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.
In the present study, we defined experimental conditions that allowed the extraction of the integral membrane protein lysophospholipid:acyl-CoA acyltransferase (LAT, EC 2.3.1.23) from membranes while maintaining the full enzyme activity using the nonionic detergent n-octyl glucopyranoside (OGP) and solutions of high ionic strength. We found that the optimal OGP concentration depended on the ionic strength of the solubilization buffer. Fluorescence measurements with 1,6-diphenyl-1,3,5-hexatriene indicated that the critical micellar concentration (CMC) of OGP decreased with increasing salt concentrations. Analogous studies revealed that the zwitterionic detergent Chaps was ineffective in extracting LAT from membranes in the absence of salt, whereas its solubilization efficiency increased with increasing salt concentrations. Detailed lipid analysis of the different protein/lipid/detergent mixed micelles showed that the protein/lipid/OGP mixed micelles were relatively enriched with sphingomyelin (SPM) compared to protein/lipid/Chaps mixed micelles, indicating that the differences in the solubilization efficiency may be due to the ability to extract more SPM from membranes. When the protein/lipid/OGP mixed micelles were dissociated into protein/detergent and lipid/detergent complexes by the addition of increasing Chaps concentrations, one-tenth of the LAT enzyme activity was preserved making the enzyme accessible to protein purification. Analysis by native PAGE revealed that in the presence of excess Chaps a high molecular mass protein complex migrated into the gel which could be photolabeled by 125I-labelled-18-(4'-azido-2'-hydroxybenzoylamino)-oleyl-CoA. This fatty acid analogue has been shown to be a competitive inhibitor of LAT enzyme activity in the dark, and an irreversible inhibitor after photolysis. Therefore, this protein complex is assumed to contain the LAT enzyme.  相似文献   

2.
Molecular properties of cytosolic Ah receptors from livers of Sprague-Dawley rats and C57BL/6N mice were assessed by velocity sedimentation on sucrose gradients and by gel permeation chromatography on Sephacryl S-300. Analyses were done under conditions of both moderate ionic strength (presence of 0.1 M KCl) and high ionic strength (0.4 M KCl). [3H] 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) was used as the radioligand. In conditions of moderate ionic strength the receptor from Sprague-Dawley rat liver sedimented at 8.8 +/- 0.05 S, had a Stokes radius of 7.0 +/- 0.21 nm, and an apparent relative molecular mass (Mr) of 257,000 +/- 7,700. In conditions of high ionic strength the Ah receptor from rat hepatic cytosol dissociated to a [3H]TCDD-binding subunit which sedimented at 5.6 +/- 0.58 S, had a Stokes radius of 5.2 +/- 0.24 nm, and an apparent Mr of 121,000 +/- 5,600. The Ah receptor from liver of C57BL/6N mice, in moderate ionic strength conditions, sedimented at 9.4 +/- 0.54 S, had a Stokes radius of 7.1 +/- 0.12 nm, and an apparent Mr of 277,000 +/- 4,800. Whereas the Ah receptor from rat liver readily dissociated into a [3H]TCDD-binding subunit during brief exposure to 0.4 M KCl, the mouse Ah receptor resisted dissociation. When exposed to 0.4 M KCl for 2 h, the mouse Ah receptor remained at the same molecular size that it had exhibited in moderate ionic strength conditions. Prolonged exposure (16 h) to 0.4 M KCl prior to analysis partially converted the mouse Ah receptor into a smaller [3H]TCDD-binding subunit which sedimented at 4.9 +/- 0.07 S, had a Stokes radius of 5.2 +/- 0.19 nm, and an apparent Mr of 105,000 +/- 3,800. The potency of seven different Ah receptor agonists in competing with [3H]TCDD for specific receptor sites was slightly different in mouse cytosol than in rat cytosol. By criteria of size, response to high ionic strength environments, and ligand binding preferences the mouse and rat Ah receptors appear to be similar but not identical molecular species.  相似文献   

3.
Molecular properties of nuclear aromatic hydrocarbon (Ah) receptor from Hepa-1c1c9 (Hepa-1) cells were assessed by velocity sedimentation on sucrose gradients and by gel permeation chromatography on Sephacryl S-300. Nuclear Ah receptor was obtained by exposing intact cells to [3H]-2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) for 1 h at 37 degrees C in culture followed by extraction of receptor from nuclei with buffers containing 0.5 M KCl. The nuclear Ah receptor was compared to the cytosolic Ah receptor from the same cells. Under conditions of low ionic strength, the Ah receptor from Hepa-1 cytosol sedimented as a single 9.4 +/- 0.63 S binding peak that had a Stokes radius of 7.1 +/- 0.12 nm and an apparent relative molecular mass of 271,000 +/- 16,000. After prolonged (24 h) exposure to high ionic strength (0.5 M KCl), cytosol labeled with [3H]TCDD exhibited two specific binding peaks. The large form of cytosolic Ah receptor seen under high ionic strength conditions sedimented at 9.4 +/- 0.46 S, had a Stokes radius of 6.9 +/- 0.19 nm, and an apparent Mr 267,000 +/- 15,000. The smaller ligand-binding subunit generated by exposing cytosol to 0.5 M KCl sedimented at 4.9 +/- 0.62 S, had a Stokes radius of 5.0 +/- 0.14 nm, and an apparent Mr 104,000 +/- 12,000. Nuclear Ah receptor, analyzed under high ionic strength conditions, sedimented at 6.2 +/- 0.20 S, had a Stokes radius of 6.8 +/- 0.19 nm, and an apparent Mr 176,000 +/- 7000. Nuclear Ah receptor from rat H4IIE hepatoma cells was analyzed and found to have physicochemical characteristics identical to those of nuclear Ah receptor from the mouse Hepa-1 cells. The molecular mass of Hepa-1 nuclear Ah receptor was found to be statistically different from both the Mr approximately 267,000 cytosolic Ah receptor and the Mr approximately 104,000 subunit which were present in cytosol under high ionic strength conditions. Hepa-1 nuclear Ah receptor could not be converted to a smaller ligand-binding subunit by treatment with alkaline phosphatase, ribonuclease, or sulfhydryl-modifying reagents or prolonged exposure to 1.0 M KCl. Cytosolic Ah receptor from Hepa-1 cells was "transformed" by heating at 25 degrees C in vitro into a form with high affinity for DNA-cellulose. The transformed cytosolic Ah receptor, when analyzed under conditions of high ionic strength, sedimented at approximately 6 S, had a Stokes radius of approximately 6.7 nm, and an apparent Mr approximately 167,000.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Two methods for extracting calelectrin, a Ca2+-regulated membrane-binding protein from the electric organ of Torpedo marmorata, have been compared and the more promising one was modified to increase the yield to 7-8 mg.kg-1 wet weight of tissue, that is 4-5 times greater than the original method. The calelectrin so obtain could be resoloved into a minor component (designated L-calelectrin) eluted from an anion-exchange column at relatively low ionic strength (100 mM NaCl) and a major component (H-calelectrin) eluted at higher ionic strength (300 mM NaCl). The two forms were also separated by chromatography on a hydrophobic resin. Electrophoresis on cellulose acetate indicated that L-calelectrin had a lower mean isoelectric point that the H-form and polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulphate showed that under reducing conditions (presence of 5% beta-mercaptoethanol) both forms migrated as single species, the L-form having a lower apparent relative molecular mass (Mr 32,000) that the H-form (34,000). Under non-reducing conditions, there was no change in the migration of L-calelectrin but the H-form was resolved into two components of Mr 34,000 and 32,000. The addition of 2 mM Ca2+ had no effect on the migration of either form. Both forms were equally recognized by an anti-calelectrin antiserum and were microheterogeneous with respect to their isoelectric points (pH 4.3-5.5) in two-dimensional gel electrophoresis. Physical measurements were carried out on the major H-form. The Stokes radius was estimated to be 3nm, corresponding to an apparent Mr of 44,000. It was unaffected by changes in ionic strength, pH or Ca2+ concentration. Analytical ultracentrifugation gave a sedimentation constant of 2.9 S and an apparent Mr of 36,000. Measurements of circular dichroism indicated that 78% of the molecule was in the alpha-helix configuration and 22% in random coil. Ca2+ had no significant effect on the conformation.  相似文献   

5.
ORD and CD measurements of spectrin, in both the dimer and tetramer association state, indicate a high proportion of alpha-helix in this protein. At temperatures below 27 degrees C and in 0.1 M NaCl, the tetramer has an apparent helix content of 73% and the dimer, 68%. The conformation of both states is dependent on salt concentration and temperature. Low ionic strength solutions of spectrin display lowered sedimentation coefficients and a decreased apparent helix content, indicating perhaps a slight refolding and expansion of the molecule. In addition, spectrin in low ionic strength solutions undergoes a broad temperature-dependent transition spread from 20 to 50 degrees C, while in the presence of salt the transition is sharp and centered on 49 degrees C. The temperature-dependent changes in low ionic strength solutions appear to parallel the dissociation of tetramer to dimer.  相似文献   

6.
Recent sedimentation equilibrium measurements of the molecular weight of tail muscle lactate dehydrogenase from the North American East Coast lobster Homarus americanus show that this enzyme does not dissociate in buffer with high ionic strength (1.2 m ammonium sulfate). However, the apparent isopotential volume φ2′ increases significantly with increasing ionic strength of the solution. Consequently, molecular weight estimates for proteins using an assumed apparent specific volume equal to that in low salt concentration solutions may lead to erroneously low values under experimental conditions of high ionic strength.  相似文献   

7.
Insulin receptors were solubilized from rat liver microsomes by the nonionic detergent Triton X-100. After gel filtration of the extract on Sepharose CL-6B, two insulin-binding species (peak I and peak II) were obtained. The structure and binding properties of both peaks were characterized. Gel filtration yielded Stokes radii of 9.2 nm (peak I) and 8.0 nm (peak II). Both peaks were glycoproteins. At 4 degrees C peak I showed optimal insulin binding at pH 8.0 and high ionic strength. In contrast, peak II had its binding optimum at pH 7.0 and low ionic strength, where peak I binding was minimal. For peak I the change in insulin binding under different conditions of pH and ionic strength was due to a change in receptor affinity only. For peak II an additional change in receptor number was found. Both peaks yielded non-linear Scatchard plots under most of the buffer conditions examined. At their binding optima at 4 degrees C the high affinity dissociation constants were 0.50 nM (peak I) and 0.55 nM (peak II). Sodium dodecyl sulfate/polyacrylamide gel electrophoresis of peak I revealed five receptor bands with Mr 400 000, 365 000, 320 000, 290 000, and 245 000 under non-reducing conditions. For peak II two major receptor bands with Mr 210 000 and 115 000 were found. The peak II receptor bands were also obtained after mild reduction of peak I. After complete reduction both peaks showed one major receptor band with Mr 130 000. The reductive generation of the peak II receptor together with molecular mass estimations suggest that the peak I receptor is the disulfide-linked dimer of the peak II receptor. Thus, Triton extracts from rat liver microsomes contain two receptor species, which are related, but differ considerably in their size and insulin-binding properties.  相似文献   

8.
We have developed a practical analytical treatment of the non-linear Poisson-Boltzmann (P-B) equation to characterize the strong but non-specific binding of charged ligands to DNA and other highly charged macromolecules. These reactions are notable for their strong salt dependence and anti-cooperativity, features which the theory fully explains. We summarize analytical results for concentration profiles and ion binding in various regimes of surface curvature and ionic strength, and show how counterion size and charge distribution may influence competitive binding. We present several practical applications of the formalism, showing how to estimate the ligand concentration needed to effectively compete with a given buffer salt, and how to calculate the amounts of counterion species bound at various distances from the DNA surface under given bulk solution conditions. We cast our results into the form of a Scatchard binding isotherm, showing how the apparent binding constant K(obs) and S = -dlog K (obs )d log[M (+)] can be predicted from the basic theory. Anti-cooperativity arises naturally without steric repulsion, and binding curves can be fitted with K(obs) and effective charge as the only free parameters. We extend the analytical P-B analysis to an arbitrary number of counterion species, and apply the results to fit and predict three-ion competition data.  相似文献   

9.
The development of a simple, reliable method for determination of detergent micelle aggregation number that relies solely on measurement of steady-state fluorescence quenching is presented. The degree of steady-state fluorescence quenching of a micelle-solubilized fluorophore (pyrene) by a quencher that partitions greatly into the micelles (coumarin 153) is dependent on the micelle concentration, which can therefore be determined. The aggregation number is calculated as the micelle concentration/detergent monomer concentration (the total detergent concentration above the critical micelle concentration). For the determination to be accurate, the partition coefficient of the quencher into the micelle phase is determined and used to calculate the micellar concentration of quencher. Also, the quenching of pyrene by a coumarin 153 molecule within the same micelle must be complete, and this was confirmed by time-resolved fluorescence measurements. Aggregation numbers were determined for one cationic and several nonionic detergents and were found to be consistent with literature values. The approach presented is an improvement on a previous luminescence quenching technique (Turro, N.J., and A. Yekta. 1978. J. Am. Chem. Soc. 100:5951-5952) and can be used on cationic, anionic, and nonionic detergents with micelles ranging greatly in size and under varying conditions, such as detergent concentration, ionic strength, or temperature.  相似文献   

10.
The normal modes for a mixture of charged macromolecules and electrolyte solution are calculated. We derive a generalized Debye relaxation time and the apparent diffusion coefficient of the macroion, which is shown to increase from its Stokes value, obtained in excess of added salt, with decreasing ionic strength. We test our result with experimental data for macromolecules with different charge densities: heparin and chondroitin sulfate. Besides, we show for this latter molecule that while the diffusion coefficient is increased, the scattered intensity is decreased but not by the same factor. Our results are compared with other theories developed in quasielastic light scattering.  相似文献   

11.
Equilibrium measurements of the binding of central nervous system myelin basic protein to sodium dodecyl sulphate, sodium deoxycholate and lysophosphatidylcholine have been obtained by gel permeation chromatography and dialysis. This protein associates with large amounts of each of these surfactants: the apparent saturation weight ratios (surfactant/protein) being 3.58 +/- 0.12 and 2.30 +/- 0.15 for dodecyl sulphate at ionic strengths 0.30 and 0.10, respectively 1.34 +/- 0.10 for deoxycholate (at 0.12 ionic strength) and 4.0 +/- 0.5 for lysophosphatidylcholine. Binding to the ionic surfactants increases markedly close to their critical micelle concentrations. Sedimentation analysis shows that at 0.30 ionic strenght in excess dodecyl sulphate the protein is monomeric. It becomes dimeric when the binding ratio falls below 1 at a free detergent concentration of approximately 0.25 mM: below this concentration much of the protein and deterent forms an insoluble complex. The amount of dodecyl sulphate bound at high concentrations and at both above-mentioned ionic strengths corresponds closely to that expected for interaction of a single poly-peptide with two micelles. Variability of deoxycholate micelle size on interaction with other molecules precludes a similar analysis for this surfactant. Association was observed only with single micelles of lysophosphatidylcholine. The results provide strong evidence for dual lipid-binding sites on basic protein and indicate that lipid bilayer cross-linking by this protein may be effected by single molecules.  相似文献   

12.
The peripheral membrane M protein of vesicular stomatitis virus purified by detergent extraction of virions and ion-exchange chromatography was determined to be a monomer in the absence of detergent at high salt concentrations. Reduction of the ionic strength below 0.2 M resulted in a rapid aggregation of M protein. This self-association was reversible by the detergent Triton X-100 even in low salt. However, aggregation was not reversible by high salt concentration alone. M protein is initially synthesized as a soluble protein in the cytosol of infected cells, thus raising the question of how the solubility of M protein is maintained at physiological ionic strength. Addition of radiolabeled M protein purified from virions to unlabeled cytosol from either infected or uninfected cells inhibited the self-association reaction. Cytosolic fractions from infected or uninfected cells were equally effective at preventing the self-association of M protein. Self-association could also be prevented by an irrelevant protein such as bovine serum albumin. Sedimentation velocity analysis indicated that most of the newly synthesized M protein is monomeric, suggesting that the solubility of M protein in the cytosol is maintained by either low-affinity interaction with macromolecules in the cytosol or interaction of a small population of M-protein molecules with cytosolic components.  相似文献   

13.
We demonstrate that bovine core histones are natively unfolded proteins in solutions with low ionic strength due to their high net positive charge at pH 7.5. Using a variety of biophysical techniques we characterized their conformation as a function of pH and ionic strength, as well as correlating the conformation with aggregation and amyloid fibril formation. Tertiary structure was absent under all conditions except at pH 7.5 and high ionic strength. The addition of trifluoroethanol or high ionic strength induced significant alpha-helical secondary structure at pH 7.5. At low pH and high salt concentration, small-angle X-ray scattering and SEC HPLC indicate the histones are present as a hexadecamer of globular subunits. The secondary structure at low pH was independent of the ionic strength or presence of TFE, as judged by FTIR. The data indicate that histones are able to adopt five different relatively stable conformations; this conformational variability probably reflects, in part, their intrinsically disordered structure. Under most of the conditions studied the histones formed amyloid fibrils with typical morphology as seen by electron microscopy. In contrast to most aggregation/amyloidogenic systems, the kinetics of fibrillation showed an inverse dependence on histone concentration; we attribute this to partitioning to a faster pathway leading to non-fibrillar self-associated aggregates at higher protein concentrations. The rate of fibril formation was maximal at low pH, and decreased to zero by pH 10. The kinetics of fibrillation were very dependent on the ionic strength, increasing with increasing salt concentration, and showing marked dependence on the nature of the ions; interestingly Gdn.HCl increased the rate of fibrillation, although much less than NaCl. Different ions also differentially affected the rate of nucleation and the rate of fibril elongation.  相似文献   

14.
A model, developed within the framework of the counterion condensation theory of linear polyelectrolytes, is presented in this paper to describe the acid-base properties of linear polyelectrolytes, consisting of several types of functional ionizable groups. This formalism has been successfully applied to Fluka humic acid under salt-free conditions, as well as in the presence of supporting simple 1:1 salt (KNO3) at three different concentrations. As part of this approach, the charge density of the humic acid is obtained from the activity coefficient measurements of potassium counterions at different humic acid concentrations at a constant degree of dissociation of the polyelectrolyte. The humic acid average charge density was found to be 0.80 +/- 0.05. Using the present model, we are able to satisfactorily describe the experimental data obtained from acid-base potentiometric titrations. Four main functional groups making up the polymer are determined through their fractional abundances (Xi) and intrinsic pK (pK0i) values. The fractional abundances remained constant and independent of the ionic strength, indicating that the humic acid constitution does not depend on the concentration of excess salts. The pK0i values show a small change with ionic strength, which can be explained by the polyelectrolytic behavior of the solution.  相似文献   

15.
The proteins and lipids of the scallop gill ciliary membrane may be reassociated through several cycles of detergent solubilization, detergent removal, and freeze-thaw, without significant change in overall protein composition. Membrane proteins and lipids reassociate to form vesicles of uniform, discrete density classes under a variety of reassociation conditions involving detergent removal and concentration. Freed of the solubilizing detergent during equilibrium centrifugation, a protein-lipid complex equilibrates to a position on a sucrose density gradient characteristic of the original membrane density. When axonemal tubulin is solubilized by dialysis, mixed with 2:1 lecithin/cholesterol dissolved in Nonidet P-40, freed of detergent, and reconstituted by freeze-thaw, vesicles of a density essentially equal to pure lipid result. If the lipid fraction is derived through chloroform-methanol extraction of natural ciliary membranes, a moderate increase in density occurs upon reconstitution, but the protein is adsorbed and most is removed by a simple low ionic strength wash, in contrast to vesicles reconstituted from membrane proteins where even high salt extraction causes no loss of protein. The proteins of the ciliary membrane dissolve with constant composition, regardless of the type, concentration, or efficiency of detergent. Analytical ultracentrifugation demonstrates that monodisperse mixed micelles form at high detergent concentrations, but that membranes are dispersed to large sedimentable aggregates by Nonidet P-40 even at several times the critical micelle concentration, which suggests reasons for the efficacy of certain detergent for the production of ATP-reactivatable cell models. In extracts freed of detergent, structured polydisperse particles, but not membrane vesicles, are seen in negative staining; vesicles form upon concentration of the extract. Membrane tubulin is not in a form that will freely undergo electrophoresis, even in the presence of detergent above the critical micelle concentration. All chromatographic attempts to separate membrane tubulin from other membrane proteins have failed; lipid and protein are excluded together by gel filtration in the presence of high concentrations of detergent. These observations support the idea that a relatively stable lipid-protein complex exists in the ciliary membrane and that in this complex membrane tubulin is tightly associated with lipids and with a number of other proteins.  相似文献   

16.
Studies were done on the effect of bile salts on the rates of hydrolysis of the N-acetylneuraminyl linkages of several sialic acid-containing compounds by the neuraminidase of Clostridium perfringens. When GM3-ganglioside, two glycolipids (glycophorin and orosomucoid) and neuraminyl-lactose were used as substrates, hydrolysis was obtained even in the absence of bile salts, but addition of this detergent, below its critical micellar concentration, increased the reaction rates; above the critical micellar concentration of the detergent rates decreased again. When a second ganglioside, GM1, was used as substrate, the requirement for bile salts was absolute; hydrolysis was not observed at all without this detergent. With increasing concentrations of bile salt and in the presence of high concentrations of enzyme, rates of hydrolysis increased, reaching maximal values at fixed ratios of bile salt to GM1-ganglioside. Physical measurements showed that mixtures of bile salt and GM1-ganglioside form mixed micelles that have a higher critical micellar concentration, a lower molecular weight and greater axial ratio than the corresponding micelles of pure GM1-ganglioside.  相似文献   

17.
The cell body of Trypanosomatidae is enclosed in densely packed, crosslinked, subpellicular microtubules closely underlying the plasma membrane. We isolated the subpellicular microtubules from bloodstream Trypanosoma brucei parasites by use of a zwitterion detergent. These cold stable structures were solubilized by a high ionic strength salt solution, and the soluble proteins that contained tubulin along with several other proteins were further fractionated by Mono S cation exchange column chromatography. Two distinct peaks were eluted containing one protein each, which had an apparent molecular weight of 52 kDa and 53 kDa. (Mr was determined by SDS-gel electrophoresis). Only the 52 kDa protein showed specific tubulin binding properties, which were demonstrated by exposure of nitrocellulose-bound trypanosome proteins to brain tubulin. When this protein was added to brain tubulin in the presence of taxol and GTP, microtubule bundles were formed with regular crosslinks between the parallel closely packed microtubules. The crosslinks were about 7.2 nm apart (center to center). Under the same conditions, but with the 53 kDA protein or without trypanosome derived proteins, brain tubulin polymerized to single microtubules. It is thus suggested that the unique structural organization of the subpellicular microtubules is dictated by specific parasite proteins and is not an inherent property of the polymerizing tubulin. The in vitro reconstituted microtubule bundles are strikingly similar to the subpellicular microtubule network of the parasite.  相似文献   

18.
R B Cornell 《Biochemistry》1991,30(24):5873-5880
The activity of phosphocholine cytidylyltransferase (CT), the regulatory enzyme in phosphatidylcholine synthesis, is dependent on lipids. The enzyme, obtained from rat liver cytosol, was purified in the presence of Triton X-100 [Weinhold et al. (1986) J. Biol. Chem. 261, 5104]. The ability of lipids to activate CT when added as Triton mixed micelles was limited to anionic lipids. The relative effectiveness of the lipids tested suggested a dependence on the negative surface charge density of the micelles. The mole percent lipid in the Triton mixed micelle required for activation decreased as the net charge of the lipid varied from 0 to -2. Evidence for the physical association of CT with micelles and vesicles containing phosphatidylglycerol was obtained by gel filtration. The activation by micelles containing PG was influenced by the ionic strength of the medium, with a higher surface charge density required for activation at higher ionic strength. The micelle surface potential required for full activation of CT was calculated to be -43 mV. A specificity toward the structure of the polar group of the acidic lipids was not apparent. CT was activated by neutral lipids such as diacylglycerol or oleyl alcohol when included in an egg PC membrane, but the activities were reduced by dilution with as little as 10 mol % Triton. Thus Triton mixed micelles are not suitable for studying the activation of CT by these neutral lipid activators. We conclude that one way that lipid composition can control CT-membrane binding and activity is by changing the surface potential of the membrane. Other distinct mechanisms involved in the activation by neutral lipids are discussed.  相似文献   

19.
P E Morin  E Freire 《Biochemistry》1991,30(34):8494-8500
The kinetic and thermodynamic parameters associated with the enzymatic reaction of yeast cytochrome c oxidase with its biological substrate, ferrocytochrome c, have been measured by using a titration microcalorimeter to monitor directly the rate of heat production or absorption as a function of time. This technique has allowed determination of both the energetics and the kinetics of the reaction under a variety of conditions within a single experiment. Experiments performed in buffer systems of varying ionization enthalpies allow determination of the net number of protons absorbed or released during the course of the reaction. For cytochrome c oxidase the intrinsic enthalpy of reaction was determined to be -16.5 kcal/mol with one (0.96) proton consumed for each ferrocytochrome c molecule oxidized. Activity measurements at salt concentrations ranging from 0 to 200 mM KCl in the presence of 10 mM potassium phosphate, pH 7.40, and 0.5 mM EDTA display a biphasic dependence of the electron transferase activity upon ionic strength with a peak activity observed near 50 mM KCl. The ionic strength dependence was similar for both detergent-solubilized and membrane-reconstituted cytochrome c oxidase. Despite the large ionic strength dependence of the kinetic parameters, the enthalpy measured for the reaction was found to be independent of ionic strength. Additional experiments involving direct transfer of the enzyme from low to high salt conditions produced negligible enthalpy changes that remained constant within experimental error throughout the salt concentrations studied (0-200 mM KCl). These results indicate that the salt effect on the enzyme activity is of entropic origin and further suggest the absence of a major conformational change in the enzyme due to changes in ionic strength.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
A large scale mutation of the Rhodobacter capsulatus reaction center M-subunit gene, sym2-1, has been constructed in which amino acid residues M205-M210 have been changed to the corresponding L subunit amino acids. Two interconvertable spectral forms of the initial electron donor are observed in isolated reaction centers from this mutant. Which conformation dominates depends on ionic strength, the nature of the detergent used, and the temperature. Reaction centers from this mutant have a ground-state absorbance spectrum that is very similar to wild-type when measured immediately after purification in the presence of high salt. However, upon subsequent dialysis against a low ionic strength buffer or the addition of positively charged detergents, the near-infrared spectral band of P (the initial electron donor) in sym2-1 reaction centers is shifted by over 30 nm to the blue, from 852 to 820 nm. Systematically varying either the ionic strength or the amount of charged detergent reveals an isobestic point in the absorbance spectrum at 845 nm. The wild-type spectrum also shifts with ionic strength or detergent with an isobestic point at 860 nm. The large spectral separation between the two dominant conformational forms of the sym2-1 reaction center makes detailed measurements of each state possible. Both of the spectral forms of P bleach in the presence of light. Electrochemical measurements of the P/P+ midpoint potential of sym2-1 reaction centers show an increase of about 30 mV upon conversion from the long-wavelength form to the short-wavelength form of the mutant. The rate constant of initial electron transfer in both forms of the mutant reaction centers is essentially the same, suggesting that the spectral characteristics of P are not critical for charge separation. The short-wavelength form of P in this mutant also converts to the long-wavelength form as a function of temperature between room temperature and 130 K, again giving rise to an isobestic point, in this case at 838 nm for the mutant. A similar, though considerably less pronounced spectral change with temperature occurs in wild-type reaction centers, with an isobestic point at about 855 nm, close to that found by titrating with ionic strength or detergent. Fitting the temperature dependence of the sym2-1 reaction center spectrum to a thermodynamic model resulted in a value for the enthalpy of the conformational interconversion between the short- and long-wavelength forms of about -6 kJ/mol and an entropy of interconversion of about -35 J/(K mol). Similar values of enthapy and entropy changes can be used to model the temperature dependence in wild-type. Thus, much of the temperature dependence of the reaction center special pair near-infrared absorbance band can be described as an equilibrium shift between two spectrally distinct conformations of the reaction center.  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