首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The thermotropic properties of bovine blood coagulation Factors IX and X, as well as the activation intermediates and products of these proteins, have been investigated by differential scanning microcalorimetry in the presence and absence of Ca2+. Bovine Factor IX displays a single thermal-denaturation transition characterized by a temperature midpoint (TM) of 54.5 ± 0.5 °C and a calorimetric enthalpy (ΔHc) of 105 ± 15 kcal/mol, in the absence of Ca2+. In the presence of Ca2+ concentrations sufficient to saturate its sites on Factor IX, the Tm value is increased to 57.0 ± 0.5 °C and the ΔHc is virtually unchanged. When the activation intermediate, Factor IXα, is similarly analyzed in the absence of Ca2+, a broad, diffuse thermogram was obtained which did not lend itself to calculation of thermodynamic parameters. In the presence of Ca2+, Factor IXα displayed thermograms characterized by a TM of 51.0 ± 0.5 °C and a ΔHc of 109 ± 10 kcal/mol. The activated product, Factor IXaα, in the absence of Ca2+ (the values in the presence of saturating Ca2+ are given in parentheses), undergoes thermal denaturation with a TM of 54.5 ± 0.5 °C (57.0 ± 0.5 °C) and a ΔHc of 158 ±10 kcal/mol (156 ± 10 kcal/mol). Similarly, the terminal-activation product, Factor IXaβ, displays a TM of 51.5 ± 0.5 °C (54.0 ± 0.5 °C) and a ΔHc of 85 ± 5 kcal/mol (126 ± 10 kcal/mol). Bovine blood coagulation Factor X has been analyzed in this same fashion, and shows very similar thermal properties to Factor IX. The thermal denaturation of Factor X is represented by a TM of 54.0 ± 0.5 °C (55.0 ± 0.5 °C) and a ΔHc of 102 ± 10 kcal/mol (118 ± 10 kcal/mol), whereas its activated form, Factor Xaβ, possesses a TM of 55.0 ± 0.5 °C (55.0 ± 0.5 °C) and a ΔHc of 92.0 ± 5 kcal/mol (136 ± 10 kcal/mol). These studies indicate that, for many of these proteins, Ca2+ induces a conformational alteration to a more thermally stable form, which also requires the absorption of greater amounts of heat for thermal denaturation.  相似文献   

2.
CD spectra and melting curves were collected for a 28 base-pair DNA fragment in the form of a DNA dumbbell (linked on both ends by T4 single-strand loops) and the same DNA sequence in the linear form (without end loops). The central 16 base pairs (bp) of the 28-bp duplex region is the poly(pu) sequence: 5′-AGGAAGGAGGAAAGAG-3′. Mixtures of the dumbbell and linear DNAs with the 16-base single-strand sequence 5′-TCCTTCCTCCTTTCTC-3′ were also prepared and studied. At 22°C, CD measurements of the mixtures in 950 mM NaCl, 10 mM sodium acetate, 1 mM EDTA, pH 5.5, at a duplex concentration of 1.8 μM, provided evidence for triplex formation. Spectroscopic features of the triplexes formed with either a dumbbell or linear substrate were quite similar. Melting curves of the duplex molecules alone and in mixtures with the third strand were collected as a function of duplex concentration from 0.16 to 2.15 μM. Melting curves of the dumbbell alone and mixtures with the third strand were entirely independent of DNA concentration. In contrast, melting curves of the linear duplex alone or mixed with the third strand were concentration dependent. At identical duplex concentrations, the dumbbell alone melts ~20°C higher than the linear duplex. The curve of the linear duplex displayed a significant pretransition probably due to end fraying. On melting curves of mixtures of the dumbbell or linear duplex with the third strand, a low temperature transition with much lower relative hyperchromicity change (~ 5%) was observed. This transition was attributed to the melting of a new molecular species, e.g., the triplex formed between the duplex and single-strand DNA molecules. In the case of the dumbbell/single-strand mixture, these melting transitions of the triplex and the dumbbell were entirely resolvable. In contrast, the melting transitions of the linear duplex and the triplex overlapped, thereby preventing their clear distinction. To analyze the data, a three-state equilibrium model is presented. The analysis utilizes differences in relative absorbance vs temperature curves of dumbbells (or linear molecules) alone and in mixtures with the third strand. From the model analysis a straightforward derivation of fT(T), the fraction of triplex as a function of temperature, was obtained. Analysis of fT vs temperature curves, in effect melting curves of the triplexes, provided evaluation of thermodynamic parameters of the melting transition. For the triplex formed with the dumbbell substrate, the total transition enthalpy is ΔHT = 118.4 ± 12.8 kcal/mol (7.4 ± 0.8 kcal/mol per triplet unit) and the total transition entropy is ΔST = 344 ± 36.8 cal/K · mol (eu) (21.5 ± 2.3 eu per triple unit). The transition curves of the triplex formed with the linear duplex substrate displayed two distinct regions. A broad pretransition region from fT = 0 to 0.55 and a higher, sharper transition above fT = 0.55. The transition parameters derived from the lower temperature region of the curve are ΔHT = 44.8 ± 9.6 kcal/mol and ΔST = 112 ± 33.6 eu (or ΔH′ = 2.8 ± 0.6 kcal/mol and ΔS′ = 7.0 ± 2.1 eu per triplet). These values are probably too small to correspond to actual melting of the triplex but instead likely reveal effects of end fraying of the duplex substrate on triplex stability. Transition parameters of the upper transition are ΔHT = 128.0 ± 2.3 kcal/mol and ΔST = 379.2 ± 6.4 eu (ΔH′ = 8.0 ± 0.2 kcal/mol and ΔS′ = 23.7 ± 0.4 eu per triplet) in good agreement (within experimental error) with the transition parameters of the triplex formed with the dumbbell substrate. Supposing this upper transition reflects actual dissociation of the third strand from the linear duplex substrate this triplex is comparable in thermodynamic stability to the triplex formed with a dumbbell substrate. Even so, the biphasic melting character of the linear triplex obscures the whole analysis, casting doubt on its absolute reliability. Apparently triplexes formed with a dumbbell substrate offer technical advantages over triplexes formed from linear or hairpin duplex substrates for studies of DNA triplex stability. © 1993 John Wiley & Sons, Inc.  相似文献   

3.
The temperature dependence for the hydrolysis of both 4-methylumbelliferyl-α-l-fucoside and p-nitrophenyl-α-l-fucoside was determined for purified α-l-fucosidase (EC 3.2.1.51) from human placenta. The inhibition of the enzymatic reaction by l-fucose was also studied using the first of these two substrates at different temperatures. The thermodynamic parameters calculated from the pKm were for the 4-methylumbelliferyl-conjugate ΔF = ?6.6 kcal/mol, ΔH = ?8.5 kcal/mol, and ΔS = ?6.3 e.u. and for the p-nitrophenylconjugate ΔF = ?5.6 kcal/mol, ΔH = ?12.2 kcal/mol, and ΔS = ?21.1 e.u. The thermodynamic parameters for l-fucose were ΔH = ?12.4 kcal/mol and ΔS = ?20.1 e.u. The lower exothermicity and negative entropy calculated for the 4-methylumbelliferyl substrate compared to the thermodynamic parameters calculated for the p-nitrophenyl substrate and l-fucose suggest the existence of a secondary hydrophobic binding site for the 4-methylumbelliferyl moiety on the enzyme. The difference in the enthalpy for both substrates is also reflected in a difference in activation energy, being 15.8 kcal/mol for the 4-methylumbelliferyl substrate and 20.7 kcal/mol for the p-nitrophenyl substrate. From these results it may be concluded that altered kinetic properties of the enzyme could be the result of the binding of the “aglycone” moiety of the fluorogenic substrate to the enzyme.  相似文献   

4.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

5.
The present study deals with the kinetics and thermodynamics of the uptake of75Se-labeled SeO 3 2? from incubation media to lymphocytes cultivated from eight normal individuals (14–55 years of age, two females). The uptake of SeO 3 2? was evaluated on the assumption of pseudo-first-order kinetics with regard to a reacting cellular receptor pool. On the basis of the experimental observations, it was assumed that the suggested pool of receptor molecules-symbolically represented by “£H4”—reacts with SeO 3 2? in the hypothetical reaction: $$\pounds H_4 + SeO_3^{2 - } + 2H^ + \underset{{ - k_1 }}{\overset{{k_1 }}{\longleftrightarrow}}\pounds Se + 3H_2 O$$ The mean value of the change in standard free energy at 25°C was calculated to be ΔG o=?141.6±1.3 kJ/mol, while the corresponding mean value of the free energy of activation at 25°C was calculated to be ΔG 2+=?7.8±0.9 kJ/mol for the forward reaction. The calculated values of the corresponding individual changes in the respective standard enthalpies and entropies were mutually interdependent for all eight donors. ΔH o=?152+315ΔS o(kJ/mol) corresponding to the common value ΔG o??152 kJ/mol at 315°K. These mutual interdependencies are possibly the effect of variable conformational states (e.g., the macromolecular compactness) of the cellular receptor pools. This suggestion may furthermore be supported by the correlation traced between ΔH o vs the biological age in years of the donors: △H °?76.7?1.0 (age)kJ/mol (r = ?0.92) The calculated values of activation enthalpy ΔH 2+ kJ/mol and activation entropy ΔS 2+ (kJ/mol K) also mutually correlated linearly (r=0.998); the regression line was: △H 2+ = ?8.9 + 305△S2+ (kJ/mol) corresponding to the common value △H 2+ △ ?8.9 (kJ/mol) at 305°K Similarly the activation enthalpy ΔH 2+ vs the biological age in years correlated linearly: ΔH 2+=67.4?0.73(age) (kJ/mol) (r=?0.76) The range of ΔH 2+ studied was from 13.8 to 53.9 kJ/mol with a linearly corresponding range in ΔS 2+ from 73 to 205 J/mol K. The thermodynamic data reveal the selenite uptake during the hypothetical standard reaction to be exergonic and endothermic. Critical pH dependencies of the selenite uptake were explained.  相似文献   

6.
Chiral discrimination in the energetics of formation of diastereomeric pairs between iron (III)-complex ions bound to poly(L -glutamate) or poly (D -glutamate) and L -catecholamines was measured by differential microcalorimetry at 25°C. When the association of substrates by Fe–polypeptide systems was virtually complete, diastereomeric discrimination energies of a few hundred calories per mol were observed, while diastereomeric discrimination entropies were found to be negligibly small. These results are consistent with the finding that the overall stereoselectivity in the electron transfer reaction between the optically active catecholamines and enantiomeric iron(III) materials—corresponding to Δ(ΔG) values around 900 cal/mol—is largely controlled by transition state effects in the kinetics.  相似文献   

7.
The standard Gibbs free energy change of hydrolysis of α-d-ribose 1-phosphate has been measured at pH 7.0, ionic strength 0.1 m, and 25 °C by combining the corresponding values of the two following reactions: adenosine + H2O ág adenine + ribose (ΔG0′ = ?2.3 ± 0.1 kcal/mol), catalyzed by adenosine nucleosidase, and ribose 1-phosphate + adenine ág adenosine + PiG0′ = ?3.1 ± 0.1 kcal/mol), catalyzed by adenosine phosphorylase. The standard Gibbs free energy changes were calculated for both reactions from the equilibrium constant. A value of -5.4 ± 0.15 kcal/mol, comparable to that of other hemiacetal phosphoric esters, was obtained for the hydrolysis of ribose 1-phosphate.  相似文献   

8.
T W Sturgill 《Biopolymers》1978,17(7):1793-1810
A self-consistent thermodynamic characterization of the binding of ethidium to yeast phenylalanine-specific tRNA at 25°C, pH 7.0, in 11 nM MgCl2, 375 nM NaCl, and 25 mM sodium phosphate has been obtained. Two ethidium molecules bind per tRNA under these conditions. The stronger site has a dissociation constant equal to 1.9 ± 0.5 μM and ΔHdis°′ = 12 ± 1 Kcal/mol, and the weaker sites has a dissociation constant equal to 24 ± 9 μM and ΔHdis°′ = 8.9 ± 1.5 Kcal/mol. The average calorimetric ΔHdis°′ for the to sites 10.6 ± 0.4 kcal/mol. The thermodynamics of binding to the stranger sites are most probably the thermodynamics of interaction between A·U (6) and A·U (7), the unique site identified by Jones and Kearns. The binding is enthalpically driven and classical hydrophobic interactions do not appear to be important in the binding reaction.  相似文献   

9.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

10.
《Inorganica chimica acta》1988,149(2):259-264
The bis(N-alkylsalicylaldiminato)nickel(II) complexes Ni(R-sal)2 with R = CH(CH2OH)CH(OH)Ph (I), R = CH(CH3)CH(OH)Ph (II) and R = CH2CH2Ph (III; Ph = phenyl) were prepared and characterized. In the solid state I and II are paramagnetic (μ = 3.2 and 3.3 BM at 20 °C, respectively), whereas III is diamagnetic. It follows from the UV-Vis spectra that in acetone solution I is six-coordinate octahedral and III is four-coordinate planar, the spectrum of II showing characteristics of both modes of coordination. Vis spectrophotometry and stopped-flow spectrophotometry were applied to study the kinetics of ligand substitution in I–III by H2salen (= N,N′-disalicylidene-ethylenediamine) in the solvent acetone at different temperatures. The kinetics follow a second-order rate law, rate = k[H2-salen] [complex]. At 20 °C the sequence of rate constants is k(III):k(II):k(I) = 11 850:40.6:1. The activation parameters are ΔH(I) = 112, ΔH(II) = 40.7, ΔH(III) = 35.7 kJ mol−1 and ΔS(I) = 92, ΔS(II) = −103, ΔS(III) = −89 J K−1 mol−1. The enormous difference in rate between complexes I, II and III, which is less pronounced in methanol, is attributed to the existence of a fast equilibrium planar ⇌ octahedral, which is established in the case of I and II by intramolecular octahedral coordination through the hydroxyl groups present in the organic group R. An A-mechanism is suggested to control the substitution in the sense that the entering ligand attacks the four-coordinate planar complex, the octahedral complex being kinetically inert.  相似文献   

11.
A new series of 2,3-disubstituted quinazolin-4(3H)-one compounds including oxadiazole and furan rings was synthesized. Their inhibitory activities on urease were assessed in vitro. All newly synthesized compounds exhibited potent urease inhibitory activity in the range of IC50 = 1.55 ± 0.07–2.65 ± 0.08 µg/mL, when compared with the standard urease inhibitors such as thiourea (IC50 = 15.08 ± 0.71 µg/mL) and acetohydroxamic acid (IC50 = 21.05 ± 0.96 µg/mL). 2,3-Disubstituted quinazolin-4(3H)-one derivatives containing furan ring (3a-e) were found to be the most active inhibitors when compared with the compounds 2a-e bearing oxadiazole ring. Compound 3a, bearing 4-chloro group on phenyl ring, was found as the most effective inhibitor of urease with the IC50 value of 1.55 ± 0.11 µg/mL. The molecular docking studies of the newly synthesized compounds were performed to identify the probable binding modes in the active site of the Jack bean urease (JBU) enzymes.  相似文献   

12.
H J Hinz  W Haar  T Ackermann 《Biopolymers》1970,9(8):923-936
The enthalpies of the helix-coil transitions of the ordered polynucleotide systems of poly(inosinic acid)–poly(cytidylic acid) [poly(I + C)], (helical duplex), and of poly (inosinic acid) [poly(I + I + I)], (proposed secondary structure: a triple-stranded helical complex), were determined by using an adiabatic twin-vessel differential calorimeter. Measuring the temperature course of the heat capacity of the aqueous polymer solutions, the enthalpy values for the dissociation of the helical duplex poly (I + C) and the three-stranded helical complex poly(I + 1 + 1), respectively, were obtained by evaluating the additional heat capacity involved in the conformational change of the polynucleotide system in the transition range. The ΔH values of the helix-coil transition of poly (I + C) resulting from the analysis of the calorimetric measurements vary between the limits 6.5 ± 0.4 kcal/mole (I + C) and 8.4 ± 0.4 kcal/mole (I + C). depending on the variation of the cation concentration ranging from 0.063 mole cations kg H2O to 1.003 mole cations/kg H2O. The calorimetric investigation of an aqueous poly I solution (cation concentration 1.0 mole/kg H2O) yielded the enthalpy value ΔH = 1.9 ± 0.4 kcal/mole (I), a result which has been interpreted qualitatively following current models of inter- and intramolecular forces of biologically significant macromolecules. Additional information on the transition behavior of poly(I+ C)Was obtained by ultraviolet and infrared absorption measurements.  相似文献   

13.
Rate parameters have been obtained for the oxidation of cuprous stellacyanin by cobalt(III) ions of the form cis(N)-[CoN2O4]?, including cis(N)-[Co(NTA)(gly)]?, cis(N)-[Co(IDA)2]?, [Co(en)(ox)2]?(μ 0.5 M(phosphate), pH 7.0), and Co(EDTA)?(μ 0.1 M(NaCl), pH 7.2, 0.001 M phosphate). An excellent isokinetic correlation between the activation parameters ΔH and ΔS exists for the reactions of aminopolycarboxylatocobalt(III) ions with reduced stellacyanin (β = 300 ± 12 K; correlation coefficient = 0.995). It is concluded that enthalpy-entropy compensation in these reactions may be understood in terms of differing orientations preferred by the various oxidants in forming precursor complexes with the reduced blue protein. While ΔH and ΔS values for electron transfer from stellacyanin to cis(N)-[CoN2O4]? ions vary over ranges of 10.7 kcal/mol and 34 cal/mol-deg, respectively, room temperature rate constants are relatively constant (3.6–34.5 M?1 sec?1), as expected from Marcus theory for outer sphere electron transfer.  相似文献   

14.
In this study, a series of novel ethylenediamine compounds were obtained by structural modification of the lead compounds with thonzylamine, and using the principle of modifying by bioisostere formation and modification with alkyl groups. In vitro assay, the biological activities showed that the target compounds have good properties in inhibiting mast cell degranulation and releasing histamine and β-aminohexidase, such as the compounds 5c, 5g, 5k, 5l and 5o, especially of compound 5k to mast cell degranulation is IC50 = 0.0106 ± 0.001 μmol?L?1, histamine release was IC50 = 0.0192 ± 0.005 μmol?L?1 and β-hexosaminidase release was IC50 = 0.0455 ± 0.002 μmol?L?1 in vitro. At the same time, in vivo biological activities assay results showed that have a good Histamie induce bronchospasm effect with relatively long duration and good protective effect in vivo, among which the protective effect of compound 5k was 79.74 ± 0.30%, compounds 5c, 5g, 5k, 5l and 5o could inhibit the capillary permeability of increasing which were caused by histamine.  相似文献   

15.
《Endocrine practice》2019,25(3):226-229
Objective: Many youth with diabetes struggle to meet glycemic targets. The new ultralong duration of action of insulin degludec (iDeg) holds potential to ameliorate missed doses of basal insulin and improve glycemic control in youth with diabetes.Methods: A retrospective chart review was undertaken of youth age 13 to <24 years in our practice with type 1 diabetes (T1D) or type 2 diabetes (T2D) who had been switched from glargine or detemir to iDeg to evaluate the impact of this transition on glycemic control.Results: Glycated hemoglobin A1c (HbA1c) in youth with T1D (n = 82) remained stable during 6 months of treatment with iDeg (10.1 ± 2.11% &lsqb;87 ± 23 mmol/mol] at start of iDeg compared to 10.1 ± 2.12% &lsqb;87 ± 23 mmol/mol] at 6 months of treatment), whereas in youth with T2D (n = 16), HbA1c significantly declined from 10.6 ± 2.3% (92 ± 25 mmol/mol) to 8.3 ± 2.2% (67 ± 24 mmol/mol) (P = .0024).Conclusion: In youth switched to iDeg, which in our practice is commonly due to ineffectiveness of the patient's current regimen, the outcome differences we saw may be due to preserved beta-cell function in youth with T2D. It remains to be seen whether there are benefits of transition to iDeg in youth with T1D beyond glycemic outcomes, such as reduction in ketosis and episodes of diabetic ketoacidosis.Abbreviations: DKA = diabetic ketoacidosis; DPV = Diabetes-Patienten-Verlaufsdokumentation (German/Austrian Prospective Diabetes Follow-Up Registry); HbA1c = glycated hemoglobin A1c; iDeg = insulin degludec; T1D = type 1 diabetes; T2D = type 2 diabetes  相似文献   

16.
The thermodynamic parameters, ΔH′, ΔG′, and ΔS′, and the stoichiometry for the binding of the substrate 2′-deoxyuridine-5′-phosphate (dUMP) and the inhibitor 5-fluoro-2′-deoxyuridine-5′-phosphate (FdUMP) to Lactobacillus casei thymidylate synthetase (TSase) have been investigated using both direct calorimetric methods and gel filtration methods. The data obtained show that two ligand binding sites are available but that the binding of the second mole of dUMP is extremely weak. Binding of the first mole of dUMP can best be illustrated by dUMP + TSase + H+?(dUMP-TSase-H+). [1] The enthalpy, ΔH1′, for reaction [1] was measured directly on a flow modification of a Beckman Model 190B microcalorimeter. Experiments in two different buffers (I = 0.10 m) show that ΔH1′ = ?28 kJ mol?1 and that 0.87 mol of protons enters into the reaction. Analysis of thermal titrations for reaction [1] indicates a free energy change of ΔG1′ = ?30 kJ mol?1 (K1 = 1.7 × 105 m?1). From these parameters, ΔS1′ was calculated to be +5 J mol?1 degree?1, showing that the reaction is almost totally driven by enthalpy changes. Gel filtration experiments show that at very high substrate concentrations, binding to a second site can be observed. Gel filtration experiments performed at low ionic strength (I = 0.05 m) reveal a stronger binding, with ΔG1′ = ?35 kJ mol?1 (K1 = 1.2 × 106 m?1), suggesting that the forces driving the interaction are, in part, electrostatic. Addition of 2-mercaptoethanol (0.10 m) had the effect of slightly increasing the dUMP binding constant. Binding of FdUMP to TSase is best illustrated by 2FdUMP + TSase + nHH+?FdUMP2 ? TSase ? (H+)nH. [2] The enthalpy for this reaction, ΔH2, was also measured calorimetrically and found to be ?30 kJ mol?1 with nH = 1.24 at pH 7.4 Assuming two FdUMP binding sites per dimer as established by Galivan et al. [Biochemistry15, 356–362 (1976)] our calorimetric results indicate different binding energies for each site. Based on the binding data, a thermodynamic model is presented which serves to rationalize much of the confusing physical and chemical data characterizing thymidylate synthetase.  相似文献   

17.
Okumura H 《Proteins》2012,80(10):2397-2416
A multibaric‐multithermal molecular dynamics (MD) simulation of a 10‐residue protein, chignolin, was performed. All‐atom model with the Amber parm99SB force field was used for the protein and the TIP3P model was used for the explicit water molecules. This MD simulation covered wide ranges of temperature between 260 and 560 K and pressure between 0.1 and 600 MPa and sampled many conformations without getting trapped in local‐minimum free‐energy states. Folding events to the native β‐hairpin structure occurred five times and unfolding events were observed four times. As the temperature and/or pressure increases, fraction of folded chignolin decreases. The partial molar enthalpy change ΔH and partial molar volume change ΔV of unfolding were calculated as ΔH = 24.1 ± 4.9 kJ/mol and ΔV = ?5.6 ± 1.5 cm3/mol, respectively. These values agree well with recent experimental results. Illustrating typical local‐minimum free‐energy conformations, folding and unfolding pathways were revealed. When chignolin unfolds from the β‐hairpin structure, only the C terminus or both C and N termini open first. It may undergo an α‐helix or 310‐helix structure and finally unfolds to the extended structure. Difference of the mechanism between temperature denaturation and pressure denaturation is also discussed. Temperature denaturation is caused by making the protein transferred to a higher entropy state and making it move around more with larger space. The reason for pressure denaturation is that water molecules approach the hydrophobic residues, which are not well hydrated at the folded state, and some hydrophobic contacts are broken. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

18.
《Inorganica chimica acta》1987,128(2):169-173
The axial adduct formation of the iron(II) complex of 2,3,9,10-tetraphenyl-l,4,8,11-tetraaza-1,3,8,10-cyclotetradecatetraene (L) with imidazole in dimethyl sulfoxide has been investigated spectrophotometrically at various temperatures and pressures. In the presence of a large excess of imidazole the reaction with the two phases has been observed. The first faster reaction is the formation of the monoimidazole complex of FeL2+, and the second slower reaction corresponds to the formation of the bisimidazole complex. Activation parameters are as follows: for the first step with k1 (25.0°C) = (6.8 ±0.2)×105 mol−1 kg s−1, ΔH31 = 47.5 ± 4.9 kJ mol−1, ΔS31 = 26±16 J K−1 mol−1, and ΔV31 (30.0°C) = 27.2±1.5 cm3 mol−1; for the second step with k2 (25.0°C) = 26.8±0.8 mol−1 kg s−1, ΔH32 = 91.6± 0.8 kJ mol−1, ΔS32 = 90±3 J K−1 mol−1, and ΔV32 (35.0°C) = 21.8±0.9 cm3 mol−1. The large positive activation volumes strongly indicate a dissociative character of the activation process.  相似文献   

19.
The helix–coil transitions of poly(Nε-methyl, Nε-carbobenzoxy-L -lysine), poly(Nδ-methyl, Nδ-carbobenzoxy-L -ornithine), and poly(Nδ-ethyl, Nδ-carbobenzoxy-L -ornithine) in chloroform–dichloroacetic acid and their corresponding decarbobenzoxylated polypeptides in alkaline solutions were followed by optical rotation measurements. The introduction of a methyl or an ethyl group to the side chains of the carbobenzoxy derivatives of poly(L -lysine) and poly(L -ornithine) appeared to weaken the helical conformation at 25°C. The thermodynamic quantities of the three water-soluble polypeptides were calculated from the data on potentiometric titrations at several temperatures. For uncharged coil-to-helix transition, ΔH = ?370 cal/mol and ΔS = ?1.1 eu/mol for poly(Nε-methyl-L -lysine), and ΔH = ?540 cal/mol and ΔS = ?1.6 eu/mol for poly(Nδ-ethyl-L -ornithine) (all on molar residue basis). The absolute values of ΔH and ΔS dropped in the region of pH-induced transition and eventually both quantities became positive. The initiation factor σ was about 2 × 10?3, which was essentially independent of temperature. For poly(Nδ-methyl-L -ornithine) the coil-to-helix transition was not complete even when the polymer was uncharged at high pH.  相似文献   

20.
The observed equilibrium constants (Kobs) for the reactions of d-2-phosphoglycerate phosphatase, d-2-Phosphoglycerate3? + H2O → d-glycerate? + HPO42?; d-glycerate dehydrogenase (EC 1.1.1.29), d-Glycerate? + NAD+ → NADH + hydroxypyruvate? + H+; and l-serine:pyruvate aminotransferase (EC 2.6.1.51), Hydroxypyruvate? + l-H · alanine± → pyruvate? + l-H · serine±; have been determined, directly and indirectly, at 38 °C and under conditions of physiological ionic strength (0.25 m) and physiological ranges of pH and magnesium concentrations. From these observed constants and the acid dissociation and metal-binding constants of the substrates, an ionic equilibrium constant (K) also has been calculated for each reaction. The value of K for the d-2-phosphoglycerate phosphatase reaction is 4.00 × 103m [ΔG0 = ?21.4 kJ/mol (?5.12 kcal/mol)]([H20] = 1). Values of Kobs for this reaction at 38 °C, [K+] = 0.2 m, I = 0.25 M, and pH 7.0 include 3.39 × 103m (free [Mg2+] = 0), 3.23 × 103m (free [Mg2+] = 10?3m), and 2.32 × 103m (free [Mg2+] = 10?2m). The value of K for the d-glycerate dehydrogenase reaction has been determined to be 4.36 ± 0.13 × 10?13m (38 °C, I = 0.25 M) [ΔG0 = 73.6 kJ/mol (17.6 kcal/mol)]. This constant is relatively insensitive to free magnesium concentrations but is affected by changes in temperature [ΔH0 = 46.9 kJ/mol (11.2 kcal/mol)]. The value of K for the serine:pyruvate aminotransferase reaction is 5.41 ± 0.11 [ΔG0 = ?4.37 kJ/mol (?1.04 kcal/mol)] at 38 °C (I = 0.25 M) and shows a small temperature effect [ΔH0 = 16.3 kJ/ mol (3.9 kcal/mol)]. The constant showed no significant effect of ionic strength (0.06–1.0 m) and a response to the hydrogen ion concentration only above pH 8.5. The value of Kobs is 5.50 ± 0.11 at pH 7.0 (38 °C, [K+] = 0.2 m, [Mg2+] = 0, I = 0.25 M). The results have also allowed the value of K for the d-glycerate kinase reaction (EC 2.7.1.31), d-Glycerate? + ATP4? → d-2-phosphoglycerate3? + ADP3? + H+, to be calculated to be 32.5 m (38 °C, I = 0.25 M). Values for Kobs for this reaction under these conditions and at pH 7.0 include 236 (free [Mg2+] = 0) and 50.8 (free [Mg2+] = 10?3m).  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