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.
D-[14C]Glucose self exchange and unidirectional efflux from human red blood cells were studied at 20 degrees C (pH 7.2) by means of the Millipore-Swinnex filtering technique whose time resolution is greater than 1 s and the continuous flow-tube method with a time resolution of greater than 2 ms. The unidirectional efflux data were analyzed using both the method of initial rates and the integrated rate equation. Simple Michaelis-Menten kinetics apply to the results obtained under both experimental conditions. In self-exchange mode, the half-saturation constant, K1/2ex, was 10 (S.E. +/- 1) mM. In unidirectional efflux mode K1/2ue was 6.6 (S.E. +/- 0.5) mM (initial rates) or by the method of integrated rates 7.7 mM, with a range of 2.7-12.1 mM, K1/2ue increasing with an increased initial intracellular glucose concentration. Our results of K1/2ex oppose previous published values of 32 mM for self exchange (Eilam and Stein (1972) Biochim. Biophys. Acta 266, 161-173) and 25 mM for unidirectional efflux (Karlish et al. (1972) Biochim. Biophys. Acta 255, 126-132) that have been used extensively in kinetic considerations of glucose transport models. Under self-exchange conditions Jmaxex was 1.8 x 10(-10) mol cm-2s-1, and in unidirectional efflux mode Jmaxue was 8.3 x 10(-11) mol cm-2s-1 (initial rates) and 8.6 x 10(-11) mol cm-2s-1 (integrated rates). We suggest that the previous high values of Jmax and in particular K1/2 are due to the use of methods with insufficient time resolution. Our results indicate that the transport system is less asymmetric than was generally accepted, and that complicated transport models developed to account for the great difference between the determined K1/2 and J max values are redundant.  相似文献   

2.
At 0 degrees C, pH 7.3, palmitate (PA) binds to human erythrocyte ghosts suspended in 0.2% bovine serum albumin (BSA) solution with molar ratios of PA to BSA, v, between 0.2 and 1.3. The binding depends on the water phase PA concentration, measured in equilibrium experiments, using BSA-filled ghosts as semipermeable bags. The saturable binding has a capacity of 19.4 +/- 7.5 nmol g-1 packed ghosts (7.2 x 10(9) cells) and Kd = 13.5 +/- 5 nM. PA exchange efflux kinetics to 0.2% BSA is recorded from ghosts without and with 0.2% BSA with a resolution time of about 1 s. Data are analyzed in terms of compartmental models. Using BSA-free ghosts the kinetics is essentially monoexponential. The rate constant is 0.0287 +/- 0.0022 s-1. Using ghosts with BSA, the kinetics is biexponential with widely different rate constants. Extrapolated zero-time values reflect, according to additional investigations, 'instantaneous' release of PA from the outer surface of the ghosts. Analyses of the biexponential curve up to about 55% tracer efflux assign unequivocally values to three model parameters. (1) k1, the dissociation rate constant of the PA-BSA complex is (1.47 +/- 0.03) x 10(-3) s-1 and (2.56 +/- 0.08) x 10(-3) s-1 and (4.08 +/- 0.13) x 10(-3) s-1 at v = 0.2, 0.6 and 1.4, respectively. (2) k3*, the overall rate constant of PA transport from the inside of the ghost membrane to the medium is 0.0269 +/- 0.0020 s-1 independent of v. (3) Qkin, the ratio of PA on the inside of the membrane to PA on BSA within the ghosts is v dependent and smaller than a corresponding ratio Qeq measured in equilibrium by a value corresponding to PA on the outer surface. This fraction is released with a rate constant, k5, which is of the order of 1 s-1. The data suggest a maximum PA transport capacity, Jmax, of 2 pmol min-1 cm-2, 0 degrees C, pH 7.3.  相似文献   

3.
There was developed a procedure for immobilization of phosphatidyl cholines (Egg yolk phosphatidyl choline and polyunsaturated soya beams phosphatidyl choline) on the modified apoHDL-Sepharose. The formation of phospholipid micelles was proved by linear dependence of the content of the sorbed phosphatidyl choline versus, the content of apoHDL bound to Sepharose. Incubation of apoHDL/PC-Sepharose with human plasma was shown to change the plasma lipid composition. The apoHDL/PC-Sepharose might be used for correction of the plasma lipid composition on vitro experiments.  相似文献   

4.
Role of apolipoproteins in cellular cholesterol efflux   总被引:1,自引:0,他引:1  
The effects of serum apolipoproteins, particle size and concentration on the effectiveness of phosphatidylcholine (PC)-containing acceptor particles in causing release of cholesterol from cells growing in culture have been investigated. The acceptor particles were prepared by detergent-dialysis procedures and were either egg PC small unilamellar vesicles (SUV) or discoidal complexes of egg PC with apoproteins from human high-density lipoprotein (HDL). Gel filtration chromatography was employed to isolate particles of defined composition and size. The half-times (t 1/2) for the unidirectional efflux of cholesterol from cells prelabeled with [3H]cholesterol were measured as a function of acceptor PC concentration in the extracellular medium. HDL apolipoprotein-egg PC discoidal complexes at 100 micrograms PC/ml gave the following t 1/2 values when incubated with rat Fu5AH hepatoma, human HepG2 hepatoma, human GM3468 skin fibroblast, L-cell and mouse J774 macrophage-tumor cells: 11 +/- 2, 22 +/- 5, 84 +/- 18, 17 +/- 2 and 32 +/- 6 h, respectively. Equivalent experiments using purified apolipoprotein A-I or the total apolipoprotein C fraction to form the egg PC complexes showed that the t 1/2 values for the hepatoma cells were unaltered. However, with the fibroblasts, L-cells and J774 macrophages, the apolipoprotein C complexes gave significantly longer t 1/2 than complexes of egg PC with either apolipoprotein A-I or HDL apolipoprotein which gave the same t 1/2. An analysis based on the theory of fast coagulation of colloid particles to describe collisions between desorbed cholesterol molecules and acceptor particles predicts that the dependence of t 1/2 for cholesterol efflux from a given cell to different acceptors should be normalized when the extracellular level of acceptors is expressed in terms of the product of the radius of the particle times the number concentration of acceptor particles. The decrease in t 1/2 for cholesterol efflux from fibroblasts when the egg PC acceptor was changed from an SUV to an apolipoprotein HDL discoidal complex is consistent with the above concepts. The primary effect of the apolipoproteins in promoting cellular cholesterol efflux seems to be the solubilization of PC so that the PC is present in the extracellular medium as many small particles.  相似文献   

5.
The present study compared ouabain-sensitive unidirectional K+ flux into (JinK) and out of (JoutK) perfused rat hindlimb skeletal muscle in situ and mouse flexor digitorum brevis (FDB) in vitro. In situ, 5 mM ouabain inhibited 54 +/- 4% of the total JinK in 28 +/- 1 min, and increased the net and unidirectional efflux of K+ within 4 min. In contrast, 1.8 mM ouabain inhibited 40 +/- 8% of the total JinK in 38 +/- 2 min, but did not significantly affect JoutK. In vitro, 1.8 and 0.2 mM ouabain decreased JinK to a greater extent (83 +/- 5%) than in situ, but did not significantly affect 42K loss rate compared with controls. The increase in unidirectional K+ efflux (JoutK) with 5 mM ouabain in situ was attributed to increased K+ efflux through cation channels, since addition of barium (1 mM) to ouabain-perfused muscles returned JoutK to baseline values within 12 min. Perfusion with 5 mM ouabain plus 2 mM tetracaine for 30 min decreased JinK 46 +/- 9% (0.30 +/- 0.03 to 0.16 +/- 0.02 micromol x min(-1) x g(-1)), however tetracaine was unable to abolish the ouabain-induced increase in unidirectional K+ efflux. In both rat hindlimb and mouse FDB, tetracaine had no effect on JoutK. Perfusion of hindlimb muscle with 0.1 mM tetrodotoxin (TTX, a Na+ channel blocker) decreased JinK by 15 +/- 1%, but had no effect on JoutK; subsequent addition of ouabain (5 mM) decreased JinK a further 32 +/- 2%. The ouabain-induced increase in unidirectional K+ efflux did not occur when TTX was perfused prior to and during perfusion with 5 mM ouabain. We conclude that 5 mM ouabain increases the unidirectional efflux of K+ from skeletal muscle through a barium and TTX-sensitive pathway, suggestive of voltage sensitive Na+ channels, in addition to inhibiting Na+/K+-ATPase activity.  相似文献   

6.
To study the effect of an unstirred layer (UL), we have investigated the exchange efflux kinetics of anandamide at 0 degrees C, pH 7.3, from albumin-free as well as from albumin-filled human red blood cell ghosts to media of various BSA concentrations ([BSA](o)). The rate constant (k(m)) of unidirectional flux from the outer membrane leaflet to BSA in the medium increased with the square root of [BSA](o) in accordance with the existence of a UL, which is a water layer adjacent to the membrane that is not subject to the same gross mixing that takes place in the rest of the medium. From k(m), it is possible to calculate the rate constant of anandamide dissociation from BSA (k(1)) if we know the membrane binding of anandamide, the equilibrium dissociation constant of BSA-anandamide complexes, and the diffusion constant of anandamide. We estimated k(1) to be 3.33 +/- 0.27 s(-1). The net flux of [(3)H]anandamide is balanced by an equal and opposite movement of nonradioactive anandamide in exchange efflux experiments. This means that our results are also valid for uptake. We show that for anandamide with rapid membrane translocation, UL causes a significant resistance to cellular uptake. Depicting the rate of anandamide uptake as a function of equilibrium water phase concentrations results in a parabolic uptake dependence. Such apparent "saturation kinetics" is often interpreted as indicating the involvement of transport proteins. The validity of such an interpretation is discussed.  相似文献   

7.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

8.
The kinetics and thermodynamics of Al(III) exchange between aluminum citrate (AlL) and human serum transferrin were investigated in the 7.2-8.9 pH range. The C-site of human serum apotransferrin in interaction with bicarbonate removes Al(III) from Al citrate with an exchange equilibrium constant K1 = (2.0 +/- 0.6) x 10(-2); a direct second-order rate constant k1 = 45 +/- 3 M(-1) x s(-1); and a reverse second-order rate constant k(-1) = (2.3 +/- 0.5) x 10(3) M(-1) x s(-1). The newly formed aluminum-protein complex loses a single proton with proton dissociation constant K1a = (15 +/- 3) nM to yield a first kinetic intermediate. This intermediate then undergoes a modification in its conformation followed by two proton losses; first-order rate constant k2 = (4.20 +/- 0.02) x 10(-2) s(-1) to produce a second kinetic intermediate, which in turn undergoes a last slow modification in the conformation to yield the aluminum-loaded transferrin in its final state. This last process rate-controls Al(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau3(-1) = (6 +/- 1) x 10(-5) x s(-1). The affinities involved in aluminum uptake by serum transferrins are about 10 orders of magnitude lower than those involved in the uptake of iron. The interactions of iron-loaded transferrins with transferrin receptor 1 occur with average dissociation constants of 3 +/- 1 and 5 +/- 1 nM for the only C-site iron-loaded and of 6.0 +/- 0.6 and 7 +/- 0.5 nM for the iron-saturated ST in the absence or presence of CHAPS, respectively. No interaction is detected between receptor 1 and aluminum-saturated or mixed C-site iron-loaded/N-site aluminum-loaded transferrin under the same conditions. The fact that aluminum can be solubilized by serum transferrin in biological fluids does not necessarily imply that its transfer from the blood stream to cytoplasm follows the receptor-mediated pathway of iron transport by transferrins.  相似文献   

9.
Inhibition of bovine erythrocyte acetylcholinesterase (free and immobilized on controlled pore glass) by separate and simultaneous exposure to malathion and malathion transformation products which are generally formed during storage or through natural or photochemical degradation was investigated. Increasing concentrations of malathion, its oxidation product malaoxon, and its isomerisation product isomalathion inhibited free and immobilized AChE in a concentration-dependent manner. KI, the dissociation constant for the initial reversible enzyme inhibitor-complex, and k3, the first order rate constant for the conversion of the reversible complex into the irreversibly inhibited enzyme, were determined from the progressive development of inhibition produced by reaction of native AChE with malathion, malaoxon and isomalathion. KI values of 1.3 x 10(-4) M(-1), 5.6 x 10(-6) M(-1) and 7.2 x 10(-6)M(-1) were obtained for malathion, malaoxon and isomalathion, respectively. The IC50 values for free/immobilized AChE, (3.7 +/- 0.2) x 10(-4) M/(1.6 +/-0.1) x 10(-4), (2.4 +/- 0.3) x 10(-6)/(3.4 +/- 0.1) x 10(-6)M and (3.2 +/- 0.3) x 10(-6) M/(2.7 +/- 0.2) x 10(-6) M, were obtained from the inhibition curves induced by malathion, malaoxon and isomalathion, respectively. However, the products formed due to photoinduced degradation, phosphorodithioic O,O,S-trimethyl ester and O,O-dimethyl thiophosphate, did not noticeably affect enzymatic activity, while diethyl maleate inhibited AChE activity at concentrations > 10mM. Inhibition of acetylcholinesterase increased with the time of exposure to malathion and its inhibiting by-products within the interval from 0 to 5 minutes. Through simultaneous exposure of the enzyme to malaoxon and isomalathion, an additive effect was achieved for lower concentrations of the inhibitors (in the presence of malaoxon/isomalathion at concentrations 2 x 10(-7) M/2 x 10(-7) M, 2 x 10(-7) M/3 x 10(-7)M and 2 x 10(-7) M/4.5 x 109-7) M), while an antagonistic effect was obtained for all higher concentrations of inhibitors. The presence of a non-inhibitory degradation product (phosphorodithioic O,O,S-trimethyl ester) did not affect the inhibition efficiencies of the malathion by-products, malaoxon and isomalathion.  相似文献   

10.
The equilibria and kinetics of the interactions of proflavine (PR) and its platinum-containing derivative [PtCl(tmen)(2)HNC(13)H(7)(NHCH(2)CH(2))(2)](+) (PRPt) with double-stranded poly(A) have been investigated by spectrophotometry and Joule temperature-jump relaxation at ionic strength 0.1 M, 25 degrees C, and pH 5.2. Spectrophotometric measurements indicate that base-dye interactions are prevailing. T-jump experiments with polarized light showed that effects due to field-induced alignment could be neglected. Both of the investigated systems display two relaxation effects. The kinetic features of the reaction are discussed in terms of a two-step series mechanism in which a precursor complex DS(I) is formed in the fast step, which is then converted to a final complex in the slow step. The rate constants of the fast step are k(1) = (2.5 +/- 0.4) x 10(6) M(-1) s(-1), k(-1) = (2.4 +/- 0.1) x 10(3) s(-1) for poly(A)-PR and k(1) = (2.3 +/- 0.1) x 10(6) M(-1) s(-1), k(-1) = (1.6 +/- 0.2) x 10(3) s(-1) for poly(A)-PRPt. The rate constants for the slow step are k(2) = (4.5 +/- 0.5) x 10(2) s(-1), k(-2) = (1.7 +/- 0.1) x 10(2) s(-1) for poly(A)-PR and k(2) = 9.7 +/- 1.2 s(-1), k(-2) = 10.6 +/- 0.2 s(-1) for poly(A)-PRPt. Spectrophotometric measurements yield for the equilibrium constants and site size the values K = (4.5 +/- 0.1) x 10(3) M(-1), n = 1.3 +/- 0.5 for poly(A)-PR and K = (2.9 +/- 0.1) x 10(3) M(-1), n = 2.3 +/- 0.6 for poly(A)-PRPt. The values of k(1) are similar and lower than expected for diffusion-limited reactions. The values of k(-1) are similar as well. It is suggested that the formation of DS(I) involves only the proflavine residues in both systems. In contrast, the values of k(2) and k(-2) in poly(A)-PRPt are much lower than in poly(A)-PR. The results suggest that in the complex DS(II) of poly(A)-PRPt both proflavine and platinum residues are intercalated. In addition, a very slow process was detected and ascribed to the covalent binding of Pt(II) to the adenine.  相似文献   

11.
Antibodies to GHRH1-44, GHRH1-29, and proinsulin were induced in guinea pigs. GHRH1-44 forms 7 S and 10 S complexes with antibodies. It is a divalent antigen. The sequence 30-44 bound 85% of the antibodies to GHRH1-44 with high affinity (3.8 +/- 0.9 x 10(9) l/mol). The fragment 1-29 bound with low affinity (0.6 +/- 0.3 x 10(9) l/mol) 15% of the antibodies (2p less than 0.001). Antibodies to GHRH1-29 had low affinity towards the native hormone (0.4 +/- 0.2 x 10(9) l/mol) and the region 1-29 (0.3 +/- 0.2 x 10(9) l/mol). Antibodies to proinsulin bound linear C-peptide with lower affinity (0.3 +/- 0.2 x 10(9) l/mol) than the C-peptide loop in proinsulin (3.4 +/- 0.9 x 10(9) l/mol). It is concluded that the conformation of the epitopes on the sequence 1-29, recognized during the immune response, i.e. on the cell membrane, is different from the conformation of GHRH1-29 or GHRH1-44 in aqueous solution.  相似文献   

12.
Tauro JR  Lee BS  Lateef SS  Gemeinhart RA 《Peptides》2008,29(11):1965-1973
To utilize biologic mechanisms to elicit controlled release in response to disease, protease-sensitive devices have been created. Hydrogels were created with pendant peptide-drug complexes. For the matrix metalloproteases (MMPs) examined, a length of six amino acids greatly improved the specificity of the peptide (k(cat)/K(m) approximately 2.4+/-0.1x10(4)M(-1)s(-1)) over shorter sequences (k(cat)/K(m) approximately 4.4+/-0.2x10(2)M(-1)s(-1)). The peptides did not exhibit anti-proliferative effects upon cancer cells, and peptide-platinum complexes showed similar anti-proliferative effects upon the cancer cells compared to the free platinum drugs. Once the peptide-drug complex was incorporated into the hydrogels, the release was dependent upon the presence of MMP in the solution with approximately 35% of platinum released from hydrogels in the presence of MMP and only 10% without MMP in the week examined. The released drug exhibited the expected anti-proliferative activity over several days of incubation. The MMP selective drug delivery holds much potential for treatment of cancer and other diseases.  相似文献   

13.
Two-dimensional (2D) (31)P magic angle spinning (MAS) nuclear magnetic resonance (NMR) exchange spectroscopy (EXSY) demonstrated that aluminum binds to the phosphate group of phosphatidylcholine (PC) in multilamellar vesicles at pH 3.2, forming preferentially 2/1, in addition to 1/1 (PC/Al) complexes in slow exchange with one another, and with free PC, on the NMR timescale. Exchange rate constants between these three co-existing species were measured as a function of temperature using one-dimensional (1D) selective inversion recovery (SIR) (31)P MAS NMR. Over the temperature range from 5 to 35 degrees C all three exchange rate constants increased by roughly an order of magnitude from k approximately 1-2 to 10-14s(-1), exhibiting Arrhenius behavior with activation energies on the order of 30-45 kJ mol(-1) and correspondingly positive enthalpies of activation. Entropies of activation were uniformly negative, consistent with an ordered transition state. From a biological perspective, the results demonstrate that aluminum binding to PC in biomembranes is transient on a biologically relevant time scale, so that the lipid bilayer portion of biomembranes is unlikely to act as a long term repository for aluminum, but rather should be viewed as a temporary reservoir of biologically available aluminum.  相似文献   

14.
This article deals with the kinetics and thermodynamics of complex formation between Fe(3+) and a series of four synthetic chelators of the 1,2-dicatecholspermidine family (LA5, LA3, LE5 and LE3). LA5 and LA3 bear a carboxylic moiety linked to the central nitrogen by either a C(5) or a C(3) chain, whereas LE5 and LE3 bear an ethyl ester moiety. The following data concern LE5, LE3, LA5 and LA3, respectively. Each species undergoes four acid-base dissociations of the hydroxyls of the catechols with, for the two hydroxyls in position 1; average pK(2a)=7.30, 7.25, 7.45, 7.34 and, for the two hydroxyl in position 2; average pK(3a)=12.35, 12.65, 12.10, 12.60. The LA5 and LA3 species also undergo proton-dissociations of their carboxylic moieties; pK(1a)=5.20 and 5.10. The four species form one-to-one iron complexes with, for the 1-hydroxyl; an average pK(22a)=2.65, 2.25, 2.95, 2.80, for the 2-hydroxyl; pK(33a)=5.20, 5.40, 6.10, 5.40 and, for the carboxylic moieties; pK(11a)=3.90 and 4.45. In the vicinity of pH 5, Fe(3+) is rapidly exchanged between FeNta and the four ligands. This occurs with direct rate constants: k(1)=(1.3+/-0.1)x10(4), (1.4+/-0.2)x10(4), (3.3+/-0.2)x10(4), (1.4+/-0.1)x10(4)M(-1)s(-1), and reverse rate constants: k(-1)=(7+/-0.5)x10(4), (9+/-1)x10(4), (1.15+/-0.15)x10(5), (7+/-0.5)x10(4)M(-1)s(-1). The kinetic data, the pK(a) values of the free ligands, those of the iron complexes and the beta value of FeNta allow us to determine the affinity constants of the four ligands for iron: logbeta(1)=33, 34, 33, 34, and pFe=23.3, 24.6, 22.2, 24.3. This implies that these ligands of the dicatecholspermidine family may act as siderophores. They may also be used as drug carriers which can utilize the bacterial iron-acquisition paths.  相似文献   

15.
Apolipoprotein A-I was purified from human high density lipoprotein and complexed with polyunsaturated phosphatidylcholine (PC) in deoxycholate (Lipostabil); the bile salt was removed subsequently by dialysis. The behavior of the resultant apoA-I/PC complexes was compared with that of Lipostabil in vitro and after injection into rabbits. In vivo apoA-I/PC complexes had the density of HDL throughout but had both alpha and pre beta electrophoretic mobility, the latter probably reflecting dissociation of apoA-I from PC. Lipostabil initially behaved like LDL but gradually acquired the density of HDL after incubation with plasma and in vivo. Both preparations increased plasma total phospholipids in normolipidemic rabbits to a similar extent, but, increments in HDL phospholipid were greater after apoA-I/PC complexes were injected. ApoHDL/PC complexes, prepared in a similar manner, appeared to promote efflux of cholesterol from perfused rabbit aortas in the presence of lecithin:cholesterol acyltransferase (LCAT) activity, consistent with a stimulatory effect on cholesterol mobilization. Injection of apoHDL/PC complexes into hyperlipidemic rabbits decreased plasma cholesterol but increased HDL cholesterol, whereas Lipostabil decreased both. These findings suggest that human apoA-I/PC complexes resemble HDL in their behavior more closely than does Lipostabil, and show that both types of liposome undergo modification upon interaction with plasma. It remains to be shown whether they possess any therapeutic potential.  相似文献   

16.
We have designed and synthesized new optically active bisviologens ([BNMV](4+)) containing a binaphthyl moiety to examine the stereoselective photoinduced electron-transfer (ET) reactions with zinc-substituted myoglobin (ZnMb) by flash photolysis. The photoexcited triplet state of ZnMb, (3)(ZnMb)*, was successfully quenched by [BNMV](4+) ions to form the radical pair of a ZnMb cation (ZnMb(.+)) and a reduced viologen ([BNMV](.3+)), followed by a thermal ET reaction to the ground state. The rate constants ( k(q)) for the ET quenching at 25 degrees C were obtained as k(q)( R)=(2.9+/-0.2)x10(7) M(-1) s(-1) and k(q)( S)=(2.2+/-0.2)x10(7) M(-1) s(-1), respectively. The ratio of k(q)( R)/ k(q)( S)=1.3 indicates that the ( R)-isomer of the chiral viologen preferentially quenches (3)(ZnMb)*. On the other hand, the rate constants ( k) for the thermal ET reaction from [BNMV](.3+) to ZnMb(-+) at 25 degrees C were k( R)=(1.2+/-0.1)x10(8) M(-1) s(-1) and k( S)=(0.47+/-0.03)x10(8) M(-1) s(-1), respectively, and the ratio remarkably increased to k( R)/ k( S)=2.6. The activation parameters, Delta H(not equal) and Delta S(not equal), were determined from the kinetic measurements at various temperatures (10-30 degrees C) to understand the ET mechanisms. In the quenching reaction, the energy differences of Delta Delta H*(R- S) and T Delta Delta S*( R- S) at 25 degrees C were calculated to be -3.9+/-1.6 and -3.3+/-0.2 kJ mol(-1), respectively, whereas Delta Delta H*( R-S)=7.7+/-1.9 kJ mol(-1 )and T Delta Delta S*( R-S)=9.9+/-0.5 kJ mol(-1 )were found for the thermal ET reaction. Therefore, the thermal ET reaction to the ground state was proved to be dominated by the entropy term, and the large stereoselectivity may arise from the decrease in charge repulsion between donor and acceptor.  相似文献   

17.
In order to investigate the role of the plasma membrane in determining the kinetics of removal of cholesterol from cells, the efflux of [3H]cholesterol from intact cells and plasma membrane vesicles has been compared. The release of cholesterol from cultures of Fu5AH rat hepatoma and WIRL-3C rat liver cells to complexes of egg phosphatidylcholine (1 mg/ml) and human high-density apolipoprotein is first order with respect to concentration of cholesterol in the cells, with half-times (t 1/2) for at least one-third of the cell cholesterol of 3.2 +/- 0.6 and 14.3 +/- 1.5 h, respectively. Plasma membrane vesicles (0.5-5.0 micron diameter) were produced from both cell lines by incubating the cells with 50 mM formaldehyde and 2 mM dithiothreitol for 90 min. The efflux of cholesterol from the isolated vesicles follows the same kinetics as the intact, parent cells: the t 1/2 values for plasma membrane vesicles of Fu5AH and WIRL cells are 3.9 +/- 0.5 and 11.2 +/- 0.7 h, respectively. These t 1/2 values reflect the rate-limiting step in the cholesterol efflux process, which is the desorption of cholesterol molecules from the plasma membrane into the extracellular aqueous phase. The fact that intact cells and isolated plasma membranes release cholesterol at the same rates indicates that variations in the plasma membrane structure account for differences in the kinetics of cholesterol release from different cell types. In order to investigate the role of plasma membrane lipids, the kinetics of cholesterol desorption from small unilamellar vesicles prepared from the total lipid isolated from plasma membrane vesicles of Fu5AH and WIRL cells were measured. Half-times of cholesterol release from plasma membrane lipid vesicles of Fu5AH and WIRL cells were the same, with values of 3.1 +/- 0.1 and 2.9 +/- 0.2 h, respectively. Since bilayers formed from isolated plasma membrane lipids do not reproduce the kinetics of cholesterol efflux observed with the intact plasma membranes, it is likely that the local domain structure, as influenced by membrane proteins, is responsible for the differences in t 1/2 values for cholesterol efflux from these cell lines.  相似文献   

18.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
We hypothesized that nitric oxide (NO) contributes to intrasplenic fluid extravasation by inducing greater relaxation in splenic resistance arteries than veins such that intrasplenic microvascular pressure (P(C)) rises. Fluid efflux was estimated by measuring the difference between splenic blood inflow and outflow. Intrasplenic infusion of the NO donor S-nitroso-N-acetylpenicillamine (SNAP) (0.3 microg. 10 microl(-1). min(-1)) caused a significant increase in intrasplenic fluid efflux (baseline: 0.8 +/- 0.4 ml/min, n = 10 vs. peak rise during SNAP infusion: 1.3 +/- 0.4 ml/min, n = 10; P < 0.05). Intrasplenic P(C) was measured in the isolated, blood-perfused rat spleen. Intrasplenic infusion of SNAP (0.1 microg. 10 microl(-1). min(-1)) caused a significant increase in P(C) (saline: 10.9 +/- 0.2 mmHg, n = 3 vs. SNAP: 12.2 +/- 0.2 mmHg, n = 3; P < 0.05). Vasoreactivity of preconstricted splenic resistance vessels to sodium nitroprusside (SNP) (1 x 10(-12)-1 x 10(-4) M) and SNAP (1 x 10(-10)-3 x 10(-4) M) was investigated with the use of a wire myograph system. Significantly greater relaxation of arterioles than of venules occurred with both SNP (%maximal vasorelaxation: artery 96 +/- 2.3, n = 9 vs. vein 26 +/- 1.9, n = 10) and SNAP (%maximal vasorelaxation: artery 50 +/- 3.5, n = 11 vs. vein 32 +/- 1.7, n = 8). These results are consistent with our proposal that differential vasoreactivity of splenic resistance arteries and veins to NO elevates intrasplenic P(C) and increases fluid extravasation into the systemic lymphatic system.  相似文献   

20.
The effect of intravesicular and extravesicular calcium concentration on the passive efflux from sarcoplasmic reticulum (SR) vesicles isolated from cardiac and skeletal muscle was determined by measuring net efflux of calcium after stopping pump-mediated fluxes. The apparent permeability, calculated as the passive efflux divided by the total intravesicular calcium, depended on calcium load. This dependence of the apparent permeability on calcium load could be explained by the presence of intravesicular calcium-binding sites with a dissociation constant less than 10(-3) M. When the intravesicular bound calcium was taken into account, passive calcium efflux was found to be linearly related to the difference in calcium concentration across the SR membrane. Thus the permeability of the SR membrane is independent of intravesicular and extravesicular calcium concentration in the ranges investigated. The average first order rate constant for passive calcium efflux for six preparations was 0.8 +/- 0.2 min-1 for skeletal and 0.7 +/- 0.1 min-1 for cardiac SR. The amount of intravesicular bound calcium for the same preparations was 33 +/- 6 nmol mg-1 for skeletal and 13 +/- 2 nmol mg-1 for cardiac SR. The first order rate constants were unaffected by Mg concentration between 0.1 +/- 15.1 mM and by the presence of an ATP-regenerating system. The results suggest that some minimal calcium load may be required in order to observe a substantial passive calcium efflux, the passive calcium efflux is not carrier mediated, and passive calcium efflux is not a likely route of calcium release during excitation-contraction coupling.  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