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 use of resealed red blood cell membranes (ghosts) allows the study of the transport of a compound in a nonmetabolizing system with a biological membrane. Transmembrane movements of anandamide (N-arachidonoylethanolamine, arachidonoylethanolamide) have been studied by exchange efflux experiments at 0 degrees C and pH 7.3 with albumin-free and albumin-filled human red blood cell ghosts. The efflux kinetics is biexponential and is analyzed in terms of compartment models. The distribution of anandamide on the membrane inner to outer leaflet pools is determined to be 0.275 +/- 0.023, and the rate constant of unidirectional flux from inside to outside is 0.361 +/- 0.023 s(-1). The rate constant of unidirectional flux from the membrane to BSA in the medium ([BSA]o) increases with the square root of [BSA]o in accordance with the theory of an unstirred layer around ghosts. Anandamide passed through the red blood cell membrane very rapidly, within seconds. At a molar ratio of anandamide to BSA of <1, membrane binding of anandamide increases with increasing temperatures between 0 degrees C and 37 degrees C, and the equilibrium dissociation constants are in the nanomolar range. The nature of membrane binding and the mechanism of membrane translocation are discussed.  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.
The whole-cell voltage-clamp technique was used in rat cardiac myocytes to investigate the kinetics of ADP binding to phosphorylated states of Na,K-ATPase and its effects on presteady-state Na(+)-dependent charge movements by this enzyme. Ouabain-sensitive transient currents generated by Na,K-ATPase functioning in electroneutral Na(+)-Na(+) exchange mode were measured at 23 degrees C with pipette ADP concentrations ([ADP]) of up to 4.3 mM and extracellular Na(+) concentrations ([Na](o)) between 36 and 145 mM at membrane potentials (V(M)) from -160 to +80 mV. Analysis of charge-V(M) curves showed that the midpoint potential of charge distribution was shifted toward more positive V(M) both by increasing [ADP] at constant Na(+)(o) and by increasing [Na](o) at constant ADP. The total quantity of mobile charge, on the other hand, was found to be independent of changes in [ADP] or [Na](o). The presence of ADP increased the apparent rate constant for current relaxation at hyperpolarizing V(M) but decreased it at depolarizing V(M) as compared to control (no added ADP), an indication that ADP binding facilitates backward reaction steps during Na(+)-Na(+) exchange while slowing forward reactions. Data analysis using a pseudo three-state model yielded an apparent K(d) of approximately 6 mM for ADP binding to and release from the Na,K-ATPase phosphoenzyme; a value of 130 s(-1) for k(2), a rate constant that groups Na(+) deocclusion/release and the enzyme conformational transition E(1) approximately P --> E(2)-P; a value of 162 s(-1)M(-1) for k(-2), a lumped second-order V(M)-independent rate constant describing the reverse reactions; and a Hill coefficient of approximately 1 for Na(+)(o) binding to E(2)-P. The results are consistent with electroneutral release of ADP before Na(+) is deoccluded and released through an ion well. The same approach can be used to study additional charge-moving reactions and associated electrically silent steps of the Na,K-pump and other transporters.  相似文献   

4.
Pre-steady-state and steady-state kinetics of nucleotide incorporation and excision were used to assess potential mechanisms by which the fidelity of the herpes simplex virus type 1 DNA polymerase catalytic subunit (Pol) is enhanced by its processivity factor, UL42. UL42 had no effect on the pre-steady-state rate constant for correct nucleotide incorporation (150 s(-1)) nor on the primary rate-limiting conformational step. However, the equilibrium dissociation constant for the enzyme in a stable complex with primer-template was 44 nm for Pol and 7.0 nm for Pol/UL42. The catalytic subunit and holoenzyme both selected against incorrect nucleotide incorporation predominantly at the level of nucleotide affinity, although UL42 slowed by 4-fold the maximum rate of incorporation of incorrect, compared with correct, nucleotide. Pol, with or without UL42, cleaved matched termini at a slower rate than mismatched ones, but UL42 did not significantly alter the pre-steady-state rate constant for mismatch excision ( approximately 16 s(-1)). The steady-state rate constant for nucleotide addition was 0.09 s(-1) and 0.03 s(-1) for Pol and Pol/UL42, respectively, and enzyme dissociation was the rate-limiting step. The longer half-life for DNA complexes with Pol/UL42 (23 s) compared with that with Pol (8 s) affords a greater probability for excision when a misincorporation event does occur, accounting predominantly for the failure of Pol/UL42 to accumulate mismatched product at moderate nucleotide concentrations.  相似文献   

5.
[3H]Palmitate, PA, exchange efflux kinetics is recorded from human erythrocyte ghosts to buffer with bovine serum albumin, BSA, at 0 degrees C. The effects have been investigated of three medium/ghost volume ratios: 36, 80 and 500, of six BSA concentrations, [BSA]: 0.01, 0.02, 0.05, 0.2, 1 and 2% (1.5, 3.0, 7.5, 30, 150 and 300 microM) and of various v, molar ratios of palmitate to BSA, between 0.15 and 0.94. Data are analyzed in terms of a virtually closed three-compartment model. In theory, the tracer efflux is biexponential and the rate coefficients differ at least 20 fold [1]. The efflux rate at 2% BSA is monoexponential beyond our resolution time of about 1 s, but nearly biexponential at or below 0.2% BSA with a well-defined smallest-rate coefficient beta. beta depends strongly on [BSA] but is remarkably v independent. The medium/ghost volume ratio has no effect on beta when [BSA] > or = 0.2%, although beta measured at 2% BSA is almost 2-fold higher than at 0.2%. This suggests the presence of an unstirred layer, USL. According to our model, the observations are understood quantitatively on basis of our previously published dissociation rate constants of the PA-BSA complex, as well as PA equilibrium bindings to ghost membranes (Bojesen, I.N. and Bojesen, E. (1991) Biochim. Biophys. Acta 1069, 297-307). Essentially, beta is theoretically a function of two terms, one comprising the membrane transport parameters and the other the medium-dependent variables. Most important is the clearance with respect to monomer concentration adjacent to the membrane. The clearance is calculated on basis of quasi-stationary diffusion in USL. The data are compatible with a planar USL of 6 microns depth and with the same area as a ghost but not with a spherical USL.  相似文献   

6.
Major zinc binding protein purified from renal brush border membrane (BBM) (R. Kumar, R. Prasad, Biochim. Biophys. Acta 1419 (1999) 23) was reconstituted into liposomes and its functional characteristics were investigated. Physical incorporation of the major zinc binding protein into the proteoliposomes was checked by SDS-PAGE, which showed a single band on silver staining. The structural integrity of the proteoliposomes was assessed by phase contrast microscopy, which revealed the proteoliposomes as globular structures and intact boundaries. Further structural integrity/leakiness of the proteoliposomes was checked by monitoring efflux of Zn(2+) from the pre-loaded proteoliposomes in the presence of either 2 mM Ca(2+) or Cd(2+) or Zn(2+). It was observed that even after 2 h of the initiation of efflux, 85-95% of Zn(2+) was retained in the proteoliposomes, thereby indicating that proteoliposomes were not leaky and maintained structural integrity during the uptake study. Zinc uptake into the proteoliposomes followed Michaelis-Menten kinetics with affinity constant (K(m)) of 1.03 mM and maximal velocity (V(max)) of 1333 nmol/mg protein per min. The uptake process followed first-order kinetics with a rate constant (k) of 1. 09x10(-3) s(-1). The specificity of zinc transport system was determined by studying the interaction of divalent cations viz. Ca(2+) and Cd(2+) with the zinc uptake. It was observed that Cd(2+) competitively inhibited the zinc uptake process with inhibitory concentration (K(i)) of 2.9 mM. Kinetic analysis of inhibitory effect of Cd(2+) on zinc uptake revealed an increase in K(m) to 1.74 mM without influencing V(max). Zn(2+) uptake into the proteoliposomes was found to be temperature sensitive and Arrhenius plot showed a breakpoint at 27 degrees C. The apparent energies of activation (E(a)) were found to be 7.09 and 2.74 kcal/mol below and above the breakpoint, respectively. The initial velocity of Zn(2+) uptake increased with the increase in outwardly directed proton gradient ([H](i) greater than [H](o)). The Zn(2+) uptake was inhibited by DCCD, thereby suggesting the involvement of -COOH groups in the translocation of Zn(2+) across the lipid bilayer. The ratio of acidic to basic amino acids (1.26) strongly indicates that it is an acidic protein. The cysteine content in this protein was insignificant, which further corroborates the possibility that the acidic amino acids might be prominent candidates for binding to zinc. The findings of the present study confirms that 40 kDa major zinc binding glycoprotein purified from renal BBM is a zinc transporter involved in the influx of Zn(2+) into the epithelial cells of the renal tubular system.  相似文献   

7.
Transport of monosaccharides in kidney-cortex cells   总被引:5,自引:4,他引:1  
1. The aerobic transport of d-glucose and d-galactose in rabbit kidney tissue at 25 degrees was studied. 2. In slices forming glucose from added substrates an accumulation of glucose against its concentration gradient was found. The apparent ratio of intracellular ([S](i)) and extracellular ([S](o)) glucose concentrations was increased by 0.4mm-phlorrhizin and 0.3mm-ouabain. 3. Slices and isolated renal tubules actively accumulated glucose from the saline; the apparent [S](i)/[S](o) fell below 1.0 only at [S](o) higher than 0.5mm. 4. The rate of glucose oxidation by slices was characterized by the following parameters: K(m) 1.16mm; V(max.) 4.5mumoles/g. wet wt./hr. 5. The active accumulation of glucose from the saline was decreased by 0.1mm-2,4-dinitrophenol, 0.4mm-phlorrhizin and by the absence of external Na(+). 6. The kinetic parameters of galactose entry into the cells were: K(m) 1.5mm; V(max) 10mumoles/g. wet wt./hr. 7. The efflux kinetics from slices indicated two intracellular compartments for d-galactose. The galactose efflux was greatly diminished at 0 degrees , was inhibited by 0.4mm-phlorrhizin, but was insensitive to ouabain. 8. The following mechanism of glucose and galactose transport in renal tubular cells is suggested: (a) at the tubular membrane, these sugars are actively transported into the cells by a metabolically- and Na(+)-dependent phlorrhizin-sensitive mechanism; (b) at the basal cell membrane, these sugars are transported in accordance with their concentration gradient by a phlorrhizin-sensitive Na(+)-independent facilitated diffusion. The steady-state intracellular sugar concentration is determined by the kinetic parameters of active entry, passive outflow and intracellular utilization.  相似文献   

8.
The small GTPase Rac cycles between the membrane and the cytosol as it is activated by nucleotide exchange factors (GEFs) and inactivated by GTPase-activating proteins (GAPs). Solubility in the cytosol is conferred by binding of Rac to guanine-nucleotide dissociation inhibitors (GDIs). To analyze the in vivo dynamics of Rac, we developed a photobleaching method to measure the dissociation rate constant (k(off)) of membrane-bound GFP-Rac. We find that k(off) is 0.048 s(-1) for wtRac and approximately 10-fold less (0.004 s(-1)) for G12VRac. Thus, the major route for dissociation is conversion of membrane-bound GTP-Rac to GDP-Rac; however, dissociation of GTP-Rac occurs at a detectable rate. Overexpression of the GEF Tiam1 unexpectedly decreased k(off) for wtRac, most likely by converting membrane-bound GDP-Rac back to GTP-Rac. Both overexpression and small hairpin RNA-mediated suppression of RhoGDI strongly affected the amount of membrane-bound Rac but surprisingly had only slight effects on k(off). These results indicate that RhoGDI controls Rac function mainly through effects on activation and/or membrane association.  相似文献   

9.
Heme transfer between phospholipid membranes and uptake by apohemoglobin   总被引:4,自引:0,他引:4  
The incorporation of CO-heme into single bilayer, egg lecithin vesicles was examined by following the spectral changes that occur when the porphyrin becomes embedded in the membranes. The rate of CO-heme uptake by liposomes is extremely fast (t1/2 less than or equal to 20 ms at 10 degrees C), and the maximum extent is roughly 1 heme/5 phospholipid molecules. This limiting stoichiometry is due to unfavorable electrostatic interactions between the propionate groups of the bound CO-heme. This effect was treated theoretically by attenuating the intrinsic heme partitioning equilibrium constant with an exponential term reflecting the surface potential of the membranes. The surface potential was assumed to be proportional to the concentration of CO-heme in the membranes, and the final expression is Kp = Kop exp[-AHb/VpCp], where Kp is the observed partition constant; Kop, the intrinsic constant; Hb, the concentration of bound heme in the suspension; Vp, the partial molar volume of egg lecithin; Cp, the concentration of lipid phosphate; and A, an empirical constant representing the capacitance of the membrane for heme. For the analysis of kinetic data, the electrostatic term is assumed to apply only to the membrane dissociation rate constant, k-1, and not the association rate constant, k1. The dissociation rate was measured independently either by following the transfer of CO-heme from one vesicle fraction to another or by monitoring heme efflux from the membranes and incorporation into apohemoglobin at high protein concentrations. The data for all three sets of experiments, heme uptake, transfer, and incorporation into globin at 10 degrees C, were fitted quantitatively to the partitioning mechanism using A = 15 M-1, Kop = 5 X 10(5), k1 = 2 X 10(6) s-1, and k0(-1) = 4 s-1. Thus, heme can spontaneously migrate across lipid-water interfaces and hence diffuse rapidly from the mitochondrial inner membrane where it is synthesized to the rough endoplasmic reticulum where it is incorporated into hemoglobin.  相似文献   

10.
We have investigated the kinetics of interaction of cationic fluorescent lipophiles (dyes) rhodamine 123, rhodamine 6G, tetramethyl rhodamine ethyl ester, safranine O, 1,1'-diethyloxacarbocyanine, 1,1'-diethyloxadicarbocyanine, and 1,1'-diethylthiadicarbocyanine iodide with isolated respiring rat-liver mitochondria (RLM). Dye flux across the RLM inner membrane was measured by following the kinetics of fluorescence signal change after mixing of dye and RLM. The time course of fluorescence was analysed in terms of a kinetic model of the binding and transport processes involved. The rate constants of dye influx and efflux were extracted from the observed effect on the apparent time constant of fluorescence change to equilibrium intensity upon mixing dye with increasing concentrations of RLM. From the influx rate constants obtained, the apparent permeability constants for dye influx (at zero potential) across the membrane were calculated and ranged from 3 to 140 x 10(-4) cm/s. The influx rate constant was found to be linearly related to relative dye lipophilicity, as predicted by the model. As another test of the model, from the ratio of the influx and efflux rate constants, the apparent trans-membrane potential, psi, was calculated and found generally to agree with reported values, but to depend on the lipophilicity of the dye used. Not predicted by the simple model was a dissymmtry observed in the influx and efflux time constants for fluorescence change to equilibrium intensity. Inferences are made relating to the utility of these dyes as probes of psi.  相似文献   

11.
The endocannabinoid anandamide is of lipid nature and may thus bind to albumin in the vascular system, as do fatty acids. The knowledge of the free water-phase concentration of anandamide is essential for the investigations of its transfer from the binding protein to cellular membranes, because a water-phase shuttle of monomers mediates such transfers. We have used our method based upon the use of albumin-filled red cell ghosts as a dispersed biological "reference binder" to measure the water-phase concentrations of anandamide. These concentrations were measured in buffer (pH 7.3) in equilibrium with anandamide bound to BSA inside resealed human red cell membranes at low molar ratios below one. Data were obtained at 0 degrees C, 10 degrees C, 23 degrees C, and 37 degrees C. The equilibrium dissociation constant (Kd) increases with temperature from 6.87 +/- 0.53 nM at 0 degrees C to 54.92 +/- 1.91 nM at 37 degrees C. Regression analyses of the data suggest that BSA has one high-affinity binding site for anandamide at all four temperatures. The free energy of anandamide binding (DeltaG0) is calculated to -43.05 kJ mol-1 with a large enthalpy (DeltaH0) contribution of -42.09 kJ mol-1. Anandamide has vasodilator activity, and the binding to albumin may mediate its transport in aqueous compartments.  相似文献   

12.
The influences of [Ca(2+)] and Ca(2+) dissociation rate from troponin C (TnC) on the kinetics of contraction (k(Ca)) activated by photolysis of a caged Ca(2+) compound in skinned fast-twitch psoas and slow-twitch soleus fibers from rabbits were investigated at 15 degrees C. Increasing the amount of Ca(2+) released increased the amount of force in psoas and soleus fibers and increased k(Ca) in a curvilinear manner in psoas fibers approximately 5-fold but did not alter k(Ca) in soleus fibers. Reconstituting psoas fibers with mutants of TnC that in solution exhibited increased Ca(2+) affinity and approximately 2- to 5-fold decreased Ca(2+) dissociation rate (M82Q TnC) or decreased Ca(2+) affinity and approximately 2-fold increased Ca(2+) dissociation rate (NHdel TnC) did not affect maximal k(Ca). Thus the influence of [Ca(2+)] on k(Ca) is fiber type dependent and the maximum k(Ca) in psoas fibers is dominated by kinetics of cross-bridge cycling over kinetics of Ca(2+) exchange with TnC.  相似文献   

13.
An analytical solution is given for the kinetics of reversible homogeneous one-dimensional growth, assuming that all association rate constants have the same value k, that all dissociation rate constants are likewise equal to k, and that the monomer concentration has a constant value, C. Such growth tends to generate a maximally polydisperse ("white") distribution of cluster concentrations ci, all approaching a limiting value equal to that of the critical nucleus, cn. Continued growth merely increases the range of cluster sizes over which this white distribution applies. A simple expression is obtained for the flux sigma infinity i = n dci/dt, which becomes constant and equal to (kC - k)cn. The monomer uptake increases with time, and is given approximately by (kC - k)2cnt.  相似文献   

14.
Protein-protein recognition results from the assembly of complementary surfaces on two molecules that form a stable, noncovalent, specific complex. Our interest was to describe kinetic aspects of the recognition in order to understand the subtle molecular mechanism of association. R67 dihydrofolate reductase (DHFR) provides an ideal model to investigate kinetic parameters of protein-protein association since it is a homotetramer resulting from the pH-dependent dimerization of homodimers. We took advantage of the presence of a tryptophan residue at the dimer-dimer interface to monitor pH-dependent oligomerization of R67 DHFR using stopped-flow fluorescence techniques. Except for pH near neutrality where dissociation exhibited biphasic kinetics, association and dissociation followed monophasic kinetics fitted on a two-state model. Apparent rate constants of association k(on) and dissociation k(off) were determined at various pHs and pointed to the key role of a histidine located at the dimer-dimer interface in the pH control of tetramerization. The values of the tetramer-dimer equilibrium dissociation constant were calculated from the ratio k(off) /k(on) and correlated well with those previously measured at equilibrium. The thermodynamic parameters and the activation energies of both the association and dissociation were determined and indicated that the association is enthalpy driven and suggested that the formation of four hydrogen bonds (one per monomer) is responsible for the thermodynamic stability of the tetramer. Detailed analysis of the biphasic kinetics led to an original model, in which protonation of the tetramer is the triggering event for the dissociation process while the association involves primarily the unprotonated dimers.  相似文献   

15.
The kinetics and thermodynamics of Bi(III) exchange between bismuth mononitrilotriacetate (BiL) and human serum transferrin as well as those of the interaction between bismuth-loaded transferrin and transferrin receptor 1 (TFR) were investigated at pH 7.4-8.9. Bismuth is rapidly exchanged between BiL and the C-site of human serum apotransferrin in interaction with bicarbonate to yield an intermediate complex with an effective equilibrium constant K(1) of 6 +/- 4, a direct second-order rate constant k(1) of (2.45 +/- 0.20) x 10(5) M(-1) s(-1), and a reverse second-order rate constant k(-1) of (1.5 +/- 0.5) x 10(6) M(-1) s(-1). The intermediate complex loses a single proton with a proton dissociation constant K(1a) of 2.4 +/- 1 nM to yield a first kinetic product. This product then undergoes a modification in its conformation followed by two proton losses with a first-order rate constant k(2) = 25 +/- 1.5 s(-1) to produce a second kinetic intermediate, which in turn undergoes a last modification in the conformation to yield the bismuth-saturated transferrin in its final state. This last process rate-controls Bi(III) uptake by the N-site of the protein and is independent of the experimental parameters with a constant reciprocal relaxation time tau(3)(-1) of (3 +/- 1) x 10(-2) s(-1). The mechanism of bismuth uptake differs from that of iron and probably does not involve the same transition in conformation from open to closed upon iron uptake. The interaction of bismuth-loaded transferrin with TFR occurs in a single very fast kinetic step with a dissociation constant K(d) of 4 +/- 0.4 microM, a second-order rate constant k(d) of (2.2 +/- 1.5) x 10(8) M(-1) s(-1), and a first-order rate constant k(-d) of 900 +/- 400 s(-1). This mechanism is different from that observed with the ferric holotransferrin and implies that the interaction between TFR and bismuth-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 bismuth incorporation by the transferrin receptor-mediated iron acquisition pathway is discussed.  相似文献   

16.
17.
The kinetics (0 to 3 h) of cholesterol efflux to delipidated apolipoprotein A-1 were investigated, and the experimental data were best fitted to a mathematical model that involves two independent pathways of cholesterol efflux. The first pathway with a rate constant of 4.6 h(-1) is fast but removes only 3-5% of total cholesterol. After preconditioning apoA-1, it was found that this pathway remains, and hence it is a property of the cholesterol-loaded cells rather than due to modification on the apolipoprotein. This fast initial efflux does not seem to contribute to cholesterol efflux at later stages (>1 h) where a second pathway predominates. However, the fast initial efflux pool can be restored if apoA-1 is withdrawn. The second slower pathway (k(membrane--media) = 0.79 h(-1)) is associated with cholesterol ester hydrolysis whose rate constant could be experimentally verified (k(cal) = 0.43, k(exp) = 0.38 +/- 0.05). The model suggests that two different plasma membrane domains are involved in the two pathways. Loading of the cells with an oxysterol, 7-ketocholesterol (7K), inhibits efflux from both pathways. The model predicts that 7K decreases the initial efflux by decreasing the available cholesterol (by possibly affecting lipid packing), while all rate constants in the second pathway are decreased. In conclusion, the kinetic model suggests that cholesterol efflux to apoA-1 is a two-step process. In the first step, some of the plasma membrane cholesterol contributes to a fast initial efflux (possibly from lipid rafts) and leads to a second pathway that mobilizes intracellular cholesterol mobilization.  相似文献   

18.
The problem of predicting the kinetics of proton efflux and the decay of the internal proton concentration for vesicles containing one or more buffers for which the internal proton concentration is initially higher than that of the surrounding medium is examined. An analytical solution is derived that describes the time course of the proton efflux from vesicles and the decay of the internal proton concentration under conditions of zero transmembrane electric potential. The effect of the internal buffers is to increase the time required for the proton concentration gradient to equilibrate across the membrane. To simplify the analysis we assume that the equilibration of the internal and external proton activity is due primarily to proton diffusion through the membrane, and not to hydroxyl ion flux. For a vesicle containing a single buffer the solution requires six independent physical parameters: the initial internal proton concentration, the external proton concentration, the ratio of the vesicle surface area to the internal volume, the permeability coefficient of the membrane for protons, the total concentration of the internal buffer, and the equilibrium constant for the dissociation of the internal buffer. Determination of these physical values is sufficient to predict the time dependence of the internal proton concentration and of the proton efflux. Over a pH range that is below or near the pK of the internal buffer the solution is complex. However, if the initial pH is one unit or more higher than the pK of the internal buffer the kinetics of the internal proton concentration and proton efflux can be described by a pseudo first order reaction. In this case the apparent rate constant depends linearly on the permeability coefficient and is dominated by the total internal buffer concentration and its pK. For example, increasing the internal buffer concentration inside a vesicle by 10-fold results in an approximately 10-fold increase in the half-time of the proton efflux kinetics. The theoretical analysis is applied to thylakiod vesicles using experimentally determined values for the physical parameters. The predictions of the analysis are compared to experimentally observed kinetics.  相似文献   

19.
L J Maher  P B Dervan  B J Wold 《Biochemistry》1990,29(37):8820-8826
Pyrimidine oligonucleotides recognize extended purine sequences in the major groove of double-helical DNA by triple-helix formation. The resulting local triple helices are relatively stable and can block DNA recognition by sequence-specific DNA binding proteins such as restriction endonucleases. Association and dissociation kinetics for the oligodeoxyribonucleotide 5'-CTCTTTCCTCTCTTTTTCCCC (bold C's indicate 5-methylcytosine residues) are now measured with a restriction endonuclease protection assay. When oligonucleotides are present in greater than 10-fold excess over the DNA target site, the binding reaction kinetics are pseudo first order in oligonucleotide concentration. Under our standard conditions (37 degrees C, 25 mM Tris-acetate, pH 6.8, 70 mM sodium chloride, 20 mM magnesium chloride, 0.4 mM spermine tetrahydrochloride, 10 mM beta-mercaptoethanol, 0.1 mg/mL bovine serum albumin) the value of the observed pseudo-first-order association rate constant, k2obs, is 1.8 x 10(3) +/- 1.9 x 10(2) L.(mol of oligomer-1.s-1. Measurement of the dissociation rate constant yields an equilibrium dissociation constant of approximately 10 nM. Increasing sodium ion concentration slightly decreased the association rate, substantially increased the dissociation rate, and thereby reduced the equilibrium binding constant. This effect was reversible by increasing multivalent cation concentration, confirming the significant role of multivalent cations in oligonucleotide-directed triple-helix formation under these conditions. Finally, a small reduction in association rate, a large increase in dissociation rate, and a resulting reduction in the equilibrium binding constant were observed upon increasing the pH between 6.8 and 7.2.  相似文献   

20.
Glucose transport in plasma membranes is the prototypic example of facilitated diffusion through biological membranes, and transport in erythrocytes is the most widely studied. One of the oldest and simplest models describing the kinetics of the transport reaction is that of alternating conformers, schematized in a cycle of four partial reactions where glucose binds and dissociates at two opposite steps, and the transporter undergoes transconformations at the other two opposite steps. The transport kinetics is entirely defined by the forward and backward rate constants of the partial reactions and the glucose and transporter concentrations at each side of the membrane, related by the law of mass action. We studied, in silico, the effect of modifications of the variables on the transient kinetics of the transport reaction. The simulations took into account thermodynamic constraints and provided results regarding initial velocities of transport, maximal velocities in different conditions, apparent influx and efflux affinities, and the turnover number of the transporter. The results are in the range of those experimentally reported. Maximal initial velocities are obtained when the affinities of the ligand for the transporter are the same at the extra- and intracellular binding sites and when the equilibrium constants of the transconformation steps are equal among them and equal to 1, independently of the obvious effect of the increase of the rate constant values. The results are well adjusted to Michaelis–Menten kinetics. A larger initial velocity for efflux than for uptake described in human erythrocytes is demonstrated in a model with the same dissociation constants at the outer and inner sites of the membrane. The larger velocities observed for uptake and efflux when transport occurs towards a glucose-containing trans side can also be reproduced with the alternating conformer model, depending on how transport velocities are measured.  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