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.
N L Gershfeld 《Biochemistry》1989,28(10):4229-4232
Thermodynamic properties of bilayer assembly have been obtained from measurements of the solubility of the sodium salt of dimyristoylphosphatidylglycerol (DMPG) in water. The standard free energy of bilayer assembly delta G degree a is shown to be RT 1n Xs + zF psi 0 where Xs is the mole fraction of dissolved lipid, F is the Faraday constant, z is the valence of the counterion (Na+), and psi 0 is the electrical double-layer potential of the ionized bilayer. The function d 1n Xs/dT was found to be discontinuous at 24 degrees C, the gel-liquid-crystal transition temperature (Tm) for DMPG. This function was unaffected when solubilities were measured in 0.001 M NaCl solutions; thus, psi 0 is constant in the experimental temperature interval (4-40 degrees C). Using a value of psi 0 = -180 mV [Eisenberg et al. (1979) Biochemistry 18, 5213-5223], and the temperature dependence of delta G degrees a, values for delta H degrees a and delta S degree a at 24 degrees C were calculated for the gel and liquid-crystal states of DMPG. For the gel, delta H degrees a and T delta S a are -26.2 and 12.7 kcal/mol, respectively; for the liquid-crystal, delta H degrees a and T delta S degrees a are -19.2 and -5.7 kcal/mol, respectively. The calculated value for the latent heat of the gel-liquid-crystal transition is 7 kcal/mol, in agreement with calorimetric measurements.  相似文献   

2.
Myosin subfragment-1 (S-1) which contains the LC2 light chain has been labelled with fluorine to allow an 19F-NMR study of the coupling and energetics of structural changes in the myosin head. Two fluorine-containing reagents, N-4-(trifluoromethyl)phenyl iodoacetamide and N-3,5-di(trifluoromethyl)phenyl iodoacetamide, have been used to label the myosin heavy chain at the unusually reactive sulfhydryl-1 (SH1) position. The chemical shift of both reagents on S-1 is sensitive to a structural transition in the region of SH1 which occurs upon increasing the temperature from 0 degrees C to 35 degrees C. The midpoint of the transition in both papain and chymotryptic S-1 is at approximately 11 degrees C at pH 7 (0.1 M CKl). The temperature dependence of the chemical shift may be fit assuming a two-state equilibrium where delta G degree' (T) = 101-110T +0.386 T2 (where T is the temperature in Kelvin). Both delta H degree' (T) and delta S degree' (T) have a small temperature dependence from 0 to 35 degrees C: at 20 degrees C, delta H degree' (T) = -33 kcal/mol. delta S degree' (T) = -116 e.u. and delta Cp = -226 cal/mol per deg (pH 7.0, 0.1 M KCl). The NMR data indicate that the presence of the LC2 light chain in papain S-1 does not modify the structure of S-1 in the vicinity of SH1, nor does it modify the energetics of the structural transition from that seen in its absence with chymotryptic S-1. The presence of calcium which is bound by the LC2 of papain S-1 also does not alter the energetics of the transition. Thus it would appear that the LC2 light chain (on myosin S-1) does not participate in the two-state transition, nor does it interact strongly with regions of the heavy chain which participate in the transition.  相似文献   

3.
N-Formylmethanofuran(CHO-MFR):tetrahydromethanopterin(H4MPT) formyltransferase (formyltransferase) from the extremely thermophilic Methanopyrus kandleri was purified over 100-fold to apparent homogeneity with a 54% yield. The monomeric enzyme had an apparent molecular mass of 35 kDa. The N-terminal amino acid sequence of the polypeptide was determined. The formyltransferase was found to be absolutely dependent on the presence of phosphate or sulfate salts for activity. The ability of salts to activate the enzyme decreased in the order K2HPO4 > (NH4)2SO4 > K2SO4 > Na2SO4 > Na2HPO4. The salts KCl, NaCl and NH4Cl did not activate the enzyme. The dependence of activity on salt concentration showed a sigmoidal curve. For half-maximal activity, 1 M K2HPO4 and 1.2 M (NH4)2SO4 were required. A detailed kinetic analysis revealed that phosphates and sulfates both affected the Vmax rather than the Km for CHO-MFR and H4MPT. At the optimal salt concentration and at 65 degrees C, the Vmax was 2700 U/mg (1 U = 1 mumol/min), the Km for CHO-MFR was 50 microM and the Km for H4MPT was 100 microM. At 90 degrees C, the temperature optimum of the enzyme, the Vmax was about 2.5-fold higher than at 65 degrees C. Thermostability as well as activity of formyltransferase was dramatically increased in the presence of salts, 1.5 M being required for optimal stabilization. The efficiency of salts in protecting formyltransferase from heat inactivation at 90 degrees C decreased in the order K2HPO4 = (NH4)2SO4 > KCl = NH4Cl = NaCl > Na2SO4 > Na2HPO4. The catalytic mechanism of formyltransferase was determined to be of the ternary-complex type. The properties of the enzyme from M. kandleri are compared with those of formyltransferase from Methanobacterium thermoautotrophicum, Methanosarcina barkeri and Archaeoglobus fulgidus.  相似文献   

4.
The fluorescence decay of 1,N6-ethenoadenosine diphosphate (epsilon ADP) bound to myosin subfragment 1 (S1) was studied as a function of temperature. The decay was biexponential, and the two lifetimes were quenched relative to the single lifetime of free epsilon ADP. The temperature dependence of the fractional intensities of the decay components showed two states of the S1.epsilon ADP complex. At pH 7.5 in 30 mM TES, 60 mM KCl, and 3 mM MgCl2, the equilibrium constant for the conversion of the low-temperature state (S1L.epsilon ADP) to the high-temperature state (S1H.epsilon ADP) was 40 at physiological temperatures, and delta H degrees = 13 kcal.mol-1 and delta S degrees = 49 cal.deg-1.mol-1. At 10 degrees C the equilibrium constant of S1 for epsilon ADP was 5, indicating that S1H.epsilon ADP was the dominant state, and that for the vanadate complex epsilon ADP.Vi was 0.7, suggesting that in S1.epsilon ADP.Vi the dominant state of the S1-nucleotide complex was converted from S1H.epsilon ADP to S1L.epsilon ADP. The single rotational correlation time of bound epsilon ADP at 10 degrees C decreased from 107 ns in S1.epsilon ADP to 74 ns in S1+.epsilon ADP.Vi. Conversion of the binary complex to the ternary vanadate complex resulted in a 3-A decrease in the energy transfer distance between bound epsilon ADP and N-[4-(dimethylamino)-3,5-dinitrophenyl]maleimide attached to SH1 and a decrease of the average distance between bound epsilon ADP and bound Co2+ from 12.6 to 8.3 A.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Fractionated polyuridylic acid with an average chain length of 55 nucleotides forms binary complexes with 30S subunits with a stoichiometry of I:I. These complexes are heterogeneous in stability. The more stable one is characterized by an association constant K2 - 5.5xI09 M-I, and the less stable-by KI = I06xM-I, at 20 mM Mg2+, 200 mM NH4(+) and 0 degrees C. The main reason for this heterogeneity is the presence or absence of the ribosomal protein SI in the presence or absence of the ribosomal protein SI in the subunits. Decrease of Mg2+ concentration down to 5 mM hardly changes the K2 values but reduction of the NH4(+) concentration to 50 mM results in a 25-fold increase of K2. Association constants K2 for the stable complex, i.e. in the presence of SI protein, were measured at different temperatures (0 - 30 degrees C) and the thermodynamic parameters of binding (delta H degrees, delta S degrees, delta G degrees) were determined. Analogous experiments were made with 70S ribosomes. K2 values as well as delta H degrees, delta S degrees, delta G degrees appeared the same both for 30S and 70S ribosomes in all conditions examined. This is strong evidence that the 50S subunits do not contribute to the interaction of poly(U) with the complete 70S ribosomes.  相似文献   

6.
K S Bruzik  M D Tsai 《Biochemistry》1987,26(17):5364-5368
The phase-transition properties of sphingomyelins were investigated in detail with totally synthetic, chemically and stereochemically pure (2S,3R)-(N-stearoylsphingosyl)-1-phosphocholine (D-erythro-C18-SPM) (1) and the corresponding 2S,3S isomer (L-threo-C18-SPM) (2). Heating scans of an unsonicated dispersion of 1 right after hydration showed a main transition (I) at 44.7 degrees C (delta H = 6.8 kcal/mol). Upon incubation at 20-25 degrees C a second transition (II) appeared at 36.0 degrees C (delta H = 5.7 kcal/mol). The two gel phases were designated as G alpha and G beta phases, respectively. The G beta phase was also metastable and relaxed to a third gel phase (G gamma) upon incubation below 10 degrees C. Conversion of the G gamma phase to the liquid-crystalline phase occurred via two new endotherms at 33.4 degrees C (2.6 kcal/mol) (III) and 43.6 degrees C (8.0 kcal/mol) (IV) as well as a main transition at 44.7 degrees C (9.5 kcal/mol). Possible interpretations have been proposed to account for the observed phase transitions. The L-threo isomer 2 showed similar thermotropic behavior to dipalmitoylphosphatidylcholine (DPPC): a "main transition" at 44.2 degrees C (6.0 kcal/mol), a "pretransition" at 43.1 degrees C (1.8 kcal/mol), and upon incubation at 7 degrees C for 2 weeks, a very broad "subtransition" at ca. 35 degrees C. The results are substantially different from previous studies of sphingomyelins using mixtures of stereoisomers. Mixing of 1 with 2, 1 with DPPC, and 2 with DPPC removed the metastability of the gel phase and resulted in a single transition.  相似文献   

7.
Extent of adsorption (gamma pw) of bovine serum albumin, beta-lactoglobulin, gelatin and myosin at the alumina-water interface has been measured as function of protein concentration (Cp) at several temperatures, pH, and ionic strengths of the medium. gamma pw for proteins in most cases increases with increase of protein concentration but it attains maximum value gamma pw(m) when Cp is high. Values of maximum adsorption have been examined in terms of molecular orientation, molecular size and shape and unfolding of the packed proteins at the interface. In few cases, gamma pw increases with increase of Cp without reaching a real state of saturation as a result of aggregation of molecules or extensive unfolding of the protein at the interface. In the case of beta-lactoglobulin at pH 5.2 and ionic strength 0.05, gamma pw in high concentration region decreases to zero value when Cp increases. For myosin at 45 degrees C and pH 6.4, and also at 27 degrees and pH 7.8, the values of gamma pw are all negative and these negative values increase with increase of Cp. All these results have been explained in terms of significant competitions of water and protein for binding to the surface sites of the powdered alumina. Adsorption of myosin has also been found to be affected in the presence of NaCl, KCl, CaCl2, KI, Na2SO4, LiCl and urea. The relative affinities of the adsorption of various proteins for the surface of alumina at different physical conditions of the system have been compared in terms of maximum values of adsorption attained when gamma pw is varied with Cp. The affinities are shown to be compared more precisely in terms of the standard free energy decrease for the saturation of the surface by protein as a result of the change in its concentration from zero to unity in the mole fraction scale.  相似文献   

8.
S H Lin  H C Cheung 《Biochemistry》1991,30(17):4317-4322
We previously reported that the nucleotide complex of myosin subfragment 1, S1.epsilon ADP, exists in two states on the basis of the temperature dependence of the fluorescence decay of bound 1,N6-ethenoadenosine diphosphate (epsilon ADP) [Aguirre, R., Lin. S.-H., Gonsoulin, F., Wang, C.-K., & Cheung, H.C. (1989) Biochemistry 28, 799-809]. We have extended the previous study of the equilibrium between the two states, S1L.ADP in equilibrium S1H.ADP, by using a fluorescently labeled myosin S1 (S1-AF). In S1 alkylated with IAF [5-(iodoacetamido)fluorescein], the decay of the label emission was biexponential both in the presence and absence of ADP and/or actin. In the presence of ADP, the two decay times were 4.30 (alpha 1 = 0.55) and 0.80 ns (alpha 2 = 0.45) at 12.4 degrees C, in a medium containing 60 mM KCl, 30 mM TES (pH 7.5), and 2 mM MgCl2. The steady-state fluorescence intensities of S1-AF, (S1-AF).ADP, acto.(S1-AF), and acto.(S1-AF).ADP were dependent on temperature over the range of 5-30 degrees C. By combining lifetime and steady-state intensity data, we obtained for the two-state transition (S1-AF)L.ADP in equilibrium (S1-AF)H.ADP the following parameters: delta H degrees = 16.1 kcal/mol (67.3 kJ/mol) and delta S degrees = 55.8 cal/(deg.mol) [233.5 J/(deg.mol)], in agreement with previous results obtained with epsilon ADP. The delta H degrees values for the two-state transition of S1-AF, acto.(S1-AF), and acto.(S1-AF).ADP are 13.0, 21.6, and 5.2 kcal/mol, respectively. The corresponding delta S degrees values are 46.9, 79.5, and 17.4 cal/(deg.mol).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Crystallization of phosphatidylserine bilayers induced by lithium   总被引:1,自引:0,他引:1  
Utilizing differential scanning calorimetry and x-ray diffraction, 1,2-dimyristoyl-L-glycero-3-phospho-L-serine (DMPS) was shown to form hydrated bilayer membrane structures exhibiting a gel leads to liquid crystalline transition at 39 degrees C (delta H = 7.2 kcal/mol). Addition of up to molar concentrations of the alkali halides NaCl, KCl, Rl Cl, and CsCl produced relatively minor changes in DMPS bilayer structure or stability. For example, in the presence of 0.5 M NaCl, the transition temperature (Tc = 42 degrees C) and transition enthalpy (delta H = 7.0 kcal/mol) show only minor changes. In marked contrast, addition of LiCl results in "'crystallization" of the DMPS bilayer membrane structure. At 0.5 M LiCl, the crystalline DMPS exhibits a bilayer gel leads to liquid crystal transition at 89 degrees C accompanied by a high enthalpy change, delta H = 16.0 kcal/mol. Thus, Li+ induces an isothermal crystallization of DMPS bilayers, the hydrocarbon chains adopting a more ordered packing mode than the "hexagonal" arrangement of the gel state. In view of the widespread use of lithium in the treatment of manic-depressive illness, we also raise the possibility that interaction of Li+ with anionic membrane phospholipids could play a role in its pharmacological action.  相似文献   

10.
Hydroxylamine oxidoreductase (HAO) of the ammonia-oxidizing bacterium Nitrosomonas catalyzes the oxidation: NH2OH + H2O----HNO2 + 2e- + 2 H+. The heme-like chromophore P460 is part of a site which binds substrate, extracts electrons and then passes them to the many c hemes of the enzyme. Reduction of the c hemes by hydroxylamine is biphasic with apparent first-order rate constants k1 and k2. CO binds to ferrous P460 with apparent first-order rate constants, k1,CO. In this work we have measured the binding of CO to ferrous P460 of hydroxylamine oxidoreductase and the reduction by substrate of some of the 24 c hemes of the ferric enzyme. These reactions have been studied in water and 40% ethylene glycol, at temperatures ranging from -15 degrees C to 20.7 degrees C and at hydrostatic pressures ranging over 0.1-80 MPa. From the measurements, thermodynamic parameters delta V+ (activation volume), delta G+, delta H+, and delta S+ have been calculated. CO binding. Binding of CO to ferrous P460 was similar to the binding of CO to ferrous horseradish peroxidase. The change of solvent had only a limited effect on delta V+ (-30 ml.mol-1), delta G+, delta H+ or delta S+ and did not cause an inflection in the Arrhenius plot or downward displacement of the linear relationship between ln k1,CO and P at a critical temperature. Binding was exothermic at high temperatures. The response of the binding of CO to solvent, temperature and pressure suggested that the CO binding site had little access to solvent and was not susceptible to change in protein conformation. Fast phase of reduction of c hemes. Changing the solvent from water to 40% ethylene glycol resulted in a decrease from 90% to 50% in the relative number of c hemes reduced during the fast phase, an increase in activation volume from -3.6 ml.mol-1 to 57 ml.mol-1 and changes in other thermodynamic parameters. The activation volume increased with decreasing temperature. The Arrhenius plot had a downward inflection at about 0 degrees C and, in water or ethylene glycol, the linear dependence of ln k1 on P was displaced downwards as the temperature changed from 3.5 degrees C to -15 degrees C. Slow phase of reduction of c hemes. Changing the solvent from water to 40% ethylene glycol resulted in an increase in the relative number of c hemes reduced during the slow phase from 10% to 50%. The activation volume, which was not measurable in water because of the low absorbance change, was -30 ml.mol-1 in ethylene glycol. The activation volume increased with increasing temperature.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
Protein fractions of a higher-oligomer (H), (alphabeta)(2)-diprotomer (D) and alphabeta-protomer (P) were separated from dog kidney Na(+)/K(+)-ATPase solubilized in the presence of NaCl and KCl. Na(+)/K(+)-dependent interconversion of the oligomers was analysed using HPLC at 0 degrees C. With increasing KCl concentrations, the content or amount of D increased from 27.6 to 54.3% of total protein, i.e. DeltaC(max) = 26.7%. DeltaC(max) for the sum of D and H was equivalent to the absolute value of DeltaC(max) for P, regardless of the anion present, indicating that K(+) induced the conversion of P into D and/or H, and Na(+) had the opposite effect. When enzymes that had been denatured to varying degrees by aging were solubilized, DeltaC(max) increased linearly with the remaining ATPase activity. The magnitude of the interconversion could be explained based on an equilibrium of D <==> 2P, assuming 50-fold difference in the K(d) between KCl and NaCl, and coexistence of unconvertible oligomers, which comprised as much as 39% of the eluted protein. Oligomeric interconversion, determined as a function of the KCl or NaCl concentration, showed K(0.5)s of 64.8 microM and 6.50 mM for KCl and NaCl, respectively, implying that oligomeric interconversion was coupled with Na(+)/K(+)-binding to their active transport sites.  相似文献   

12.
Tris buffer causes acyl chain interdigitation in phosphatidylglycerol   总被引:2,自引:0,他引:2  
The structure of the gel phase and the properties of the acyl chain disordering transition of dipalmitoyl phosphatidylglycerol (DPPG) have been studied using differential scanning calorimetry, differential scanning dilatometry, and X-ray diffraction. In the presence of small, monovalent cations, DPPG at 22 degrees C exists in a lamellar phase in which the hydrocarbon chains are tilted from the perpendicular to the bilayer surface. Around 34 degrees C, there is a small pretransition (delta H less than 1 kcal/mol) followed by the main transition at 40.4 degrees C (delta H = 8.3 kcal/mol; delta V = 0.0381 ml/g). If DPPG is suspended in Tris-HCl buffer in the absence of other monovalent cations, X-ray diffraction data show that at 22 degrees C, the gel phase consists of interdigitated acyl chains perpendicular to the plane of the bilayer. No pretransition is observed and the main transition occurs at 41.3 degrees C with delta H = 9.1 kcal/mol and delta V = 0.0514 ml/g. If sufficient Na+ or K+ ions are added to the Tris-buffered DPPG, the phase behavior reverts to what is observed in the absence of Tris. Analysis of the energetics of the main transition shows that the increase in van der Waals interaction energy resulting from the larger delta V in Tris can be compensated by the favorable energetics of removing terminal methyl groups from the bilayer surface. The amount of disordering, i.e. formation of gauche rotamers, is likely to be the same in Tris as it is in buffers without amphiphilic cations.  相似文献   

13.
The kinetics of association and dissociation for the ouabain-Na+,K+- dependent ATPase complex have been studied in intact turkey erythrocytes as a function of external Na+ concentration, K+ concentration, and temperature. At free ligand concentrations substantially exceeding the concentration of available binding sites, the association reaction exhibits pseudo-first-order kinetics with an association rate constant (k1) that is conveniently determined over a wide range of temperatures (5-37 degrees C). The dissociation reaction exhibits strict first-order kinetics with a dissociation rate constant (k-1) that has the unusual property, in the turkey cell, of being sufficiently great to permit its direct determination even at temperatures as low as 5 degrees C. Values for the equilibrium binding constant for the ouabain-ATPase complex (KA) predicted from the ratio of the association and dissociation rate constants agree closely with independently measured values of KA determined directly under conditions of equilibrium binding. KA is a sensitive function of the composition of the external ionic environment, rising with increasing Na+ concentration and falling with increasing K+ concentration. These changes in KA are shown to be quantitatively attributable to changes in the rate constant k1, k-1 in contrast being unaffected at any given temperature by even very large changes in Na+ or K+ concentration. Arrhenius plots of k1 and k-1 both yield straight lines over the entire temperature range corresponding to activation energies for association and dissociation of 29.5 and 24.2 kcal/mol, respectively. These observations have made it possible to calculate the following standard values for the ouabain binding reaction in the presence of 150 mM Na+: delta G degree = -9.8 kcal/mol; delta H degree = +5.3 kcal/mol; delta S degree = +48.7 cal/degree/mol. The large positive value of delta S degree presumably reflects a highly ordered configuration of the ouabain-free ATPase molecule that is lost upon ouabain binding and that "drives" the reaction despite the positive value of delta H degree.  相似文献   

14.
We studied the interaction between the 35 kDa apolipoprotein of canine pulmonary surfactant (SP 35) and five saturated phosphatidylcholines: distearoyl (DSPC), diheptadecanoyl (DHPC), dipalmitoyl (DPPC), dimyristoyl (DMPC), and dilauroyl (DLPC); and two monoenoic unsaturated phosphatidylcholines: dioleoyl (DOPC) and dielaidyl (DEPC), using temperatures at which all of the phospholipids except DOPC were in both the gel and liquid-crystalline states. The experiments were carried out in a buffer without Ca2+. The amount of apolipoprotein which was bound by both small unilamellar and multilayered vesicles of these lipids decreased as the temperature was increased. Moreover, near the temperatures of the phase transitions of all lipids except DLPC, there was an abrupt and marked reduction in binding of protein, in that over a 3-4 degree change in temperature there was an abrupt decrease in bound apolipoprotein. A similar change in binding occurred using DLPC, although the relatively large changes in bound protein occurred at about 10 and 20 degrees C, temperatures which are above the phase transition temperature of this lipid. Experiments using DOPC were limited to temperatures above the phase transition, and apolipoprotein binding was low. Experiments monitoring the intrinsic fluorescence of the protein, and the fluorescence of bis-1-anilino-8-naphthalene sulfonic acid bound to the protein, revealed a possible conformational change at about 40 degrees C. Measurement of intrinsic fluorescence provided the same result whether or not the protein was associated with lipid. DSC of the apolipoprotein indicated that this change was not associated with a measurable thermogenic process. We found that the interaction with DPPC was reversible at 42 degrees C, and we measured the thermodynamic parameters of the interaction at this temperature. These were: delta G0 = -8.0 kcal/mol apolipoprotein; delta H0 = -88 kcal/mol; delta S0 = -254 cal/Cdeg per mol. We conclude that the interaction between SP 35 and saturated phosphatidylcholines is temperature sensitive, and this probably reflects differences in the ability of gel and liquid-crystalline phospholipids to bind this protein. Both the delta H0 and delta S0 of the interaction are negative, and may reflect an immobilization of phospholipid around the apolipoprotein to form a boundary layer. This hypothesis is consistent with the findings obtained by DSC, in which the enthalpy of the phase transition of DMPC in lipid-apolipoprotein recombinants was found to be about 60% of that expected for a pure and unperturbed multilamellar dispersion.  相似文献   

15.
The time course of the precipitin reactions of concanavalin A with glycogen, dextran and ovalbumin was investigated by a light-scattering method near 30 degrees C in 10 mM-Tris/HCl buffer, pH 7.4, containing neutral salts, i.e. NaCl, KCl, NaBr, KI and NaClO4. With 0.8 microM-lectin and 0.36 mg of glycogen/ml, the half-life, t 1/2, of the precipitin reaction was independent of salt concentration between 0.1 M and 1.5 M, and was the same (175s) in the presence of NaCl, KCl, NaBr and KI but was significantly (27%) higher in NaClO4. In contrast, the five salts caused significant to marked enhancement in t 1/2 for the reactions of concanavalin A with dextran and ovalbumin. Likewise, whereas the turbidity produced in 1 h as a result of lectin-glycogen precipitation remained unchanged, those measured for the binding of dextran and ovalbumin were decreased in the presence of three salts. The increase in t 1/2 and decrease in turbidity were found to be higher with NaClO4, followed by KI; NaBr produced moderate and NaCl (or KCl) small but generally significant inhibition of the precipitin reactions with dextran and ovalbumin. The results showed that the lectin-ligand precipitin reactions involve salt-sensitive polar interactions that are less pronounced with compactly folded ligands such as glycogen.  相似文献   

16.
Methanobacterium thermoautotrophicum delta H and Marburg were adapted to grow in medium containing up to 0.65 M NaCl. From 0.01 to 0.5 M NaCl, there was a lag before cell growth which increased with increasing external NaCl. The effect of NaCl on methane production was not significant once the cells began to grow. Intracellular solutes were monitored by nuclear magnetic resonance (NMR) spectroscopy as a function of osmotic stress. In the delta H strain, the major intracellular small organic solutes, cyclic-2,3-diphosphoglycerate and glutamate, increased at most twofold between 0.01 and 0.4 M NaCl and decreased when the external NaCl was 0.5 M. M. thermoautotrophicum Marburg similarly showed a decrease in solute (cyclic-2,3-diphosphoglycerate, 1,3,4,6-tetracarboxyhexane, and L-alpha-glutamate) concentrations for cells grown in medium containing > 0.5 M NaCl. At 0.65 M NaCl, a new organic solute, which was visible in only trace amounts at the lower NaCl concentrations, became the dominant solute. Intracellular potassium in the delta H strain, detected by atomic absorption and 39K NMR, was roughly constant between 0.01 and 0.4 M and then decreased as the external NaCl increased further. The high intracellular K+ was balanced by the negative charges of the organic osmolytes. At the higher external salt concentrations, it is suggested that Na+ and possibly Cl- ions are internalized to provide osmotic balance. A striking difference of strain Marburg from strain delta H was that yeast extract facilitated growth in high-NaCl-containing medium. The yeast extract supplied only trace NMR-detectable solutes (e.g., betaine) but had a large effect on endogenous glutamate levels, which were significantly decreased. Exogenous choline and glycine, instead of yeast extract, also aided growth in NaCl-containing media. Both solutes were internalized with the choline converted to betaine; the contribution to osmotic balance of these species was 20 to 25% of the total small-molecule pool. These results indicate that M. thermoautotrophicum shows little changes in its internal solutes over a wide range of external NaCl. Furthermore, they illustrate the considerable differences in physiology in the delta H and Marburg strains of this organism.  相似文献   

17.
The gel formed by warming the 100,000 g supernate of isotonic extracts of sea urchin eggs to 40 degrees C is made up of actin and two additional proteins of mol wt of 58,000 and 220,000. Actin and 58,000 form a characteristic structural unit which has now been identified in the microvilli of the urchin egg and in the filopods of urchin coelomocytes. However, egg extract gels did not contract as those from other cell types do, and the aim of these experiments was to determine the reason for this lack of contraction. Although the extracts are dialyzed to a low ionic strength, myosin is present in soluble form and makes up approximately 1% of the protein of the extract. It becomes insoluble in the presence of high ATP concentrations at 0 degrees C, and the precipitate formed under these conditions consists almost entirely of myosin. This procedure provides a simple method of isolating relatively pure myosin without affecting other extract components and functions. Contraction will follow gelation in these extracts if the temperature and time of incubation used to induce actin polymerization are reduced to minimize myosin inactivation. At the optimal ATP and KCl concentration for contraction, the contracted material has an additional 250,000 component and contains very little 58,000. The conditions found to provide maximum gel yields favor the formation of the actin-58,000-220,000 structural gel, while reduced temperature and increase in KCl concentration results in a contractile gel whose composition is similar to those reported from amoeboid cell types. Both the structural protein cores found in the egg microvilli and a gel contraction related to the amoeboid motion which is seen in later urchin embryonic development can thus be induced in vitro in the same extract.  相似文献   

18.
P Alexander  S Fahnestock  T Lee  J Orban  P Bryan 《Biochemistry》1992,31(14):3597-3603
We have cloned, expressed, and characterized two naturally occurring variations of the IgG-binding domain of streptococcal protein G. The domain is a stable cooperative folding unit of 56 amino acids, which maintains a unique folded structure without disulfide cross-links or tight ligand binding. We have studied the thermodynamics of the unfolding reaction for the two versions of this domain, designated B1 and B2, which differ by six amino acids. They have denaturation temperatures of 87.5 degrees C and 79.4 degrees C, respectively at pH 5.4, as determined by differential scanning calorimetry. Thermodynamic state functions for the unfolding reaction (delta G, delta H, delta S, and delta Cp) have been determined and reveal several interesting insights into the behavior of very small proteins. First, though the B1 domain has a heat denaturation point close to 90 degrees C, it is not unusually stable at physiologically relevant temperatures (delta G = 25 kJ/mol at 37 degrees C). This behavior occurs because the stability profile (delta G vs temperature) is flat and shallow due to the small delta S and delta Cp for unfolding. Related to this point is the second observation that small changes in the free energy of unfolding of the B-domain due to mutation or change in solvent conditions lead to large shifts in the heat denaturation temperature. Third, the magnitude and relative contributions of hydrophobic vs nonhydrophobic forces (per amino acid residue) to the total free energy of folding of the B-domain are remarkably typical of other globular proteins of much larger size.  相似文献   

19.
Dietary potassium (K(+)) restriction and hypokalemia have been reported to change the abundance of most renal Na(+) and K(+) transporters and aquaporin-2 isoform, but results have not been consistent. The aim of this study was to reexamine Na(+), K(+) and H(2)O transporters' pool size regulation in response to removing K(+) from a diet containing 0.74% NaCl, as well as from a diet containing 2% NaCl (as found in American diets) to blunt reducing total diet electrolytes. Sprague-Dawley rats (n = 5-6) were fed for 6 days with one of these diets: 2% KCl, 0.74% NaCl (2K1Na, control chow) compared with 0.03% KCl, 0.74% NaCl (0K1Na); or 2% KCl, 2%NaCl (2K2Na) compared with 0.03% KCl, 2% NaCl (0K2Na, Na(+) replete). In both 0K1Na and 0K2Na there were significant decreases in: 1) plasma [K(+)] (<2.5 mM); 2) urinary K(+) excretion (<5% of control); 3) urine osmolality and plasma [aldosterone], as well as 4) an increase in urine volume and medullary hypertrophy. The 0K2Na group had the lowest [aldosterone] (172.0 ± 17.4 pg/ml) and lower blood pressure (93.2 ± 4.9 vs. 112.0 ± 3.1 mmHg in 2K2Na). Transporter pool size regulation was determined by quantitative immunoblotting of renal cortex and medulla homogenates. The only differences measured in both 0K1Na and 0K2Na groups were a 20-30% decrease in cortical β-ENaC, 30-40% increases in kidney-specific Ste20/SPS1-related proline/alanine-rich kinase, and a 40% increase in medullary sodium pump abundance. The following proteins were not significantly changed in both the 0 K groups: Na(+)/H(+) exchanger isoform 3; Na(+)-K(+)-Cl(-) cotransporter; Na(+)-Cl(-) cotransporter, oxidative stress response kinase-1; renal outer medullary K(+) channel; autosomal recessive hypercholesterolemia; c-Src, aquaporin 2 isoform; or renin. Thus, despite profound hypokalemia and renal K(+) conservation, we did not confirm many of the changes that were previously reported. We predict that changes in transporter distribution and activity are likely more important for conserving K(+) than changes in total abundance.  相似文献   

20.
Adsorption isotherms of BSA at the solid-water interfaces have been studied as a function of protein concentration, ionic strength of the medium, pH and temperature using silica, barium sulphate, carbon, alumina, chromium, ion-exchange resins and sephadex as solid interfaces. In most cases, isotherms for adsorption of BSA attained the state of adsorption saturation. In the presence of barium sulphate, carbon and alumina, two types in the isotherms are observed. Adsorption of BSA is affected by change in pH, ionic strength and temperature of the medium. In the presence of metallic chromium, adsorbed BSA molecules are either denatured or negatively adsorbed at the metallic interface. Due to the presence of pores in ion-exchange resins, adsorption of BSA is followed by preferential hydration on resin surfaces in some cases. Sometimes two steps of isotherms are also observed during adsorption of BSA on the solid resins in chloride form. Adsorption of BSA, beta-lactoglobulin, gelatin, myosin and lysozyme is negative on Sephadex surface due to the excess adsorption of water by Sephadex. The negative adsorption is significantly affected in the presence of CaCl2, KSCN, LiCl, Na2SO4, NaI, KCl and urea. The values of absolute amounts of water and protein, simultaneously adsorbed on the surface of different solids, have been evaluated in some cases on critical thermodynamic analysis. The standard free energies (delta G0) of excess positive and negative adsorption of the protein per square meter at the state of monolayer saturation have been calculated using proposed universal scale of thermodynamics. The free energy of adsorption with reference to this state is shown to be strictly comparable to each other. The magnitude of standard free energy of transfer (delta G0B) of one mole of protein or a protein mixture at any type of physiochemical condition and at any type of surface is observed to be 38.5 kJ/mole.  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