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 93 毫秒
1.
The action of near-ultraviolet (UV-365 nm) radiation in cellular inactivation (biological measurements) and induction and repair of DNA strand breaks (physical measurements) were studied in a repair-proficient strain and in polA-, recA-, uvrA-, and polA uvrA-deficient strains of Escherichia coli K-12. The induction of breaks in the polA and polA uvrA strains was linear with dose (4.0 and 3.7 X 10(-5) breaks/2.5 X 10(9) daltons/Jm-2, respectively). However, in the recA-, uvrA-, and repair-proficient strains, there was an initial lag in break induction at low doses and then a linear induction of breaks at higher doses with rates of 4.6, 2.8, and 3.2 X 10(-5) breaks/2.5 X 10(9) daltons/Jm-2, respectively. We interpret these strain differences as indicating simultaneous induction and repair of breaks in polymerase 1 (polA)-proficient strains under the 0 degrees C, M9 buffer irradiation conditions that, for maximum efficiency, require both the polA and recA gene products. Strand-break rejoining also occurred at 30 degrees C in complete growth medium. We propose that at least three (and possibly four) distinct types of pathways can act to reduce the levels of 365-nm radiation-induced strand breaks. A quantitative comparison of the number of breaks remaining with the number of lethal events remaining after repair in complete medium at 30 degrees C showed that between one and three breaks remain per lethal event in the wild-type and recA strains, whereas in the polA strain one order of magnitude more breaks were induced.  相似文献   

2.
The survival of the yeast mutant rad 54-3, which is temperature-conditional for the repair of double-strand breaks, was measured after exposure to UV-light (254 nm) and incubation at 23 degrees C and 36 degrees C. It was found that survival was drastically reduced with incubation at the restrictive temperature. Temperature-shift experiments indicated that repair of UV-induced damage which is controlled by the rad 54 gene proceeds with a half-value-time of about 7 h.  相似文献   

3.
Arrhenius plots of various enzyme and transport systems associated with the liver mitochondrial inner membranes of ground squirrels exhibit changes in slope at temperatures of 20-25 degrees C in nonhibernating but not in hibernating animals. It has been proposed that the Arrhenius breaks observed in nonhibernating animals are the result of a gel to liquid-crystalline phase transition of the mitochondrial membrane lipids, which also occurs at 20-25 degrees C, and that the absence of such breaks in hibernating animals is due to a major depression of this lipid phase transition to temperatures below 4 degrees C. In order to test this hypothesis, we have examined the thermotropic phase behavior of liver inner mitochondrial membranes from hibernating and nonhibernating Richardson's ground squirrels, Spermophilus richardsonii, by differential scanning calorimetry and by 19F nuclear magnetic resonance and fluorescence polarization spectroscopy. Each of these techniques indicates that no lipid phase transition occurs in the membranes of either hibernating or nonhibernating ground squirrels within the physiological temperature range of this animal (4-37 degrees C). Moreover, differential scanning calorimetric measurements indicate that only a small depression of the lipid gel to liquid-crystalline phase transition, which is centered at about -5 degrees C in nonhibernating animals and at about -9 degrees C in hibernators, occurs. We thus conclude that the Arrhenius plot breaks observed in some membrane-associated enzymatic and transport activities of nonhibernating animals are not the result of a lipid phase transition and that a major shift in the gel to liquid-crystalline lipid phase transition temperature is not responsible for seasonal changes in the thermal behavior of these inner mitochondrial membrane proteins.  相似文献   

4.
Freezing of the enteropathogenic bacterium Yersinia enterocolitica to -18 and -75 degrees C caused 7 and 42% cell death, respectively, and 0.329 and 0.588 single-strand breaks per 10(8) daltons of DNA, respectively, while radiation to one D10 dose (10% cell survival) combined with freezing to 2 to 0, -18, and -75 degrees C induced 0.05, 0.75, and 5.04 single-strand breaks, respectively. The increase in the effectiveness of radiation with respect to the yield of single-strand breaks at -18 and -75 degrees C is contrary to expectation and seems to be due to arrest of repair of single-strand breaks by these low temperatures.  相似文献   

5.
The kinetics of transfer of two electrons from a photodonor (a system containing eosin and NADH or 4;,5;-dibromofluorescein and NADH) to Fe-protein (Av2) and the kinetics of transfer of the first and second electrons from Av2 to Mo-Fe-protein (Av1) were studied by kinetic laser spectroscopy of nitrogenase from Azotobacter vinelandii. The effects of the substrates of nitrogenase (nitrogen, acetylene, and protons) on the intramolecular electron transfer in nitrogenase were studied. Analysis of the effect of photodonor excitation radiation intensity on the rate of electron transfer was used to determine the transfer rate constants for the first (k1) and second (k2) electrons from Av2 to Av1. In the presence of MgATP, two electrons are sequentially transferred from Av2 to Av1, and no delay between these reactions was detected. The first electron transferred from Av2 to Av1 is not targeted to the substrate; k1 = 154 +/- 15 sec-1 at 23 degrees C for the system 4;,5;-dibromofluorescein-NADH; k2 = 53 +/- 5 sec-1, 95 +/- 9 sec-1, and 24 +/- 2 sec-1 at 23 degrees C in the presence of nitrogen, acetylene, and argon, respectively. An unidentified slow step (k3 = 18 +/- 2 sec-1 at 23 degrees C) may be associated with electron transfer within Av1.  相似文献   

6.
The hypothesis that after irradiation a competition exists between fixation of radiation damage and its repair and that this competition determines cell survival was to be tested. Postirradiation temperature of holding was employed as a means of modulating rate of damage repair, and the postirradiation rates of repair of DNA strand breaks (both single and double) were monitored using elution assays. At temperatures below 37 degrees C following irradiation the rates of rejoining were decreased markedly, although rejoining of single-strand breaks was seen even at 10 degrees C and rejoining of double-strand breaks still occurred at 16 degrees C. However, 3 h incubation of cells at these lowered temperatures had no observable effect on cell survival parameters. It is concluded that either damage fixation and damage repair have the same dependence on temperature, or simple measurements of rejoining of breaks are insufficient to detect the details of the competition between repair and fixation (some measure of fidelity of repair is needed).  相似文献   

7.
1. The superoxide anion radical (O2-) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2- and ferrocytochrome c. 2. At 20 degrees C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4-10(6) M-1. S -1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2- and the form of cytochrome c which exists above pH approximately 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2- reacts with the form present below pH 7.45 with k = 1.4-10(6) M-1 - S-1, the form above pH 7.45 with k = 3.0- 10(5) M-1 - S-1, and the form present above pH 9.2 with k = 0. 3. The reaction has an activation energy of 20 kJ mol-1 and an enthalpy of activation at 25 degrees C of 18 kJ mol-1 both above and below pH 7.45. It is suggested that O2- may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2-6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5-10(5)-5-10(6) M-1 - S-1.  相似文献   

8.
The kinetics and energetics of the binding between barley alpha-amylase/subtilisin inhibitor (BASI) or BASI mutants and barley alpha-amylase 2 (AMY2) were determined using surface plasmon resonance and isothermal titration calorimetry (ITC). Binding kinetics were in accordance with a 1:1 binding model. At pH 5.5, [Ca(2+)] = 5 mM, and 25 degrees C, the k(on) and k(off) values were 8.3 x 10(+4) M(-1) s(-1) and 26.0 x 10(-4) s(-1), respectively, corresponding to a K(D) of 31 nM. K(D) was dependent on pH, and while k(off) decreased 16-fold upon increasing pH from 5.5 to 8.0, k(on) was barely affected. The crystal structure of AMY2-BASI shows a fully hydrated Ca(2+) at the protein interface, and at pH 6.5 increase of [Ca(2+)] in the 2 microM to 5 mM range raised the affinity 30-fold mainly due to reduced k(off). The K(D) was weakly temperature-dependent in the interval from 5 to 35 degrees C as k(on) and k(off) were only increasing 4- and 12-fold, respectively. A small salt dependence of k(on) and k(off) suggested a minor role for global electrostatic forces in the binding and dissociation steps. Substitution of a positively charged side chain in the mutant K140L within the AMY2 inhibitory site of BASI accordingly did not change k(on), whereas k(off) increased 13-fold. ITC showed that the formation of the AMY2-BASI complex is characterized by a large exothermic heat (Delta H = -69 +/- 7 kJ mol(-1)), a K(D) of 25 nM (27 degrees C, pH 5.5), and an unfavorable change in entropy (-T Delta S = 26 +/- 7 kJ mol(-1)). Calculations based on the thermodynamic data indicated minimal structural changes during complex formation.  相似文献   

9.
Mukaiyama A  Takano K  Haruki M  Morikawa M  Kanaya S 《Biochemistry》2004,43(43):13859-13866
Equilibrium and kinetic studies were carried out under denaturation conditions to clarify the energetic features of the high stability of a monomeric protein, ribonuclease HII, from a hyperthermophile, Thermococcus kodakaraensis (Tk-RNase HII). Guanidine hydrochloride (GdnHCl)-induced unfolding and refolding were measured with circular dichroism at 220 nm, and heat-induced denaturation was studied with differential scanning calorimetry. Both GdnHCl- and heat-induced denaturation are very reversible. It was difficult to obtain the equilibrated unfolding curve of Tk-RNase HII below 40 degrees C, because of the remarkably slow unfolding. The two-state unfolding and refolding reactions attained equilibrium at 50 degrees C after 2 weeks. The Gibbs energy change of GdnHCl-induced unfolding (DeltaG(H(2)O)) at 50 degrees C was 43.6 kJ mol(-1). The denaturation temperature in the DSC measurement shifted as a function of the scan rate; the denaturation temperature at a scan rate of 90 degrees C h(-1) was higher than at a scan rate of 5 degrees C h(-1). The unfolding and refolding kinetics of Tk-RNase HII were approximated as a first-order reaction. The ln k(u) and ln k(r) values depended linearly on the denaturant concentration between 10 and 50 degrees C. The DeltaG(H(2)O) value obtained from the rate constant in water using the two-state model at 50 degrees C, 44.5 kJ mol(-1), was coincident with that from the equilibrium study, 43.6 kJ mol(-1), suggesting the two-state folding of Tk-RNase HII. The values for the rate constant in water of the unfolding for Tk-RNase HII were much smaller than those of E. coli RNase HI and Thermus thermophilus RNase HI, which has a denaturation temperature similar to that of Tk-RNase HII. In contrast, little difference was observed in the refolding rates among these proteins. These results indicate that the stabilization mechanism of monomeric protein from a hyperthermophile, Tk-RNase HII, with reversible two-state folding is characterized by remarkably slow unfolding.  相似文献   

10.
Temperature dependence of photoinhibition and photoprotective mechanisms (10-35 degrees C) was investigated for Chenopodium album leaves grown at 25 degrees C under 500 micro mol quanta m(-2) s(-1). The fraction of active photosystem II (PSII) was determined after photoinhibitory treatment at different temperatures in the presence and absence of lincomycin, an inhibitor of chloroplast-encoded protein synthesis. In the absence of lincomycin, leaves were more tolerant to photoinhibition at high (25-35 degrees C) than at low (11-15 degrees C) temperatures. In the presence of lincomycin, the variation in the tolerance to photoinactivation became relatively small. The rate constant of photoinactivation (k(pi)) was stable at 25-35 degrees C and increased by 50% with temperature decrease from 25 to 11 degrees C. The rate constant of recovery of inactivated PSII (k(rec)) was more sensitive to temperature; it was very low at 11 degrees C and increased by an order of magnitude at 35 degrees C. We conclude that the recovery of photoinactivated PSII plays an essential role in photoprotection at 11-35 degrees C. Partitioning of light energy to various photoprotective mechanisms was further analyzed to reveal the factor responsible for k(pi). The fraction of energy utilized in photochemistry was lower at lower temperatures. Although the fraction of heat dissipation increased with decreasing temperatures, the excess energy that is neither utilized by photochemistry nor dissipated by heat dissipation was found to be greater at lower temperatures. The k(pi) value was strongly correlated with the excess energy, suggesting that the excess energy determines the rate of photoinactivation.  相似文献   

11.
Rate constants for the hydrolysis of L-proline methyl ester to form proline and methanol in D(2)O buffered at neutral pD and 25 degrees C and the deuterium enrichment of the proline product determined by electrospray ionization mass spectrometry are reported. The data give k(DO) = 5.3 +/- 0.5 M(-1) s(-1) as the second-order rate constant for carbon deprotonation of N-protonated proline methyl ester by deuterioxide ion in D(2)O at 25 degrees C and I = 1.0 (KCl). The data provide good estimates of carbon acidities of pK(a) = 21 for N-protonated proline methyl ester and pK(a) = 29 for proline zwitterion in water and of the second-order rate constant k(HO) = 4.5 x 10(-5) M(-1) s(-1) for carbon deprotonation of proline zwitterion by hydroxide ion at 25 degrees C. There is no detectable acceleration of the deprotonation of N-protonated proline methyl ester by the Br?nsted base 3-quinuclidinone in water, and it is not clear that such Br?nsted catalysis would make a significant contribution to the rate acceleration for deprotonation of bound proline at proline racemase. A comparison of the first-order rate constants k(HO)[HO(-)] = 4.5 x 10(-11) s(-1) for deprotonation of free proline zwitterion in water at pH 8 and k(cat) = 2600 s(-1) for deprotonation of proline bound to the active site of proline racemase at pH 8 shows that the enzymatic rate acceleration for proline racemase is ca. 10(13)-fold. This corresponds to a 19 kcal/mol stabilization of the transition state for deprotonation of the enzyme-bound carbon acid substrate by interaction with the protein catalyst. It is suggested that (1) much of the rate acceleration of the enzymatic over the nonenzymatic reaction in water may result from transfer of the substrate proline zwitterion from the polar solvent water to a nonpolar enzyme active site and (2) the use of thiol anions rather than oxygen anions as Br?nsted bases at this putative nonpolar enzyme active site may be favored, because of the smaller energetic price for desolvation of thiol anions than for desolvation of the more strongly solvated oxygen anions.  相似文献   

12.
Human adenovirus type 2 (Ad 2) was irradiated with 137Cs gamma rays in the liquid state at 0 degree C. DNA breaks were correlated with the inactivation of several viral functions and compared to results obtained previously for irradiation of Ad 2 under frozen conditions at -75 degrees C. Irradiation at 0 degree C induced 170 +/- 20 single-strand breaks and 2.6 +/- 0.4 double-strand breaks/Gy/10(12) Da in the viral DNA. Viral adsorption to human KB cells was inactivated with a D0 of 9.72 +/- 1.18 kGy, whereas the inactivation of Ad 2 plaque formation had a D0 of 0.99 +/- 0.14 or 1.1 +/- 0.29 kGy when corrected for the effect of radiation on virus adsorption. For the adsorbed virus, an average of 4.3 +/- 1.7 single-strand and 0.065 +/- 0.02 double-strand breaks were induced in the viral DNA per lethal hit. In contrast, irradiation of Ad 2 at -75 degrees C results in 2.6- to 3.4-fold less DNA breakage per Gy and a 5.6-fold increase in D0 for plaque formation of the adsorbed virus. Furthermore, although host cell reactivation (HCR) of Ad 2 viral structural antigen production for irradiated virus was substantially reduced in the xeroderma pigmentosum fibroblast strain (XP25RO) compared to normal strains for irradiation at -75 degrees C (57% HCR), it was only slightly reduced compared to normal for irradiation at 0 degree C (88% HCR). These results indicate that the spectrum of DNA damage is both quantitatively and qualitatively different for the two conditions of irradiation.  相似文献   

13.
When a mutation in an essential gene shows a temperature-sensitive phenotype, one usually assumes that the protein is inactive at nonpermissive temperature. DNA gyrase is an essential bacterial enzyme composed of two subunits, GyrA and GyrB. The gyrB652 mutation results from a single base change that substitutes a serine residue for arginine 436 (R436-S) in the GyrB protein. At 42 degrees C, strains with the gyrB652 allele stop DNA replication, and at 37 degrees C, such strains grow but have RecA-dependent SOS induction and show constitutive RecBCD-dependent DNA degradation. Surprisingly, the GyrB652 protein is not inactive at 42 degrees C in vivo or in vitro and it doesn't directly produce breaks in chromosomal DNA. Rather, this mutant has a low k(cat) compared to wild-type GyrB subunit. With more than twice the normal mean number of supercoil domains, this gyrase hypomorph is prone to fork collapse and topological chaos near the terminus of DNA replication.  相似文献   

14.
E P Lennette  B V Plapp 《Biochemistry》1979,18(18):3933-3938
The reaction of the imidazole group of histidine hydantoin with bromoacetate was studied as a model for carboxymethylation of histidine residues in proteins. pK values of 6.4 and 9.1 (25 degrees C) and apparent heats of ionization of 7.8 and 8.7 kcal/mol were determined for the imidazole and hydantoin rings, respectively. At pH values corresponding to the isoelectric points for histidine hydantoin, the rates of carboxymethylation at 12, 25, 37, and 50 degrees C were determined; the modified hydantoins were hydrolyzed to the corresponding histidine derivatives for quantitative amino acid analysis. At pH 7.72 and 25 degrees C, the imidazole tele-N was alkylated (k = 3.9 X 10(-5) M-1 s-1) twice as fast as the pros-N. The monocarboxymethyl derivatives were carboxymethylated at the same rate at the pros-N (k = 2.1 X 10(-5) M-1 s-1) but 3 times faster at the tele-N (k = 11 X 10(-5) M-1 s-1). The enthalpies of activation determined for carboxymethylation of the imidazole ring and its monocarboxymethyl derivatives were similar (15.9 +/- 0.7 kcal/mol). delta S for the four carboxymethylations was -25 +/- 2 eu. The electrostatic component of delta S (delta S es) was calculated from the influence of the dielectric constant on the reaction rate at 25 degrees C. delta S es was slightly negative (-4 +/- 1 eu) for mono- or dicarboxymethylations, indicating some charge separation in the transition state. The nonelectrostatic entropy of activation was -21 +/- 2 eu for all four carboxymethylations.  相似文献   

15.
We have made experimental studies into the enzymatic hydrolysis of cellobiose within the temperature range of 40 degrees C to 70 degrees C at pH 4.9, by using beta-1,4-glucosidase from Aspergillus niger. At 70 degrees C there was significant enzyme deactivation, which could be fitted to a potential deactivation model with values of n equal to 1.09 and k(d) to 0.1564 (g/l)(-0.09) min(-1), whereas the rate of hydrolysis could be fitted to the Michaelis-Menten equation. Between 40 degrees C and 60 degrees C we noted a substrate inhibition and that the CEC compound formed contributed to glucose production. The apparent activation energies had values of 4.66, 8.45, 4.82, and 3.99 kJ/mol for the kinetic constants k(a) and k(a2) the Michaelis constant and the substrate inhibition constant, respectively.  相似文献   

16.
Yeast cells deficient in DNA ligase were also deficient in their capacity to rejoin single-strand scissions in prelabeled nuclear DNA. After high-dose-rate gamma irradiation (10 and 25 krads), cdc9-9 mutant cells failed to rejoin single-strand scissions at the restrictive temperature of 37 degrees C. In contrast, parental (CDC9) cells (incubated with mutant cells both during and after irradiation) exhibited rapid medium-independent DNA rejoining after 10 min of post-irradiation incubation and slower rates of rejoining after longer incubation. Parental cells were also more resistant than mutant cells to killing by gamma irradiation. Approximately 2.5 +/- 0.07 and 5.7 +/- 0.6 single-strand breaks per 10(8) daltons were detected in DNAs from either CDC9 or cdc9-9 cells converted to spheroplasts immediately after 10 and 25 krads of irradiation, respectively. At the permissive temperature of 23 degrees C, the cdc9-9 cells contained 2 to 3 times the number of DNA single-strand breaks as parental cells after 10 min to 4 h of incubation after 10 krads of irradiation, and two- to eightfold more breaks after 10 min to 2.5 h of incubation after 25 krads of irradiation. Rejoining of single-strand scissions was faster in medium. After only 10 min in buffered growth medium and after 10 krads of irradiation, the number of DNA single-strand breaks was reduced to 0.32 +/- 0.3 (at 23 degrees C) or 0.21 +/- 0.05 (at 37 degrees C) per 10(8) daltons in parental cells, but remained at 2.1 +/- 0.06 (at 23 degrees C) or 2.3 +/- 0.07 (at 37 degrees C) per 10(8) daltons in mutant cells. After 10 or 25 krads of irradiation plus 1 h of incubation in medium at 37 degrees C, only DNA from CDC9 cells was rejoined to the size of DNA from unirradiated cells, whereas at 23 degrees C, DNAs in both strains were completely rejoined.  相似文献   

17.
Using a high-sensitivity differential scanning microcalorimeter capable of performing cooling scans, we have examined the phase behavior of small unilamellar vesicles (SUV) as a function of time of storage above their order-disorder phase transition. Vesicles composed of dipalmitoylphosphatidylcholine (DPPC) and dimyristoylphosphatidylcholine (DMPC) were examined. Cooling scans on fresh (5-7-h postsonication) samples revealed broad, relatively simple heat capacity peaks (peak temperatures: 19.9 degrees C for DMPC, 37.8 degrees C for DPPC) free of high-temperature spikes or shoulders. Subsequent heating scans displayed a sharp peak characteristic of previously described fusion products formed below the phase transition. SUV samples stored for 1 or more days above their phase transition displayed a moderately broad, high-temperature shoulder (23.8 degrees C for DMPC and 40.2 degrees C for DPPC) in the cooling profile. For DMPC, the enthalpy associated with this peak increased in a first-order fashion with time. Hydrolysis products were not detected until 12-20 days of storage. Both the rate and extent of shoulder appearance increased with temperature (k = 0.0017 h-1, fraction of total enthalpy = 0.1 at 36 degrees C; k = 0.0037 h-1, fraction = 0.2 at 42 degrees C). Freeze-fracture electron micrographs confirmed that an intermediate-sized vesicle population (diameters 400-500 A) appeared in SUV samples stored above their phase transition. Also, the trapped volume of DMPC SUV increased from 0.26 microL/mumol after 17 h of storage to 0.54 microL/mumol after storage for 16 days at 36 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Cytosolic and nuclear estrogen receptor forms of chicken oviduct have been studied by (1) measuring hormone dissociation kinetics and by (2) sucrose density gradient analysis on high salt gradients. Estradiol dissociates from the receptor in chicken oviduct cytosol at 22 degrees C following a two-phase exponential process. The fraction of receptor with a fast dissociation rate (k = 120 X 10(-3) min-1) decreases as a function of the pre-incubation at 22 degrees C; after prolonged pre-incubation only the slowly dissociating (k = 12.3 X 10(-3) min-1) form remains. Dissociation of moxestrol, a synthetic estrogen with a higher affinity, from the cytosol receptor at 30 degrees C is similar, showing a transition of a fast dissociating form (k = 120 X 10(-3) min-1) to a slowly dissociating form (k = 7.6 X 10(-3) min-1) as a result of pre-incubation at 30 degrees C. A concomitant temperature-dependent shift of the estrogen receptor from a 4.8 S to a 6.1 S form was observed with moxestrol but not with estradiol as a ligand. Sodium molybdate (20 mM) and NaSCN (400 mM) inhibit the temperature-dependent increase in sedimentation coefficient, but molybdate allows the formation of a receptor form which shows intermediary dissociation kinetics. Estrogen receptor, precipitated with ammonium sulfate (0-35%) shows monophasic dissociation kinetics of estradiol (k = 39.5 X 10(-3) min-1) and for moxestrol (k = 10.8 X 10(-3) min-1), suggesting full receptor activation only with moxestrol as a ligand. Moxestrol-receptor complexes obtained by ammonium sulfate precipitation sediment at 0 degree C at 4.8 S. Only after subsequent incubation at 30 degrees C a shift from 4.8 S to 5.9 S is observed, suggesting that the formation of the slowly dissociating form of the receptor may precede the formation of a stable transformed receptor complex. The nuclear estrogen receptor with estradiol as a ligand shows biphasic dissociation kinetics at 22 degrees C (k = 70 X 10(-3) min-1; k = 14.0 X 10(-3) min-1). The ratio of both components (1:1) does not change after preincubation of the nuclear receptor extract at 22 degrees C. Moxestrol dissociates from the nuclear receptor at 30 degrees C monophasically with a slow rate (k = 6.1 X 10(-3) min-1), suggesting that it is extracted as an activated hormone-receptor complex.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Heptafluorostearic acid, an isogeometric derivative of stearic acid, has a pK(a) value of about 0.5. To evaluate the suitability of heptafluorostearate as model compound for anions of long-chain fatty acids in membrane transport, monolayer and liposome studies were performed with lipid mixtures containing phospholipids;-cholesterol-heptafluorostearate or stearate (100:40:20 molar ratios). Transfer of heptafluorostearate and stearate from liposomes to bovine serum albumin (BSA) was followed by measuring the intrinsic fluorescence of BSA. The percentage of heptafluorostearate, equivalent to the amount placed in their outer monolayer, transferred from liposomes (120;-130 nm diameter) to BSA was 55.7 +/- 3.7% within 10 min at 25 degrees C and 55 +/- 2% within 5 min at 37 degrees C. Slow transfer of 22.7 +/- 2.5% of heptafluorostearate at 25 degrees C followed with a half-life of 2.3 +/- 0.4 h and of 20 +/- 4% at 37 degrees C with a half-life of 0.9 +/- 0.1 h until the final equilibrium distributions between BSA and liposomes were reached, 79 +/- 6% to 21 +/- 5% at 25 degrees C and 75 +/- 5% to 25 +/- 4% at 37 degrees C. The pseudounimolecular rate constants for flip-flop of heptafluorostearate equal k(FF,25) = 0.24 +/- 0.05 h(-) and k(FF,37) = 0.6 +/- 0.1 h(-), respectively. By comparison, transfer of stearate required only 3 min to reach equilibrium distribution.The difference between heptafluorostearate and stearate may be explained by a rapid flip-flop movement of the un-ionized fatty acids which exist in different concentrations in accordance with their pK(a) values. Half-life of flip-flop of heptafluorostearate makes it suitable to study mediated membrane transport of long-chain fatty acid anions.  相似文献   

20.
Activated bovine plasma protein C (APC) was not reactive with the substrate p-nitrophenyl p-guanidinobenzoate (NPGB) in the absence of cations. In the presence of increasing concentrations of Na+, the acylation rate constant, k2,app, at 7 degrees C, progressively increased from 0.32 +/- 0.03 s-1 at 12.5 mM Na+ to 1.15 +/- 0.10 s-1 at 62.5 mM Na+. A linear dependence of the reciprocal of k2,app with [Na+]-2 was observed, indicating that at least two monovalent cation sites, or classes of sites, are necessary for the catalytic event to occur. From this latter plot, the k2,max for APC catalysis of NPGB hydrolysis, at saturating [Na+] and [NPGB], was calculated to be 1.21 +/- 0.10 s-1, and the Km for Na+ was found to be 21 +/- 3 mM. The dissociation constant, Ks, for NPGB to APC, at 7 degrees C, was not altered as [Na+] was increased, yielding a range of values of 18.5 X 10(-5) to 19.9 X 10(-5) M as [Na+] was varied from 12.5 to 62.5 mM. The deacylation rate constant, k3, for p-guanidinobenzoyl-APC hydrolysis was also independent of [Na+], with a value of (3.8 +/- 1.0) X 10(-3) s-1 in the absence of Na+ or in the presence of concentrations of Na+ up to 200 mM. Identical kinetic behavior was observed when Cs+ was substituted for Na+ in the above enzymic reaction. The pre-steady-state kinetic parameters were calculated according to the same methodology as described above.(ABSTRACT TRUNCATED AT 250 WORDS)  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