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.
We study the unbiased folding/unfolding thermodynamics of the Trp‐cage miniprotein using detailed molecular dynamics simulations of an all‐atom model of the protein in explicit solvent using the Amberff99SB force field. Replica‐exchange molecular dynamics simulations are used to sample the protein ensembles over a broad range of temperatures covering the folded and unfolded states at two densities. The obtained ensembles are shown to reach equilibrium in the 1 μs/replica timescale. The total simulation time used in the calculations exceeds 100 μs. Ensemble averages of the fraction folded, pressure, and energy differences between the folded and unfolded states as a function of temperature are used to model the free energy of the folding transition, ΔG(P, T), over the whole region of temperatures and pressures sampled in the simulations. The ΔG(P, T) diagram describes an ellipse over the range of temperatures and pressures sampled, predicting that the system can undergo pressure‐induced unfolding and cold denaturation at low temperatures and high pressures, and unfolding at low pressures and high temperatures. The calculated free energy function exhibits remarkably good agreement with the experimental folding transition temperature (Tf = 321 K), free energy, and specific heat changes. However, changes in enthalpy and entropy are significantly different than the experimental values. We speculate that these differences may be due to the simplicity of the semiempirical force field used in the simulations and that more elaborate force fields may be required to describe appropriately the thermodynamics of proteins. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

2.
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   

3.
The unfolding at pH 8 of chicken cardiac aquometmyoglobin was examined as a function of temperature and concentration of guanidinium chloride using the two-state model. The isothermal unfolding data at 25°C were fitted to Tanford's transfer model and the binding model of Aune and Tanford. The estimates obtained for ΔGD) were virtually identical, viz., 8.3 ±0.3 kcal mol?1. The chicken metmyoglobin is thus some 5.3 kcal mol?1 less stable than that of sperm whale metmyoglobin. The unfolding parameters α and Δn were decreased 20% from those of mammalian myoglobins thus far examined, suggesting nonidentity of native conformations. The apparent enthalpy change on unfolding was dependent on both temperature and denaturant concentration. The decreases in the isothermal unfolding parameters from those of sperm whale are principally assigned to three of the 46 sequence changes.  相似文献   

4.
The Pressure Dependence of the Helix-Coil Transition Temperature (Tm) of Poly[d(G-C)] was studied as a function of sodium ion concentration in phosphate buffer. The molar volume change of the transition (ΔV) was calculated using the Clapeyron equation and calorimetrically determined enthalpies. The ΔV of the transition increased from +4.80 (±0.56) to +6.03 (±0.76) mL mol?1 as the sodium ion concentration changed from 0.052 to 1.0M. The van't Hoff enthalpy of the transition calculated from the half-width of the differentiated transition displayed negligible pressure dependence: however, the value of this parameter decreased with increasing sodium ion concentration, indicating a decrease in the size of the cooperative unit. The volume change of the transition exhibits the largest magnitude of any double-stranded DNA polymer measured using this technique. For poly[d(G-C)] the magnitude of the change in ΔV with sodium ion concentration (0.94 ± 0.05 mL mol?1) is approximately one-half that observed for either poly[d(A-T)] or poly (dA)·poly(dT). The ΔV values are interpreted as arising from changes in the hydration of the polymer due to the release of counterions and changes in the stacking of the bases of the coil form. As a consequence of solvent electrostriction, the release of counterions makes a net negative contribution to the total ΔV, implying that disruption of the slacking interactions contributes a positive volume change to the total ΔV. The larger magnitude of the ΔV compared with that of other double-stranded polymers may be due in part to the high helix-coil transition temperature of poly[d(G-C)], which will attenuate the contribution of electrostriction to the total volume change. The data in addition show that in the absence of other cellular components, the covalent structure of DNA is stabile under conditions of temperature and pressure more extreme than those experienced by any known organism. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
The reversible thermal unfolding of the archaeal histone-like protein Ssh10b from the extremophile Sulfolobus shibatae was studied using differential scanning calorimetry and circular dichroism spectroscopy. Analytical ultracentrifugation and gel filtration showed that Ssh10b is a stable dimer in the pH range 2.5–7.0. Thermal denaturation data fit into a two-state unfolding model, suggesting that the Ssh10 dimer unfolds as a single cooperative unit with a maximal melting temperature of 99.9 °C and an enthalpy change of 134 kcal/mol at pH 7.0. The heat capacity change upon unfolding determined from linear fits of the temperature dependence of ΔHcal is 2.55 kcal/(mol K). The low specific heat capacity change of 13 cal/(mol K residue) leads to a considerable flattening of the protein stability curve (ΔG (T)) and results in a maximal ΔG of only 9.5 kcal/mol at 320 K and a ΔG of only 6.0 kcal/mol at the optimal growth temperature of Sulfolobus.  相似文献   

6.
Abstract

This research is focussed on kinetic, thermodynamic and thermal inactivation of a novel thermostable recombinant α-amylase (Tp-AmyS) from Thermotoga petrophila. The amylase gene was cloned in pHIS-parallel1 expression vector and overexpressed in Escherichia coli. The steady-state kinetic parameters (Vmax, Km, kcat and kcat/Km) for the hydrolysis of amylose (1.39?mg/min, 0.57?mg, 148.6?s?1, 260.7), amylopectin (2.3?mg/min, 1.09?mg, 247.1?s?1, 226.7), soluble starch (2.67?mg/min, 2.98?mg, 284.2?s?1, 95.4) and raw starch (2.1?mg/min, 3.6?mg, 224.7?s?1, 61.9) were determined. The activation energy (Ea), free energy (ΔG), enthalpy (ΔH) and entropy of activation (ΔS) at 98?°C were 42.9?kJ mol?1, 74?kJ mol?1, 39.9?kJ mol?1 and ?92.3 J mol?1 K?1, respectively, for soluble starch hydrolysis. While ΔG of substrate binding (ΔGE-S) and ΔG of transition state binding (ΔGE-T) were 3.38 and ?14.1?kJ mol?1, respectively. Whereas, EaD, Gibbs free energy (ΔG*), increase in the enthalpy (ΔH*) and activation entropy (ΔS*) for activation of the unfolding of transition state were 108, 107, 105?kJ mol?1 and ?4.1 J mol?1 K?1. The thermodynamics of irreversible thermal inactivation of Tp-AmyS revealed that at high temperature the process involves the aggregation of the protein.  相似文献   

7.
Heating and cooling temperature jumps (T-jumps) were performed using a newly developed technique to trigger unfolding and refolding of wild-type ribonuclease A and a tryptophan-containing variant (Y115W). From the linear Arrhenius plots of the microscopic folding and unfolding rate constants, activation enthalpy (ΔH#), and activation entropy (ΔS#) were determined to characterize the kinetic transition states (TS) for the unfolding and refolding reactions. The single TS of the wild-type protein was split into three for the Y115W variant. Two of these transition states, TS1 and TS2, characterize a slow kinetic phase, and one, TS3, a fast phase. Heating T-jumps induced protein unfolding via TS2 and TS3; cooling T-jumps induced refolding via TS1 and TS3. The observed speed of the fast phase increased at lower temperature, due to a strongly negative ΔH# of the folding-rate constant. The results are consistent with a path-dependent protein folding/unfolding mechanism. TS1 and TS2 are likely to reflect X-Pro114 isomerization in the folded and unfolded protein, respectively, and TS3 the local conformational change of the β-hairpin comprising Trp115. A very fast protein folding/unfolding phase appears to precede both processes. The path dependence of the observed kinetics is suggestive of a rugged energy protein folding funnel.  相似文献   

8.
X‐ray crystallography is the most powerful method for determining three‐dimensional structures of proteins to (near‐)atomic resolution, but protein crystallization is a poorly explained and often intractable phenomenon. Differential Scanning Calorimetry was used to measure the thermodynamic parameters (ΔG, ΔH, ΔS) of temperature‐driven unfolding of two globular proteins, lysozyme, and ribonuclease A, in various salt solutions. The mixtures were categorized into those that were conducive to crystallization of the protein and those that were not. It was found that even fairly low salt concentrations had very large effects on thermodynamic parameters. High concentrations of salts conducive to crystallization stabilized the native folded forms of proteins, whereas high concentrations of salts that did not crystallize them tended to destabilize them. Considering the ΔH and TΔS contributions to the ΔG of unfolding separately, high concentrations of crystallizing salts were found to enthalpically stabilize and entropically destabilize the protein, and vice‐versa for the noncrystallizing salts. These observations suggest an explanation, in terms of protein stability and entropy of hydration, of why some salts are good crystallization agents for a given protein and others are not. This in turn provides theoretical insight into the process of protein crystallization, suggesting ways of predicting and controlling it. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 642–652, 2016.  相似文献   

9.
The interaction of the nonsteroidal anti‐inflammatory drug flurbiprofen (FBP) with human serum albumin (HSA) hardly influences the fluorescence of the protein's single tryptophan (Trp). Therefore, in addition to fluorescence, heavy atom‐induced room‐temperature phosphorescence is used to study the stereoselective binding of FBP enantiomers and their methyl esters to HSA. Maximal HSA phosphorescence intensities were obtained at a KI concentration of 0.2 M. The quenching of the Trp phosphorescence by FBP is mainly dynamic and based on Dexter energy transfer. The Stern–Volmer plots based on the phosphorescence lifetimes indicate that (R)‐FBP causes a stronger Trp quenching than (S)‐FBP. For the methyl esters of FBP, the opposite is observed: (S)‐(FBPMe) quenches more than (R)‐FBPMe. The Stern–Volmer plots of (R)‐FBP and (R)‐FBPMe are similar although their high‐affinity binding sites are different. The methylation of (S)‐FBP causes a large change in its effect on the HSA phosphorescence lifetime. Furthermore, the quenching constants of 3.0 × 107 M?1 s?1 of the R‐enantiomers and 2.5 × 107 M?1 s?1 for the S‐enantiomers are not influenced by the methylation and indicate a stereoselectivity in the accessibility of the HSA Trp to these drugs. Chirality 24:840–846, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
To understand how environmental changes have influenced forest productivity, stemwood biomass (B) dynamics were analyzed at 1267 permanent inventory plots, covering a combined 209 ha area of unmanaged temperate‐maritime forest in southwest British Columbia, Canada. Net stemwood production (ΔB) was derived from periodic remeasurements of B collected over a 40‐year measurement period (1959–1998) in stands ranging from 20 to 150 years old. Comparison between the integrated age response of net stemwood production, ΔB(A), and the age response of stemwood biomass, B(A), suggested a 58 ± 11% increase in ΔB between the first 40 years of the chronosequence period (1859–1898) and the measurement period. To estimate extrinsic forcing on ΔB, several different candidate models were developed to remove variation explained by intrinsic factors. All models exhibited temporal bias, with positive trends in (observed minus predicted) residual ΔB ranging between of 0.40 and 0.64% yr?1. Applying the same methods to stemwood growth (G) indicated residual increases ranging from 0.43 and 0.67% yr?1. Higher trend estimates corresponded with models that included site index (SI) as a predictor, which may reflect exaggeration of the age‐decline in SI tables. Choosing a model that excluded SI, suggested that ΔB increased by 0.40 ± 0.18% yr?1, while G increased by 0.43 ± 0.12% yr?1 over the measurement period. Residual G was significantly correlated with atmospheric carbon dioxide (CO2), temperature (T), and climate moisture index (CMI). However, models driven with climate and CO2, alone, could not simultaneously explain long‐term and measurement‐period trends without additional representation of indirect effects, perhaps reflecting compound interest on direct physiological responses to environmental change. Evidence of accelerating forest regrowth highlights the value of permanent inventories to detect and understand systematic changes in forest productivity caused by environmental change.  相似文献   

11.
The influence of proline cis-trans isomerization on the kinetics of lysozyme unfolding was examined carefully according to the theory of Hagerman and Baldwin [(1976) Biochemistry 15, 1462–1473]. As a result, the kinetics of lysozyme unfolding was found to follow the two-state transition model well. The temperature dependencies of kuf and kf over a wide temperature range showed that ΔC = 0 and ΔC = ?6.7 kJ K?1 mol?1 in solutions of different concentrations of GuHCl. The data observed in solutions containing other denaturants also supported the conclusion that ΔC is nearly equal to zero. The activation enthalpies of unfolding (ΔH) were observed at various concentrations of several kinds of denaturants. They were independent of species and concentrations of denaturants ΔH = 200 kJ mol?1). These facts indicate that the aspect of interaction between protein and different kinds of solvent molecules varies only slightly during the unfolding to the transition state, that is, the transition state is at compact as the native one. Therefore, it is also suggested that ΔH of 200 kJ mol?1 is primarily required for the disruption of long-range interactions among different structural domains through a subtle conformational change. We compared the effects of several kinds of denaturants on the unfolding rate. The addition of PrOH more remarkably increases the unfolding rate than do other hydrophilic denaturants. This is probably because PrOH molecules can penetrate into the hydrophobic core of lysozyme, but hydrophilic reagents cannot because of the compactness of the transition state.  相似文献   

12.
We have measured the thermodynamic parameters of the slow-fast tail-fiber reorientation transition on T2L bateriophage. Proportions of the virus in each form were determined from peak-height measurements in sedimention-velocity runs and from average diffusion coefficients obtained by quasielastic laser light scattering. Computer simulation of sedimentation confirmed that there were no undetected intermediates in the transition, which was analyzed as a two-state process. Van't Hoff-type plots of the apparent equilibrium constant and of the pH midpoint of the transition as function of reciprocal temperature led to the following estimates of the thermodynamic parameters for the transition at pH 6.0 and 20°C: ΔH° = ?139 ± 18Kcal mol?1, ΔS° = ?247 ± 46 cal K?1 mol?1, and ΔG° = ?66 ± 22 kcal mol?1. Per mole of protons taken up in the transition, the analogous quantities were ?15.9 ± 1.7 kcal mol?1, ?26.3 ± 2.2 cal K?1 mol?1, and ?8.22 ± 1.8 kcal mol?1. The net number of protons taken up was about 8.5 ± 1.5. The large values of the thermodynamic functions are consistent with a highly cooperative reaction and with multiple interactions between the fibres and the remainder of the phage. The negative entropy of the transition is probably due to immobilization of the fibres.  相似文献   

13.
Lactic dehydrogenase from pig heart can be reversibly dissociated at hydrostatic pressures above 1000 bar. The breakdown of the native quaternary structure occurs at lower pressures compared to the isoenzyme from pig, skeletal muscle. As shown by hybridization experiments of the two isoenzymes the final product of dissociation is the homogeneous monomer. Fluorescence emission spectra of the monomeric enzyme at elevated pressure are characterized by a decrease in fluorescence intensity without any red shift, indicating that no significant unfolding occurs upon high-pressure dissociation. The spectral changes are comparable to those observed after acid dissociation. The amount and rate of deactivation depend on pressure and on the conditions of the solvent. The presence of various anions (Cl?, SO2?4. HPO42?) has no effect on the stability of ihe enzyme towards pressure. High-pressure denaturation (as monitored by intrinsic protein fluorescence), and deactivalion (measured immediately after decompression) run parallel; the pressure dependence of their first-order rate constants is characterized by an activation volume ΔVDc = ?140 = 10 cm3/mol. As taken from the yield of reconstitution, dissociation, denaturation and deactivation are found to be fully reversible provided the pressure does not exceed a limiting value (p = 1000 bar in Tris. pH 7.6: 24 h incubation at 20°C). After extended incubation beyond the limiting, pressure of 1000 bar. “irreversible high-pressure denaturation” occurs which is accompanied bv partial aggregation after decompression. The coenzyme, NAD+ stabilizes the native tetramer shifting the dissociation equilibrium to higher pressures. The overall dissociation-association reaction can be quantitatively described by a consecutive dissociation/unfolding mechanism N?4 M'?4 M (where N is the native tetramer. and M' and M two different conformations of the monomer). The reaction volume of the dissociation reaction N?4 M' is found to be ΔVDiss = ?360 = 30 cm3/mol: as indicated by the pressure dependence of the yield of reconstitution, the reaction volume of the equilibrium M'?MXXX is also negative.  相似文献   

14.
High‐density lipoproteins (HDLs) are complexes of lipids and proteins (termed apolipoproteins) that remove cell cholesterol and protect from atherosclerosis. Apolipoproteins contain amphipathic α‐helices that have high content (≥1/3) and distinct distribution of charged and apolar residues, adopt molten globule‐like conformations in solution, and bind to lipid surfaces. We report the first pressure perturbation calorimetry (PPC) study of apolipoproteins. In solution, the main HDL protein, apoA‐I, shows relatively large volume contraction, ΔVunf = ?0.33%, and an apparent reduction in thermal expansivity upon unfolding, Δαunf ≤ 0, which has not been observed in other proteins. We propose that these values are dominated by increased charged residue hydration upon α‐helical unfolding, which may result from disruption of multiple salt bridges. At 5°C, apoA‐I shows large thermal expansion coefficient, α(5°) = 15·10?4 K?1, that rapidly declines upon heating from 5 to 40°C, α(40°) ? α(5°) = ?4·10?4 K?1; apolipoprotein C‐I shows similar values of α(5°) and α(40°). These values are larger than in globular proteins. They indicate dominant effect of charged residue hydration, which may modulate functional apolipoprotein interactions with a broad range of their protein and lipid ligands. The first PPC analysis of a protein–lipid complex is reported, which focuses on the chain melting transition in model HDL containing apoA‐I or apoC‐I, dimyristoyl phosphatidylcholine, and 0–20% cholesterol. The results may provide new insights into volumetric properties of HDL that modulate metabolic lipoprotein remodeling during cholesterol transport. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

15.
K Sasahara  M Sakurai  K Nitta 《Proteins》2001,44(3):180-187
The influence of hydrostatic pressure (< or =100 MPa) on denaturant-induced unfolding of hen egg white lysozyme was investigated by means of ultraviolet spectroscopy at various temperatures. Assuming a two-state transition model, the dependence of Gibbs free-energy change of unfolding on the denaturant concentration was calculated. Under applied hydrostatic pressure, these data were interpreted as suggesting that a two-state model is not applicable in a restricted temperature range; the dominant effect of hydrostatic pressure is to affect the cooperativity in protein unfolding due to a chemical equilibrium shift in the direction of the reduction in the system volume. The deviation from the two-state transition model appears to be rationalized by assuming that applied pressure induces an intermediate conformation between the native and unfolded states of the protein. The implication of the thermodynamic stability of protein under pressure was discussed.  相似文献   

16.
The antidepressant drug tetramezine [1,2‐bis‐(3,3‐dimethyldiaziridin‐1‐yl)ethane] consists of two bridged diaziridine moieties with four stereogenic nitrogen centers, which are stereolabile and, therefore, are prone to interconversion. The adjacent substituents at the nitrogen atoms of the diaziridines moieties exist only in an antiperiplanar conformation, which results in a coupled interconversion. Therefore, three stereoisomers exist (meso form and two enantiomeric forms), which epimerize when the diaziridine moieties are regarded as stereogenic units due to the coupled interconversion. Here, we have investigated the epimerization between the meso and enantiomeric forms by dynamic gas chromatography. Temperature‐dependent measurements were performed, and reaction rate constants were determined using the unified equation of chromatography implemented in the software DCXplorer. The activation barriers of the epimerization were found to be ΔG = 100.7 kJ mol?1 at 25°C and ΔG = 104.5 kJ mol?1 at 37°C, respectively. The activation enthalpy and entropy were determined to be ΔH = 70.3 ± 0.4 kJ mol?1 and ΔS = ?102 ± 2 J mol?1 K?1. Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

17.
Spectroscopic and calorimetric melting studies of 28 DNA hairpins were performed. These hairpins form by intramolecular folding of 16 base self‐complementary DNA oligomer sequences. Sequence design dictated that the hairpin structures have a six base pair duplex linked by a four base loop and that the first five base pairs in the stem are the same in every molecule. Only loop sequence and identity of the duplex base pair closing the loop vary for the set of hairpins. For these DNA samples, melting studies were carried out to investigate effects of the variables on hairpin stability. Stability of the 28 oligomers was ascertained from their temperature‐induced melting transitions in buffered 115 mM Na+ solvent, monitored by ultraviolet absorbance and differential scanning calorimetry (DSC). Experiments revealed the melting temperatures of these molecules range from 32.4 to 60.5°C and are concentration independent over strand concentrations of 0.5 to 260 μM; thus, as expected for hairpins, the melting transitions are apparently unimolecular. Model independent thermodynamic transition parameters, ΔHcal, ΔScal, and ΔGcal, were determined from DSC measurements. Model dependent transition parameters, ΔHvH, ΔSvH, and ΔGvH were estimated from a van't Hoff (two‐state) analysis of optical melting transitions. Results of these studies reveal a significant sequence dependence to DNA hairpin stability. Thermodynamic parameters evaluated by either procedure reveal the transition enthalpy, ΔHcalHvH) can differ by as much as 20 kcal/mol depending on sequence. Similarly, values of the transition entropy ΔScalSvH) can differ by as much as 60 cal/Kmol (eu) for different molecules. Differences in free energies ΔGcalGvH) are as large as 4 kcal/mol for hairpins with different sequences. Comparisons between the model independent calorimetric values and the thermodynamic parameters evaluated assuming a two‐state model reveal that 10 of the 28 hairpins display non‐two‐state melting behavior. The database of sequence‐dependent melting free energies obtained for the hairpins was employed to extract a set of n‐n (nearest‐neighbor) sequence dependent loop parameters that were able to reproduce the input data within error (with only two exceptions). Surprisingly, this suggests that the thermodynamic stability of the DNA hairpins can in large part be reasonably represented in terms of sums of appropriate nearest‐neighbor loop sequence parameters. © 1999 John Wiley & Sons, Inc. Biopoly 50: 425–442, 1999  相似文献   

18.
N‐alkylated trans‐diaziridines are an intriguing class of compounds with two stereogenic nitrogen atoms which easily interconvert. In the course of our investigations of the nature of the interconversion process via nitrogen inversion or electrocyclic ring opening ring closure, we synthesized and characterized the three constitutionally isomeric diaziridines 1,2‐di‐n‐propyldiaziridine 1 , 1‐isopropyl‐2‐n‐propyldiaziridine 2 , and 1,2‐diisopropyldiaziridine 3 to study the influence of the substituents on the interconversion barriers. Enantiomer separation was achieved by enantioselective gas chromatography on the chiral stationary phase Chirasil‐β‐Dex with high separation factors α (1‐isopropyl‐2‐n‐propyldiaziridine: 1.18; 1, 2‐diisopropyldiaziridine: 1.24; 100°C 50 kPa He) for the isopropyl substituted diaziridines. These compounds showed pronounced plateau formation between 100 and 150°C, and peak coalescence at elevated temperatures. The enantiomerization barriers ΔG? and activation parameters ΔH? and ΔS? were determined by enantioselective dynamic gas chromatography (DGC) and direct evaluation of the elution profiles using the unified equation implemented in the software DCXplorer. Interestingly, 1‐isopropyl‐2‐n‐propyldiaziridine and 1,2‐diisopropyldiaziridine exhibit similar high interconversion barriers ΔG? (100°C) of 128.3 ± 0.4 kJ mol?1 and 129.8 ± 0.4 kJ mol?1, respectively, which indicates that two sterically demanding substituents do not substantially increase the barrier as expected for a distinct nitrogen inversion process. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
Anoxybacillus beppuensis TSSC-1 (GenBank Number, EU710556), a thermophilic bacterium isolated from a hot spring reservoir, was found to optimally secrete a monomeric α-amylase at 55 °C and pH 7. The enzyme was purified to homogeneity by a single-step purification on phenyl sepharose 6FF, achieving a 58% yield, 10,000 U/mg specific activity and 19.5 fold purification. The molecular weight, Km and Vmax were 43 kD, 0.5 mg ml?1 and 3571.42 μmol ml?1 m?1, respectively. The enzymatic catalysis of soluble starch was optimum at 80 °C and pH 7. The thermodynamic parameters, Kd, t1/2, ΔH*, ΔS*, E and ΔG*, were consistent. The very compact structure of the enzyme and the transitional enzyme–substrate complex resisted denaturation at extreme temperatures and alkaline pH. The Kd and t1/2 measurements were consistent with the high thermostability and pH tolerance observed. The structural stability of the enzyme was also reflected by the values of ΔH*, ΔS*, E and ΔG*. While the enzyme did not exhibit metal ion dependency, it was resistant to chemical denaturation. The broad thermo- and pH-tolerance of this enzyme suggests potential commercial opportunities.  相似文献   

20.
We performed complementary inelastic neutron scattering (INS) experiments and molecular dynamics (MD) simulations to study the influence of pressure on the low‐frequency vibrational modes of lysozyme in aqueous solution in the 1 atm–6 kbar range. Increasing pressure induces a high‐frequency shift of the low‐frequency part (<10 meV = 80 cm?1) of the vibrational density of states (VDOS), g(ω), of both lysozyme and water that reveals a stiffening of the interactions ascribed to the reduction of the protein and water volumes. Accordingly, high pressures increase the curvature of the free energy profiles of the protein quasiharmonic vibrational modes. Furthermore, the nonlinear influence of pressure on the g(ω) of lysozyme indicates a change of protein dynamics that reflects the nonlinear pressure dependence of the protein compressibility. An analogous dynamical change is observed for water and stems from the distortion of its tetrahedral structure under pressure. Moreover, our study reveals that the structural, dynamical, and vibrational properties of the hydration water of lysozyme are less sensitive to pressure than those of bulk water, thereby evidencing the strong influence of the protein surface on hydration water. Proteins 2013. © 2012 Wiley Periodicals, Inc.  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