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 conformational behavior of DNA minihairpin loops is sensitive to the directionality of the base pair that closes the loop. Especially tailored circular dumbbells, consisting of a stem of three Watson–Crick base pairs capped on each side with a minihairpin loop, serve as excellent model compounds by means of which deeper insight is gained into the relative stability and melting properties of hairpin loops that differ only in directionality of the closing pair: C-G vs G-C. For this reason the thermodynamic properties of the circular DNA decamers 5′-d〈pCGC-TT-GCG-TT〉-3′( I ) and reference compounds 5′-d〈pGGC-TT-GCC-TT≤-3′( II ) and 5′-d(GCG-TC-CGC)-3′( III ) are studied by means of nmr spectroscopy. Molecules I and II adopt dumbbell structures closed on both sides by a two-membered hairpin hop. At low temperature I consists of a mixture of two slowly exchanging forms, denoted L2L2 and L2L4 . The low-temperature L2L2 form is the fully intact minihairpin structure with three Watson–Crick C-G base pairs. The high-temperature form, L2L4 ,contains a partially disrupted closing G-C base pair in the 5′-GTTC-3′ loop, with the cytosine base placed in a syn orientation. The opposite 5′-CTTG-3′ loop remains stable. A study of the noncircular hairpin structure III shows similar conformational behavior for the 5′-GTTC-3′ loop as found in I a syn orientation for C(6) and two slowly exchanging imino proton signals for G(3). The melting point Tm of II was estimated to lie above 365 K. The Tm value of the duplex stem and the 5′-CTTG-3′ loop of the L2L4 form ofIis 352 ± 2 K. The ΔH° is calculated as ?89 ± 10 kJ/mol. The Tm value determined for the individual residues of the 5′-GTTC-3′ loop lies 4°–11° lower. The enthalpy ΔH° of melting the thymine residues in the 5′-GTTC-3′ loop is calculated to be -61± 7 kJ/mol. Thermodynamic data of the equilibrium between the slowly exchanging two- and four-membered loop conformers of I reveal an upper limit for ΔH° of +30 kJ/mol in going from a two-memberedto a four-membered loop, in agreement with the enthalpy difference of +28 k.j/mol between the two loops at the Tm midpoint. For hairpin III the upper limit for ΔH° going from a two-membered to a four-membered loop amounts to ±21 kJ/mol. The mutual exchange rate between the L2 and L4 form in III is estimated as 13.6 s?1. Our results clearly suggest that small four-way DNA junctions(model for immobilized Holliday junctions) can be designed that consist of a single DNA strandthat features -CTTG-caps on three of the four arms of the junction. © 1995 John Wiley & Sons, Inc.  相似文献   

2.
3.
Plants offer the potential for selective removal and sequestration of toxic heavy metals from contaminated soil. Phytoextraction of metal ions involve their transport through the plant’s root system and into its shoots and leaves. This study investigates the thermodynamics of Eu(III) ion chemical interactions with Datura innoxia plant root materials under solution conditions of pH 4.0 and 5.0. Both changes in enthalpies (?H) and entropies (?S) of metal binding were elucidated from isotherms collected under varied temperature conditions using regularized regression data analysis and conditional affinity spectra. ?H values for binding to root materials at pH 4.0 and 5.0 were each calculated to be +30 kJ/mol. Values of ΔS for these same materials were found to be +170 and +153 J/mol K for solution conditions of pH 4.0 and 5.0, respectively. These results suggest binding to the root material to be entropically driven (?S° > 0 and ΔH > 0) through possible displacement of waters of solvation.  相似文献   

4.
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   

5.
The most effective way of enzymatic synthesis of biodiesel is through lipase-catalyzed transesterification, while its performance and economic feasibility should still be improved. In this study, lipase produced by an isolated Burkholderia sp. was immobilized on microsize Celite materials functionally modified with long alkyl groups. The specific activity of the immobilized lipase was 1,154 U/g. The methanolysis of olive oil catalyzed by the immobilized lipase obeyed Ping Pong Bi Bi model with an estimated V max, K m,TG, K m,M and K i,M value of 0.61 mol/(L min), 7.93 mol/L, 1.01 mol/L, and 0.24 mol/L, respectively. The activation energy of the enzymatic reaction is estimated as 15.51 kJ/mol. The immobilized lipase exhibits high thermal stability with thermal deactivation energy of 83 kJ/mol and a long half-life. The enthalpy, Gibb’s free energy, and entropy of the immobilized lipase were in the range of 80.02–80.35 kJ/mol, 88.35–90.13 kJ/mol, and ?28.22 to ?25.11 J/(mol K), respectively.  相似文献   

6.
Phenolic compounds have increasingly attracted global concerns in recent years due to their strong bioaccumulation and potential toxicity. In this study, the adsorption of 2, 4-Dichlorophenol (2, 4-DCP), 2, 4-Dinitrophenol (2, 4-DNP) and 2, 4-Dimethyphenol (2, 4-DMP) on sediment at different temperatures was studied. Adsorption isotherms fitted well to the Freundlich model and the adsorption capacity was increased when the temperature increased from 5°C to 25°C. It was found that the adsorption process could be modeled well using pseudo-second-order kinetic equation. The thermodynamic parameter ΔG° for 2, 4-DCP, 2, 4-DNP and 2, 4-DMP varied between ?8.82 and ?13.68, ?4.9 and ?8.05, and ?7.52 and ?10.55 kJ/mol, respectively, with ΔH° (kJ/mol) and ΔS°(J/(mol·K)) measuring 55.397 and 0.2263, 40.121 and 0.1585, and 38.012 and 0.16, respectively. The calculated thermodynamic parameters suggested that adsorption of the three selected phenolic compounds was a spontaneous (ΔG°< 0), endothermic (ΔH°> 0), and entropy-driven reaction (ΔS° > 0). The thermodynamic data also suggested that the three selected phenolic compounds to sediment were closed to chemisorption because ΔH° was around 40.  相似文献   

7.
Calorimetric studies of the reduction of free oxygen in solution by sodium dithionite are in agreement with a stoichiometry of 2 moles Na2S2O4 per mole of oxygen. The reaction is biphasic with ΔHt - 118±7 kcal mol?1 (?494 ± 29 kJ mol?1). The initial phase of the reaction proceeds with an enthalpy change of ca ?20 kcal (?84 kJ) and occurs when 0.5 moles of dithionite have been added per mole dioxygen present. This could be interpreted as the enthalpy change for the addition of a single electron to form the superoxide anion. Further reduction of the oxygen to water by one or more additional steps is accompanied by an enthalpy change of ca ?100 kcal (?418. 5 kJ). Neither of these reductive phases is consistent with the formation of hydrogen peroxide as an intermediate. The reduction of hydrogen peroxide by dithionite in 0.1 M phosphate buffer, pH 7.15, is a much slower process and with an enthalpy change of ca ? 74 kcal mol?1 (?314 kJ mol?1). Dissociation of oxyhemoglobin induced by the reduction of free oxygen tension with dithionite also shows a stoichiometry of 2 moles dithionite per mole oxygen present and an enthalpy change of ca. ?101 ±9 kcal mol?1 (?423± 38 kJ mol?1). The difference in the observed enthalpies (reduction of dioxygen vs. oxyhemoglobin) has been attributed to the dissociation of oxyhemoglobin, which is 17 kcal mol?1 (71 kJ mol?1).  相似文献   

8.
The feasibility for the removal of Acid Blue25 (AB25) by Bengal gram fruit shell (BGFS), an agricultural by-product, has been investigated as an alternative for high-cost adsorbents. The impact of various experimental parameters such as dose, different dye concentration, solution pH, and temperature on the removal of Acid Blue25 (AB25) has been studied under the batch mode of operation. pH is a significant impact on the sorption of AB25 onto BGFS. The maximum removal of AB25 was achieved at a pH of 2 (83.84%). The optimum dose of biosorbent was selected as 200 mg for the removal of AB25 onto BGFS. Kinetic studies reveal that equilibrium reached within 180 minutes. Biosorption kinetics has been described by Lagergren equation and biosorption isotherms by classical Langmuir and Freundlich models. Equilibrium data were found to fit well with the Langmuir and Freundlich models, and the maximum monolayer biosorption capacity was 29.41 mg g?1 of AB25 onto BGFS. The kinetic studies indicated that the pseudo-second-order (PSO) model fitted the experimental data well. In addition, thermodynamic parameters have been calculated. The biosorption process was spontaneous and exothermic in nature with negative values of ΔG° (?1.6031 to ?0.1089 kJ mol?1) and ΔH° (?16.7920 kJ mol?1). The negative ΔG° indicates the feasibility of physical biosorption process. The results indicate that BGFS could be used as an eco-friendly and cost-effective biosorbent for the removal of AB25 from aqueous solution.  相似文献   

9.
In this study, the acid-soluble collagen (ASC), extracted from the fish scales of the Caspian white fish (Rutilus Firisikutum) was studied. The thermo-gravimetric analysis (TGA) showed the maximum demineralization accomplished after 48 h of EDTA treatment. SDS-PAGE and FT-IR spectroscopy confirmed that extracted ASC was mainly type I collagen. FE-SEM images confirmed the porous and filamentary structure. The denaturation temperature (Td) of ASC was 19 °C, and the transition heat achieved 9.6 J/g. Collagen self-assembly exhibit important potential because for biomedical applications and green technologies. Various inter- and intra-molecular no-covalent interactions such as hydrogen bonding, hydrophobic, electrostatic and Van der Waals interactions influence the formation of self-assembled collagen. Therefore, critical factors as concentration of ASC, temperature, pH, and ionic strength play crucial role in function integration and structural modulation. The impacts of those external triggers on the kinetic self-assembly of ASC demonstrated a two-phase kinetic process, a sigmoidal plot. ACS showed pronounced self-assembly behavior when temperature and concentration reach above 14 °C and 0.125 mg/ml, higher concentration and/or temperature could stimulate the ASC self-assembly. The optimum pH value for ASC self-assembly was pH = 7. The effect of ionic strength on ASC self-assembly showed the turbidity increases significantly in 131.2 mM salt concentration. The process of self-assembly is mainly driven by thermodynamics. The thermodynamic study of collagen self-assembly illustrated that the activation energy, Ea = 44.3 kJ/mol, the frequency factor, A = 117 × 105 s?1, the enthalpy transition, ΔH? = 42.98 kJ/mol, and the entropy transition, ΔS? = ?0.12 kJ/mol.K, respectively. These findings show that kinetics factors not only influence the self-assembly structure of ASC but also regulate the activation complex structure in the transition state.  相似文献   

10.
All life requires energy to drive metabolic reactions such as growth and cell maintenance; therefore, fluctuations in energy availability can alter microbial activity. There is a gap in our knowledge concerning how energy availability affects the growth of extreme chemolithoautotrophs. Toward this end, we investigated the growth of thermoacidophile Acidianus ambivalens during sulfur oxidation under aerobic to microaerophilic conditions. Calorimetry was used to measure enthalpy (ΔHinc) of microbial activity, and chemical changes in growth media were measured to calculate Gibbs energy change (ΔGinc) during incubation. In all experiments, Gibbs energy was primarily dissipated through the release of heat, which suggests enthalpy‐driven growth. In microaerophilic conditions, growth was significantly more efficient in terms of biomass yield (defined as C‐mol biomass per mole sulfur consumed) and resulted in lower ΔGinc and ΔHinc. ΔGinc in oxygen‐limited (OL) and oxygen‐ and CO2‐limited (OCL) microaerophilic growth conditions resulted in averages of ?1.44 × 103 kJ/C‐mol and ?7.56 × 102 kJ/C‐mol, respectively, and average ΔHinc values of ?1.11 × 105 kJ/C‐mol and ?4.43 × 104 kJ/C‐mol, respectively. High‐oxygen experiments resulted in lower biomass yield values, an increase in ΔGinc to ?1.71 × 104 kJ/C‐mol, and more exothermic ΔHinc values of ?4.71 × 105 kJ/C‐mol. The observed inefficiency in high‐oxygen conditions may suggest larger maintenance energy demands due to oxidative stresses and a preference for growth in microaerophilic environments.  相似文献   

11.
The rhodium-catalysed hydroacylation of alkene is one of the most useful C–H bond activation processes. The C–C bond-forming reactions via C–H bond activation have extensively been the focus of study in the fields of organic and organometallic chemistry. In this work, density functional theory has been used to study Rh(I)-catalysed hydroacylation and hydrogenation of ethene with formic acid. All the intermediates and the transition states were optimised completely at the B3LYP/6-311++G(d,p) level (LANL2DZ(d) for Rh, P). Calculation results confirm that Rh(I)-catalysed hydroacylation of ethene is exothermic and the released Gibbs free energy is ? 60.39 kJ/mol. Rh(I)-catalysed hydrogenation of ethene is also exothermic and the released Gibbs free energy is ? 150.97 kJ/mol. Rh(I)-catalysed hydroacylation of ethene is the dominant reaction mode for Rh(I)-catalysed hydroacylation and hydrogenation of ethene with formic acid. In Rh(I)-catalysed hydroacylation of ethene, the H-transfer reaction is prior to the C–C bond-forming reaction. Therefore, the reaction mode ‘a’ (i.e. ca → M1 → TS1 → M2 → TS2a → M3a → TS3a → M4 → P1) is the dominant reaction pathway for Rh(I)-catalysed hydroacylation and hydrogenation of ethene. The theoretically predicted dominant product is propane acid.  相似文献   

12.
The establishment of renewable biofuel and chemical production is desirable because of global warming and the exhaustion of petroleum reserves. Sebacic acid (decanedioic acid), the material of 6,10-nylon, is produced from ricinoleic acid, a carbon-neutral material, but the process is not eco-friendly because of its energy requirements. Laccase-catalyzing oxidative cleavage of fatty acid was applied to the production of dicarboxylic acids using hydroxy and oxo fatty acids involved in the saturation metabolism of unsaturated fatty acids in Lactobacillus plantarum as substrates. Hydroxy or oxo fatty acids with a functional group near the carbon–carbon double bond were cleaved at the carbon–carbon double bond, hydroxy group, or carbonyl group by laccase and transformed into dicarboxylic acids. After 8 h, 0.58 mM of sebacic acid was produced from 1.6 mM of 10-oxo-cis-12,cis-15-octadecadienoic acid (αKetoA) with a conversion rate of 35% (mol/mol). This laccase-catalyzed enzymatic process is a promising method to produce dicarboxylic acids from biomass-derived fatty acids.  相似文献   

13.
Direct enantiomer separation of hypericin, pseudohypericin, and protohypericin was accomplished by high‐performance liquid chromatography (HPLC) using immobilized polysaccharide‐type chiral stationary phases (CSPs). Enantioselectivities up to 1.30 were obtained in the polar‐organic elution mode whereby for hypericin and pseudohypericin Chiralpak IC [chiral selector being cellulose tris(3,5‐dichlorophenylcarbamate)] and for protohypericin Chiralpak IA (chiral selector being the 3,5‐dimethylphenylcarbamate of amylose) gave favorable results. Enantiomers were distinguished by on‐line electronic circular dichroism detection. Optimized enantioselective chromatographic conditions were the basis for determining stereodynamic parameters of the enantiomer interconversion process of hypericin and pseudohypericin. Rate constants delivered by computational simulation of dynamic HPLC elution profiles (stochastic model, consideration of peak tailing) were used to calculate averaged enantiomerization barriers (ΔG) of 97.6–99.6 kJ/mol for both compounds (investigated temperature range 25–45°C). Complementary variable temperature off‐column (i.e., in solution) racemization experiments delivered ΔG = 97.1–98.0 kJ/mol (27–45°C) for hypericin and ΔG = 98.9–101.4 kJ/mol (25–55°C) for pseudohypericin. An activation enthalpy of ΔH# = 86.0 kJ/mol and an activation entropy of ΔS# = ?37.7 J/(K mol) were calculated from hypericin racemization kinetics in solution, whereas for pseudohypericin these figures amounted to 74.1 kJ/mol and ?82.6 J/(K mol), respectively. Although the natural phenanthroperylene quinone pigments hypericin and pseudohypericin as well as their biological precursor protohypericin are chiral and can be separated by enantioselective HPLC low enantiomerization barriers seem to prevent the occurrence of an excess of one enantiomer under typical physiological conditions—at least as long as stereoselective intermolecular interactions with other chiral entities are absent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
Although the dehydration of α-D-glucose monohydrate is an important aspect of several industrial processes, there is uncertainty with regard to the applicable rate law and other factors that affect dehydration. Therefore, the dehydration of three glucose monohydrate samples has been studied using isothermal gravimetric analysis. Dehydration follows a one-dimensional contraction (R1) rate law for the majority of kinetic runs, and an activation energy of 65.0±3.9 kJ mol(-1) results when the rate constants are fitted to the Arrhenius equation. Fitting the rate constants to the Eyring equation results in values of 62.1±3.7 kJ mol(-1) and -77.8±4.7 J mol(-1)K(-1) for ΔH(?) and ΔS(?), respectively. The impedance effect on the loss of water vapor has also been investigated to determine the values for activation energy, enthalpy, and entropy for diffusion of water. The results obtained for the activation parameters are interpreted in terms of the absolute entropies of anhydrous glucose and the monohydrate.  相似文献   

15.
A simple method for evaluating a range of molecular movements in crystals has been developed. This estimate is needed to calculate the entropy of binding, in particular in protein–ligand complexes. The estimate is based on experimental data concerning the enthalpy of sublimation and saturated vapor pressure obtained for 15 organic crystals with melting temperatures of 25–80°С. For this set, we calculated the values of the average range and the corresponding average amplitude of molecular movements in crystals that constituted 0.75 ± 0.14 Å and 0.18 ± 0.03 Å, respectively. The entropy of sublimation calculated based on the average range of molecular movements in crystals was well consistent with the experimental data.  相似文献   

16.
The macromolecular structural transition of Pf1 filamentous bacterial virus detected by X-ray diffraction analysis has been studied in virus solutions by density, circular dichroism, and microcalorimetric measurements. The reversible structural change occurring between 5 °C and 25 °C has a calorimetrically determined transition enthalpy ΔHt,cal of 14·5 ± 1.5 kJ (mol protein)?1. The transition curves resulting from the density, circular dichroism, and calorimetric measurements have been analysed in terms of a two-state process to extract the van't Hoff enthalpy. Comparison of the effective transition enthalpy and the calorimetric ΔHt,cal values gives about 26 protein subunits as the size of the co-operative unit. Parallel heat capacity and density measurements on fd virus show no such transition, in agreement with X-ray diffraction studies.  相似文献   

17.
The aquation equilibria for cis-dibromodiammineplatinum(II) have been measured. For the first and second aquations at 25°C: K1 = 1.13 mM. ΔH° = 3.9 ± .7 kcal/mol, K2 = 0.042 mM, ΔH° = 10 ± 3 kcal/mol. Some isotopic exchange of 82Br between c-Pt(NH3)2Br2 and PtBr42− occurs without the formation of a free bromide ion and with a rate expression: Rate Exchange = (kac′ + kac″/[Br])[Pt(NH3)2Br2] [PtBr42−]. It has been shown that the presence of c-Pt(NH3)2Br(H2O)+ serves to catalyze the aquation of PtBr42−, but that the catalysis step does not account for all the bromide dependent exchange.  相似文献   

18.
A sedimentation analysis has been used to determine the proportion of protein present as monomer and aggregate in 0.5 and 1.0 g/dl solutions of β-casein A in pH 7 phosphate buffer over the temperature range 10–40°C. The amount and molecular weight of the aggregate increase with temperature; under the conditions used, the aggregation number (n) of β-casein is given approximately by n = 0.6t + 2 with t in degrees centigrade. The concentration of β-casein in monomeric and aggregated states at different temperatures is used to calculate the standard enthalpy of aggregation ΔH° (Van't Hoff) by assuming that β-casein undergoes a cooperative, two-state, micellization process; aggregation is an endothermic process and ΔH° = 66.0 ± 2.6 kJ mol?1. Combination of this ΔH° with the amount of protein calculated to dissociate when 1 g/dl solutions are diluted isothermally to 0.5 g/dl gives the heat of dilution at various temperatures. These calculated heats of dilution are compared with the experimental values obtained by carrying out the same dilutions in a microcalorimeter. The heat of dilution decreases linearly with β-casein concentration, but the extrapolated zero-concentration values of 65.8 ± 1.6 kJ mol?1 is the same as the Van't Hoff enthalpy. This agreement in the enthalpy values indicates that the micellization of β-casein occurs cooperatively. The effect of modifying the hydrophobic/hydrophilic balance of the system on the micellization of β-casein A has been investigated. The hydrophobic interaction between the protein molecules is decreased by removing the three C-terminal residues (Ileu Ileu Val) with carboxypeptidase-A. This modification drastically reduces the ability of the β-casein molecule to form micelles. Substitution of 2H2O for H2O at constant temperature perturbs the monomer–micelle equilibrium in favor of micelles because of enhanced hydrophobic interactions in the former solvent. The results are consistent with β-casein micellization involving a delicate balance of the hydrophobic forces favoring aggregation and electrostatic forces opposing it.  相似文献   

19.
为探讨人载脂蛋白A-Ⅰ(apoA-Ⅰ,apolipoproteinA-Ⅰ)α螺旋不同位点的半胱氨酸突变后,对蛋白二级结构和脂质结合能力的影响,利用定点诱变技术构建apoA-Ⅰ的天然半胱氨酸突变体apoA-ⅠMilano(R173C),及其它α螺旋片段上的半胱氨酸突变体,分别为apoA-Ⅰ(S52C),apoA-Ⅰ(N74C),apoA-Ⅰ(L107C),apoA-Ⅰ(K129C),和apoA-Ⅰ(L195C).观察比较各种野生型及突变apoA-Ⅰ单体蛋白的α螺旋含量和二级结构稳定性及其脂质结合能力.结果显示,野生型apoA-Ⅰ,apoA-Ⅰ(S52C),apoA-Ⅰ(N74C),apoA-Ⅰ(L107C),apoA-Ⅰ(K129C),apoA-ⅠMilano和apoA-Ⅰ(L195C)的α螺旋含量分别为54±4%,49±4%,50±2%,51±6%,56±4%,52±3%,和54±1%,各种蛋白的α螺旋含量无显著性差异(P>0.05).野生型apoA-Ⅰ的变性标准自由能(ΔG0D)为10.5kJ/mol;apoA-Ⅰ(S52C)和apoA-ⅠMilano的ΔG0D比野生型低2.1kJ/mol;而apoA-Ⅰ(K129C)的ΔG0D比野生型apoA-Ⅰ高1.6kJ/mol.与野生型apoA-Ⅰ相比,apoA-Ⅰ(K129C)和apoA-Ⅰ(L195C)两个突变体与脂质结合能力明显下降(P<0.05),而其它半胱氨酸突变体(包括apoA-ⅠMilano)在脂质结合动力学方面与野生型apoA-Ⅰ无明显差异.以上结果提示,不同位点发生的半胱氨酸突变对apoA-Ⅰ单体蛋白的α螺旋含量无明显影响,但对蛋白的二级结构稳定性和脂质结合能力影响不尽相同.  相似文献   

20.
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   

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