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.
G Renger  B Hanssum 《FEBS letters》1992,299(1):28-32
The temperature dependence of the rate constants of the univalent redox steps YzoxSi----YzSi + 1 (i = 0,1,2) and YzoxS3----(YzS4)----YzSo + O2 in the water oxidase was investigated by measuring time resolved absorption changes at 355 nm induced by a laser flash train in dark adapted PS II membrane fragments from spinach. Activation energies of 5.0, 12.0 and 36.0 kJ/mol were obtained for the reactions YzoxSi----YzSi + 1 with i = 0,1 and 2, respectively. The reaction YzoxS3----(YzS4)----YzS0 + O2 exhibits a temperature dependence with a characteristic break point at 279 K with activation energies of 20 kJ/mol (T greater than 279 K) and 46 kJ/mol (T less than 279 K). Evaluation of the data within the framework of the classical Marcus theory of nonadiabatic electron transfer [(1985) Biochim. Biophys. Acta 811, 265-322] leads to the conclusion that the S2 oxidation to S3 is coupled with significant structural changes. Furthermore, the water oxidase in S3 is inferred to attain two different conformational states with populations that markedly change at a characteristic transition temperature.  相似文献   

2.
The Michaelis constant of membrane-bound adenylate cyclase increased from 1.1 to 1.8 mM between 7 and 38 degrees C (delta H = 13 kJ/mol). Over this temperature range, the maximum velocity increased 10-fold, and the Arrhenius plot was nearly linear, with an average delta H* of 51 kJ/mol. The temperature-dependence of the reaction rate at 2 mM-ATP was examined in more detail: for Lubrol-dispersed enzyme, Arrhenius plots were nearly linear with average delta H* values of 45 and 68 kJ/mol, respectively, for untreated and gel-filtered enzymes; for membrane-bound enzyme, delta H changed from 40 kJ/mol above about 21 degrees C to 62 kJ/mol below 21 degrees C, but this behaviour does not necessarily indicate an abrupt, lipid-induced, transition in the reaction mechanism.  相似文献   

3.
Dodson ML  Walker RC  Lloyd RS 《PloS one》2012,7(2):e31377
In order to suggest detailed mechanistic hypotheses for the formation and dehydration of a key carbinolamine intermediate in the T4 pyrimidine dimer glycosylase (T4PDG) reaction, we have investigated these reactions using steered molecular dynamics with a coupled quantum mechanics-molecular mechanics potential (QM/MM). We carried out simulations of DNA abasic site carbinolamine formation with and without a water molecule restrained to remain within the active site quantum region. We recovered potentials of mean force (PMF) from thirty replicate reaction trajectories using Jarzynski averaging. We demonstrated feasible pathways involving water, as well as those independent of water participation. The water-independent enzyme-catalyzed reaction had a bias-corrected Jarzynski-average barrier height of approximately (6.5 kcal mol(-1) (27.2 kJ mol(-1)) for the carbinolamine formation reaction and 44.5 kcal mol(-1) (186 kJ mol(-1)) for the reverse reaction at this level of representation. When the proton transfer was facilitated with an intrinsic quantum water, the barrier height was approximately 15 kcal mol(-1) (62.8 kJ mol(-1)) in the forward (formation) reaction and 19 kcal mol(-1) (79.5 kJ mol(-1)) for the reverse. In addition, two modes of unsteered (free dynamics) carbinolamine dehydration were observed: in one, the quantum water participated as an intermediate proton transfer species, and in the other, the active site protonated glutamate hydrogen was directly transferred to the carbinolamine oxygen. Water-independent unforced proton transfer from the protonated active site glutamate carboxyl to the unprotonated N-terminal amine was also observed. In summary, complex proton transfer events, some involving water intermediates, were studied in QM/MM simulations of T4PDG bound to a DNA abasic site. Imine carbinolamine formation was characterized using steered QM/MM molecular dynamics. Dehydration of the carbinolamine intermediate to form the final imine product was observed in free, unsteered, QM/MM dynamics simulations, as was unforced acid-base transfer between the active site carboxylate and the N-terminal amine.  相似文献   

4.
An intramolecular electron-transfer process has previously been shown to take place between the Cys3--Cys26 radical-ion (RSSR-) produced pulse radiolytically and the Cu(II) ion in the blue single-copper protein, azurin [Farver, O. & Pecht, I. (1989) Proc. Natl Acad. Sci. USA 86, 6868-6972]. To further investigate the nature of this long-range electron transfer (LRET) proceeding within the protein matrix, we have now investigated it in two azurins where amino acids have been substituted by single-site mutation of the wild-type Pseudomonas aeruginosa azurin. In one mutated protein, a methionine residue (Met44) that is proximal to the copper coordination sphere has been replaced by a positively charged lysyl residue ([M44K]azurin), while in the second mutant, another residue neighbouring the Cu-coordination site (His35) has been replaced by a glutamine ([H35Q]azurin). Though both these substitutions are not in the microenvironment separating the electron donor and acceptor, they were expected to affect the LRET rate because of their effect on the redox potential of the copper site and thus on the driving force of the reaction, as well as on the reorganization energies of the copper site. The rate of intramolecular electron transfer from RSSR- to Cu(II) in the wild-type P. aeruginosa azurin (delta G degrees = -68.9 kJ/mol) has previously been determined to be 44 +/- 7 s-1 at 298 K, pH 7.0. The [M44K]azurin mutant (delta G degrees = -75.3 kJ/mol) was now found to react considerably faster (k = 134 +/- 12 s-1 at 298 K, pH 7.0) while the [H35Q]azurin mutant (delta G degrees = -65.4 kJ/mol) exhibits, within experimental error, the same specific rate (k = 52 +/- 11 s-1, 298 K, pH 7.0) as that of the wild-type azurin. From the temperature dependence of these LRET rates the following activation parameters were calculated: delta H++ = 37.9 +/- 1.3 kJ/mol and 47.2 +/- 0.7 kJ/mol and delta S++ = -86.5 +/- 5.8 J/mol.K and -46.4 +/- 4.4 J/mol.K for [H35Q]azurin and [M44K]azurin, respectively. Using the Marcus relation for intramolecular electron transfer and the above parameters we have determined the reorganization energy, lambda and electronic coupling factor, beta. The calculated values fit very well with a through-bond LRET mechanism.  相似文献   

5.
The effect of temperature on the activation energies of mitochondrial enzymes of the yeast Saccharomyces cerevisiae was examined. Non-linear Arrhenius plots with discontinuities in the temperature range 14-19 degrees C and 19-22 degrees C were observed for the respiratory enzymes and mitochondrial ATPase (adenosine triphosphatase) respectively. A straight-line Arrhenius plot was observed for the matrix enzyme, malate dehydrogenase. The activation energies of the enzymes associated with succinate oxidation, namely, succinate oxidase, succinate dehydrogenase and succinate-cytochrome c oxidoreductase, were in the range 60-85kJ/mol above the transition temperature and 90-160kJ/mol below the transition temperature. In contrast, the corresponding enzymes associated with NADH oxidation showed significantly lower activation energies, 20-35kJ/mol above and 40-85kJ/mol below the transition temperature. The discontinuities in the Arrhenius plots were still observed after sonication, treatment with non-ionic detergents or freezing and thawing of the mitochondrial membranes. Discontinuities for cytochrome c oxidase activity were only observed in freshly isolated mitochondria, and no distinct breaks were observed after storage at -20 degrees C. Mitochondrial ATPase activity still showed discontinuities after sonication and freezing and thawing, but a linear plot was observed after treatment with non-ionic detergents. The results indicate that the various enzymes of the respiratory chain are located in a similar lipid macroenvironment within the mitochondrial membrane.  相似文献   

6.
The relationship between rate of ATP synthesis, JATP, and value of the proton electrochemical gradient, delta mu H, has been analyzed in intact mitochondria. Onset of phosphorylation causes a depression of delta mu H of 1.5 kJ/mol. There is a close parallelism between inhibition of JATP and restoration of delta mu H to its state-4 value during titrations with oligomycin or atractyloside. Titrations with ionophores display the following features: (a) delta mu H can be depressed by 3-4 kJ/mol by valinomycin + K+ without affecting the rate of ATP synthesis; (b) uncouplers abolish JATP completely while depressing delta mu H by 3 kJ/mol; (c) complete abolition of ATP synthesis by inhibitors of electron transport is accompanied by a depression of delta mu H of only 1 kJ/mol. The results indicate that: (a) there is a close functional relationship between redox and ATPase H+ pumps, whereby inhibition of electron transfer is accompanied by simultaneous inhibition of the ATPase H+ pumps; and (b) uncoupling of oxidative phosphorylation is not due to depression of delta mu H per se. The consistence of the present data with either a chemiosmotic model where delta mu H is the sole and obligatory intermediate for energy coupling, or models where there is a direct transfer of energy between the two pumps is discussed.  相似文献   

7.
The relationship between molecular mobility (tauR) of the polar spin probe 3-carboxy-proxyl and water content and temperature was established in pea axes by electron paramagnetic resonance (EPR) and saturation transfer EPR. At room temperature, tauR increased during drying from 10(-11) s at 2.0 g water/g dry weight to 10(-4) s in the dry state. At water contents below 0.07 g water/g dry weight, tauR remained constant upon further drying. At the glass transition temperature, tauR was constant at approximately 10(-4) s for all water contents studied. Above Tg, isomobility lines were found that were approximately parallel to the Tg curve. The temperature dependence of tauR at all water contents studied followed Arrhenius behavior, with a break at Tg. Above Tg the activation energy for rotational motion was approximately 25 kJ/mol compared to 10 kJ/mol below Tg. The temperature dependence of tauR could also be described by the WLF equation, using constants deviating considerably from the universal constants. The temperature effect on tauR above Tg was much smaller in pea axes, as found previously for sugar and polymer glasses. Thus, although glasses are present in seeds, the melting of the glass by raising the temperature will cause only a moderate increase in molecular mobility in the cytoplasm as compared to a huge increase in amorphous sugars.  相似文献   

8.
N Sone  A Naqui  C Kumar    B Chance 《The Biochemical journal》1984,221(2):529-533
Reaction of O2 and CO with a caa3-type terminal cytochrome oxidase (EC 1.9.3.1) from the thermophilic bacterium PS3 grown with high aeration was studied at low temperatures. The CO recombination at the temperature range studied (-50 degrees C to -80 degrees C) followed first-order kinetics with an activation energy of 29.3 kJ/mol (7.0 kcal/mol). In the presence of O2 at -113 degrees C the photolysed reduced form binds O2 to form an 'oxy' intermediate similar to Compound A. At a higher temperature (-97 degrees C) another intermediate, similar to Compound B, is formed as a result of electron transfer from the enzyme to the liganded O2.  相似文献   

9.
《Biorheology》1995,32(6):643-653
The effect of temperature on the flow of diluted blood [Hct = 0.21], through 5 μm Nuclepore filters, is described by the Arrhenius equation with an energy of activation of 27.7 kJ/mol. Plasma, diluted with PBS, is almost three times less sensitive to temperature, with an energy of activation of 9.8 kJ/mol, while red cells are of intermediate sensitivity, with an energy of activation of 14.7 kJ/mol. The most sensitive elements to changes in temperature are leukocytes, with energies of activation of 31 kJ/mol and 35 kJ/mol for fast-flowing leukocytes (granulocytes and lymphocytes) and slow-flowing leukocytes (monocytes) respectively. Hence, the major determinants of the decline in filterability of blood through micropore filters are the leukocytes. This effect is compounded when blood is kept for 10 minutes or more at 10° C due to activation of granulocytes, which leads to permanent pore blocking when the affected blood is filtered at room temperature. The combination of increased passage time of leukocytes through peripheral areas at abnormally low temperatures and subsequent activation might influence the flow of blood in non-affected tissues. The effect of temperature on the filterability of red blood cells through 3 μm filters is not described by the Arrhenius equation and the deviations are seen as a gradual change of slope rather than a sharp break between two straight lines. The data are consistent with a gradual shift in rate limiting step away from the entry event into pores, which dominates at low temperature but becomes progressively less important at elevated temperatures. The changing parameter is probably the volume of the red cell, which is less important when flow is measured through 5 μm pores.  相似文献   

10.
Isolated rat and mouse extensor digitorum longus (EDL) and soleus muscles were studied under isometric and isotonic conditions at temperatures from approximately 8 degrees -38 degrees C. The rate constant for the exponential rise of tension during an isometric tetanus had a Q10 of approximately 2.5 for all muscles (corresponding to an enthalpy of activation, delta H = 66 kJ/mol, if the rate was determined by a single chemical reaction). The half-contraction time, contraction time, and maximum rate of rise for tension in an isometric twitch and the maximum shortening velocity in an isotonic contraction all had a similar temperature dependence (i.e., delta H approximately 66 kJ/mol). The Mg++ ATPase rates of myofibrils prepared from rat EDL and soleus muscles had a steeper temperature dependence (delta H = 130 kJ/mol), but absolute rates at 20 degrees C were lower than the rate of rise of tension. This suggests that the Mg++ ATPase cycle rate is not limiting for force generation. A substantial fraction of cross-bridges may exist in a resting state that converts to the force-producing state at a rate faster than required to complete the cycle and repopulate the resting state. The temperature dependence for the rate constant of the exponential decay of tension during an isometric twitch or short tetanus (and the half-fall time of a twitch) had a break point at approximately 20 degrees C, with apparent enthalpy values of delta H = 117 kJ/mol below 20 degrees C and delta H = 70 kJ/mol above 20 degrees C. The break point and the values of delta H at high and low temperatures agree closely with published values for the delta H of the sarcoplasmic reticulum (SR) Ca++ ATPase. Thus, the temperature dependence for the relaxation rate of a twitch or a short tetanus is consistent with that for the reabsorption rate of Ca++ into the SR.  相似文献   

11.
The Arrhenius plots for the membrane-bound ATPase and its soluble form purified from Micrococcus lysodeikticus, presented discontinuities near 30 degrees C at pH 7.5. Glycerol-containing lipids were not responsible for these discontinuities. The values of the enthalpies of activation were 12 (soluble) and 22 (membrane-bound) kcal/mol (50.2 and 92.0 kJ/mol) above 30 degrees C and 42 (soluble) and 29 (membrane-bound) kcal/mol (175.7 and 121.3 kJ/mol) below that temperature. The results suggested that both molecular forms of the ATPase were able to adopt at least two different structures, above and below the critical temperature. Of the two, only the high-temperature structure seemed to be enzymically active. In the case of lipid-dependent ATPases, such as the Escherichia coli enzyme, the transition between both enzyme structures probably occurred with simultaneous "melting" of their lipid microenvironment.  相似文献   

12.
Enthalpy changes (Delta H) of the photointermediates that appear in the photolysis of octopus rhodopsin were measured at physiological temperatures by the laser-induced transient grating method. The enthalpy from the initial state, rhodopsin, to bathorhodopsin, lumirhodopsin, mesorhodopsin, transient acid metarhodopsin, and acid metarhodopsin were 146 +/- 15 kJ/mol, 122 +/- 17 kJ/mol, 38 +/- 8 kJ/mol, 12 +/- 5 kJ/mol, and 12 +/- 5 kJ/mol, respectively. These values, except for lumirhodopsin, are similar to those obtained for the cryogenically trapped intermediate species by direct calorimetric measurements. However, the Delta H of lumirhodopsin at physiological temperatures is quite different from that at low temperature. The reaction volume changes of these processes were determined by the pulsed laser-induced photoacoustic method along with the above Delta H values. Initially, in the transformation between rhodopsin and bathorhodopsin, a large volume expansion of +32 +/- 3 ml/mol was obtained. The volume changes of the subsequent reaction steps were rather small. These results are compared with the structural changes of the chromophore, peptide backbone, and water molecules within the membrane helixes reported previously.  相似文献   

13.
This communication addresses a simple question by means of density functional calculations: Why is iron used as the metal in iron-sulfur clusters? While there may be several answers to this question, it is shown here that one feature - the well-defined inner-sphere reorganization energy of self-exchange electron transfer - is very much favored in iron-sulfur clusters as opposed to metal substituted analogues of Mn, Co, Ni, and Cu. Furthermore, the conclusion holds for both 1Fe and 2Fe type iron-sulfur clusters. The results show that only iron provides a small inner-sphere reorganization energy of 21 kJ/mol in 1Fe (rubredoxin) and 46 kJ/mol in 2Fe (ferredoxin) models, whereas other metal ions exhibit values in the range 57-135 kJ/mol (1Fe) and 94-140 kJ/mol (2Fe). This simple result provides an important, although partial, explanation why iron alone is used in this type of clusters. The results can be explained by simple orbital rules of electron transfer, which state that the occupation of anti-bonding orbitals should not change during the redox reactions. This rule immediately suggests good and poor electron carriers.  相似文献   

14.
Effect of heparin on thrombin inactivation by antithrombin-III.   总被引:4,自引:4,他引:0       下载免费PDF全文
The inactivation of thrombin by heat and by its physiological inhibitor, antithrombin-III, shows quite different dependence on heparin concentration. Heparin at 250 microgram/ml protects thrombin against heat inactivation, and thrombin behaves heterogeneously in this reaction. In the absence of heparin, the thermodynamic activation parameters change with temperature (deltaH+ = 733 kJ/mol and 210 kJ/mol at 50 and 58 degrees C respectively). When heparin is present, heat inactivation of the protected thrombin species proceeds with deltaH+ = 88 kJ/mol and is independent of temperature in the same range. On the other hand, heparin at 0.125-2.5 microgram/ml accelerates the thrombin-antithrombin-III reaction. Thrombin does not show heterogeneity in this reaction and the time courses at any heparin concentration and any temperature between 0 and 37 degrees C appear to follow first-order kinetics. Activation enthalpy is independent of heparin concentration or temperature, deltaH+ = 82-101 kJ/mol, varying slightly with antithrombin-III concentration and thrombin specific activity. Heparin seems to exert its effect by increasing activation entropy. On the basis of these data we suggest a mechanism of action of heparin in the thrombin-antithrombin-III reaction which accounts for all the important features of the latter and seems to unify the different hypotheses that have been advanced.  相似文献   

15.
Arthrospira platensis was cultivated photoautotrophically at 6.0 klux light intensity in 5.0-L open tanks, using a mineral medium containing urea as nitrogen source. Fed-batch experiments were performed at constant flowrate. A central composite factorial design combined to response surface methodology (RSM) was utilized to determine the relationship between the selected response variables (cell concentration after 10 days, X(m), cell productivity, P(X), and nitrogen-to-cell conversion factor, Y(X/N)) and codified values of the independent variables (pH, temperature, T, and urea flowrate, K). By applying the quadratic regression analysis, the equations describing the behaviors of these responses as simultaneous functions of the selected independent variables were determined, and the conditions for X(m) and P(X) optimization were estimated (pH 9.5, T = 29 degrees C, and K = 0.551 mM/day). The experimental data obtained under these conditions (X(m) = 749 mg/L; P(X) = 69.9 mg/L.day) were very close to the estimated ones (X(m) = 721 mg/L; P(X) = 67.1 mg/L.day). Additional cultivations were carried out under the above best conditions of pH control and urea flowrate at variable temperature. Consistently with the results of RSM, the best growth temperature was 29 degrees C. The maximum specific growth rates at different temperatures were used to estimate the thermodynamic parameters of growth (DeltaH* = 59.3 kJ/mol; DeltaS* = -0.147 kJ/mol.K; DeltaG* = 103 kJ/mol) and its thermal inactivation (DeltaH(D) (o) = 72.0 kJ/mol; DeltaS(D) (o) = 0.144 kJ/mol.K; DeltaG(D) (o) = 29.1 kJ/mol).  相似文献   

16.
A double mutant cycle (DMC) approach was employed to estimate the effect of temperature on the contribution of two highly conserved salt bridges to protein stability in the hyperthermophilic protein Ssh10b. The coupling free energy were 2.4 +/- 0.4 kJ/mol at 298 K and 2.2 +/- 0.4 kJ/mol at 353 K for Glu-54/Arg-57, and 6.0 +/- 0.2 kJ/mol at 298 K and 5.9 +/- 0.6 kJ/mol at 353 K for Glu-36/Lys-68. The stability free energy of Ssh10b decrease greatly with increasing temperature, while the direct contribution of these two salt bridges to protein stability remain almost constant, providing evidence supporting the theoretical prediction that salt bridges are extremely resilient to temperature increases and thus are specially suited to improving protein stability at high temperatures. The reason for the difference in coupling free energy between salt bridges Glu-54/Arg-57 and Glu-36/Lys-68 is discussed. Comparing our results with published DMC data for the contribution of salt bridges to stability in other proteins, we found that the energy contribution of a salt bridge formed by two charged residues far apart in the primary sequence is higher than that of those formed between two very close ones. Implications of this finding are useful for engineering proteins with enhanced thermostability.  相似文献   

17.
The temperature- and solvent-induced denaturation of both the SCP2 wild-type and the mutated protein c71s were studied by CD measurements at 222 nm. The temperature-induced transition curves were deconvoluted according to a two-state mechanism resulting in a transition temperature of 70.5 degrees C and 59.9 degrees C for the wild-type and the c71s, respectively, with corresponding values of the van't Hoff enthalpies of 183 and 164 kJ/mol. Stability parameters characterizing the guanidine hydrochloride denaturation curves were also calculated on the basis of a two-state transition. The transitions of the wild-type occurs at 0.82 M GdnHCl and that of the c71s mutant at 0.55 M GdnHCl. These differences in the half denaturation concentration of GdnHCl reflect already the significant stability differences between the two proteins. A quantitative measure are the Gibbs energies DeltaG(0)(D)(buffer) at 25 degrees C of 15.5 kJ/mol for the wild-type and 8.0 kJ/mol for the mutant. We characterized also the alkyl chain binding properties of the two proteins by measuring the interaction parameters for the complex formation with 1-O-Decanyl-beta-D-glucoside using isothermal titration microcalorimetry. The dissociation constants, K(d), for wild-type SCP2 are 335 microM at 25 degrees C and 1.3 mM at 35 degrees C. The corresponding binding enthalpies, DeltaH(b), are -21. 5 kJ/mol at 25 degrees C and 72.2 kJ/mol at 35 degrees C. The parameters for the c71s mutant at 25 degrees C are K(d)=413 microM and DeltaH(b)=16.6 kJ/mol. These results suggest that both SCP2 wild-type and the c71s mutant bind the hydrophobic compound with moderate affinity.  相似文献   

18.
The drying of reduced glutathione from a series of aqueous–ethanol binary solutions at 300 K (below human body temperature) and 330 K (above human body temperature) was investigated in detail by steered molecular simulation and an umbrella sampling method with the Gromacs software package and Gromos96(53a6) united atomic force field. The results show that electrostatic interactions between glutathione and solvent represent the main resistance to drying. When the aqueous solution was gradually changed to pure ethanol, the energy of electrostatic interaction between glutathione and solvent molecules increased by 445.088 kJ/mol, and the drying potential of mean force (PMF) free energy also fell by 253.040 kJ/mol. However, an increase in temperature from 300 to 330 K in the aqueous solution only results in an increase of 23.013 kJ/mol in electrostatic interaction energy and a decrease of 34.956 kJ/mol in drying PMF free energy. Furthermore, we show that hydrogen bonding is the major form of electrostatic interaction involved, and directly affects the drying of glutathione. Therefore, choosing water-miscible solvents that minimise hydrogen-bond formation with glutathione will enhance its drying rate, and this is likely to be more efficient than increasing the temperature of the process. Thus, a power-saving technology can be used to produce the high bioactivity medicines.  相似文献   

19.
The cellular autolytic reaction system in Streptococcus faecalis ATCC 9790 was analyzed for relative increases in reaction rates with increasing temperature by determination of Arrhenius activation energies (E). The systems examined were: (i) an isolated wall-enzyme complex in 0.01 M sodium phosphate, pH 6.9; (ii) exponential-phase cells suspended in 0.01 or o.3 M sodium phosphate pH 6.8, or in 0.04 M ammonium acetate, pH 6.8, (iii) growing cultures deprived of glucose or lysine; and (iv) cultures treated in growth media with the nonionic detergent, Triton X-100. For detergent-treated cells, E values were between 23.9 and 27.4 kcal/mol (ca. 100.1 to 174.7 kJ/mol) at concentrations of Triton X-100 between about 0.03 and 0.072 mg/ml. E values dropped sharply to 11.5 to 13.0 kcal/m-l (ca. 48.2 to 54.4 kJ/mol) at Triton X-100 concentrations of 0.12 mg/ml or higher. For the remaining systems, E values ranged from 16 to 20 kcal/mol (ca. 67.0 to 83.7 kJ/mol) (wall lysis, cellular autolysis in 0.01 M sodium phosphate or in 0.04 M ammonium acetate, and autolysis of glucose-starved cells) to 31 to 38 kcal/mol (ca 129.8 to 159.1 kJ/mol) (cellular autolysis in 0.3 M sodium phosphate or autolysis of lysine-starved cells). High concentrations of Triton X-100 appear to lower the E values below the 16 to 20 kcal/mol observed for the autolysis of isolated walls. This effect may be related to disruption by the detergent of a hydrophobic complex regulating cellular autolysis in vivo.  相似文献   

20.
The rate constants for internalization of surface-bound asialo-orosomucoid by hepatocytes were 0.040 min-1 at 20 degrees C, 0.18 min-1 at 30 degrees C and 0.28 min-1 at 40 degrees C. At 40 degrees C, internalization accounted for most of the increase in cell-associated radioactivity. The activation energy over the temperature range 20 to 40 degrees C was 68 +/- 7 (S.D.) kJ/mol. At 10 degrees C, most of the cell-associated asialo-orosomucoid was bound to the cell surface in a reaction which followed ordinary chemical kinetics. Pre-incubation of hepatocytes with a large concentration of unlabelled asialo-orosomucoid did not influence the uptake of subsequently added 125I-asialofetuin; neither was degradation of 125I-asialo-fetuin affected in this experiment. The fractional rate of degradation (the fraction of cell-associated asialo-fetuin which was degraded per unit time) was constant over a twelve-fold range of intracellular asialo-fetuin concentrations. Increasing the temperature from 20 to 30 degrees C produced approximately a ten-fold increase in the rate of degradation of either asialo-fetuin or asialo-orosomucoid. The average activation energies of degradation over the range 20 to 40 degrees C were 125 kJ/mol for asialo-fetuin and 149 kJ/mol for asialo-orosomucoid; however, the Arrhenius plots were not straight lines over this temperature range.  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