with a Kd of 2.7 x 10(-7) M and stabilizes the anionic flavosemiquinone and the benzoquinoid form of 8-mercapto-FMN. Steady-state analysis at pH 8.3, 4 degrees C, yielded a Km = 1 x 10(-3) M for glycolate and Km = 2.1 x 10(-4) M for oxygen. The turnover number has been determined to be 20 s-1. Stopped-flow studies of the reductive (k = 25 s-1) and oxidative (k = 8.5 x 10(4) M-1 s-1) half-reactions have identified the reduction of glycolate oxidase to be the rate-limiting step.  相似文献   

19.
Vitamin A (retinol reacts extremely rapidly (k = 1.4 x 10(9) M-1 s-1) with thiyl free radicals derived from glutathione to form a free radical with a very strong visible absorption (lambda max. = 380 nm, E max. = 4.0 x 10(4) M-1 cm-1). Arachidonate, linolenate, linoleate and ascorbate also react readily but much more slowly (k = 2.2 x 10(7), 1.9 x 10(7), 1.3 x 10(7) and 3.6 x 10(8) M-1 s-1 respectively). These results support the possibility that vitamin A might play a role in protecting lipid membranes against thiyl free radical mediated damage.  相似文献   

20.
The displacement of NADH from cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) from sheep liver was studied by using NAD+, 1,10-phenanthroline, ADP-ribose, deamino-NAD+ and pyridine-3-aldehyde-adenine dinucleotide as displacing agents, by following the decrease in fluorescence as a function of time. The data obtained could be fitted by assuming two first-order processes were occurring, a faster process with an apparent rate constant of 0.85 +/- 0.20 s-1 and a relative amplitude of 60 +/- 10% and a slower process with an apparent rate constant of 0.20 +/- 0.05 s-1 and a relative amplitude of 40 +/- 10% (except for pyridine-3-aldehyde-adenine dinucleotide, where the apparent rate constant for the slow process was 0.05 s-1). The displacement rates did not change significantly when the pH was varied from 6.0 to 9.0. Kinetic data are also reported for the dependence of the rate of binding of NADH to the enzyme on the total concentration of NADH. Detailed arguments are presented based on the isolation and purification procedures, the equilibrium coenzyme-binding studies and the kinetic data, which lead to the following model for the release of NADH from the enzyme: (formula: see article). The parameters that best fit the data are: k + 1 = 0.2 s-1; k - 1 = 0.05 s-1; k + 2 = 0.8 s-1 and k - 2 = 5 X 10(5)litre-mol-1-s-1. The slow phase of the NADH release is similar to the steady-state turnover number for substrates such as acetaldehyde and propionaldehyde and appears to contribute significantly to the limitation of the steady-state rate.  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