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 thermal stability of Taq DNA polymerase is well known, and is the basis for its use in PCR. A comparative thermodynamic characterization of the large fragment domains of Taq (Klentaq) and E. coli (Klenow) DNA polymerases has been performed by obtaining full Gibbs‐Helmholtz stability curves of the free energy of folding (ΔG) versus temperature. This analysis provides the temperature dependencies of the folding enthalpy and entropy (ΔH and ΔS), and the heat capacity (ΔCp) of folding. If increased or enhanced non‐covalent bonding in the native state is responsible for enhanced thermal stabilization of a protein, as is often proposed, then an enhanced favourable folding enthalpy should, in general, be observed for thermophilic proteins. However, for the KlenowKlentaq homologous pair, the folding enthalpy (ΔHfold) of Klentaq is considerably less favorable than that of Klenow at all temperatures. In contrast, it is found that Klentaq's extreme free energy of folding (ΔGfold) originates from a significantly reduced entropic penalty of folding (ΔSfold). Furthermore, the heat capacity changes upon folding are similar for Klenow and Klentaq. Along with this new data, comparable extended analysis of available thermodynamic data for 17 other mesophilic–thermophilic protein pairs (where enough applicable thermodynamic data exists) shows a similar pattern in seven of the 18 total systems. When analyzed with this approach, the more familiar “reduced ΔCp mechanism” for protein thermal stabilization (observed in a different six of the 18 systems) frequently manifests as a temperature dependent shift from enthalpy driven stabilization to a reduced‐entropic‐penalty model. Proteins 2014; 82:785–793. © 2013 Wiley Periodicals, Inc.  相似文献   

2.
Extension of the (isothermal) Gibbs-Helmholtz equation for the heat capacity terms (ΔC(p)) allows formulating a temperature function of the free (Gibbs) energy change (ΔG). An approximation of the virtually unknown ΔC(p) temperature function enables then to determine and numerically solve temperature functions of thermodynamic parameters ΔH and ΔS (enthalpy and entropy change, respectively). Analytical solutions and respective numeric procedures for several such approximation formulas are suggested in the presented paper. Agreement between results obtained by this analysis with direct microcalorimetric measurements of ΔH (and ΔC(p) derived from them) was approved on selected cases of biochemical interactions presented in the literature. Analysis of several ligand-membrane receptor systems indicates that temperature profiles of ΔH and ΔS are parallel, largely not monotonic, and frequently attain both positive and negative values within the current temperature range of biochemical reactions. Their course is determined by the reaction change of heat capacity: temperature extremes (maximum or minimum) of both ΔH and ΔS occur at ΔC(p)=0, for most of these systems at roughly 285-305 K. Thus, the driving forces of these interactions may change from enthalpy-, entropy-, or enthalpy-entropy-driven in a narrow temperature interval. In contrast, thermodynamic parameters of ligand-macromolecule interactions in solutions (not bound to a membrane) mostly display a monotonic course. In the case of membrane receptors, thermodynamic discrimination between pharmacologically defined groups-agonists, partial agonists, antagonists-is in general not specified and can be achieved, in the best, solely within single receptor groups.  相似文献   

3.
4.
J Alvarez  R Biltonen 《Biopolymers》1973,12(8):1815-1828
The heats of solution of thymine in water and ethanol have been determined calorimetrically as a function of temperature. These data, along with solubility data, have been used to calculate the thermodynamic quantities (ΔGt, ΔHt, ΔSt and ΔCp,t) associated with the transfer of thymine from ethanol to water. Since ΔSt = ?2 cal/mole deg and ΔCp,t = 0, it has been concluded that hydrophobic bonding does not play an important role in the thermocynamic stability of nucleic acids. However, large heat capacities of solution of thymine are observed in both solvents (ΔC°p2 = 45 ± 4 cal/mole deg). This is explained in terms of temperature variation in the degree of solvent–solute hydrogen bonding. It is our proposal that the components of macromolecules (i.e., nucleic acid bases and amino acids) do not make all possible hydrogen bonds with the solvent in the vicinity of room temperature. Thus the thermodynamic contribution of hydrogen bonding to the stability of macromolecules in aqueous solution must be reassessed.  相似文献   

5.
The interactions of imidazolium bashed ionic liquid-type cationic gemini surfactant ([C12-4-C12im]Br2) with HSA were studied by fluorescence, time-resolved fluorescence, UV-visible, circular dichroism, molecular docking and molecular dynamic simulation methods. The results showed that the [C12-4-C12im]Br2 quenched the fluorescence of HSA through dynamic quenching mechanism as confirmed by time-resolved spectroscopy. The Stern–Volmer quenching constant (Ksv) and relevant thermodynamic parameters such as enthalpy change (ΔH), Gibbs free energy change (ΔG) and entropy change (ΔS) for interaction system were calculated at different temperatures. The results revealed that hydrophobic forces played a major role in the interactions process. The results of synchronous fluorescence, UV-visible and CD spectra demonstrated that the binding of [C12-4-C12im]Br2 with HSA induces conformational changes in HSA. Inquisitively, the molecular dynamics study contribute towards understanding the effect of binding of [C12-4-C12im]Br2 on HSA to interpret the conformational change in HSA upon binding in aqueous solution. Moreover, the molecular modelling results show the possible binding sites in the interaction system.  相似文献   

6.
Our understanding of the energetics that govern antigen–antibody recognition lags behind the increasingly rapid accumulation of structural information on antigen–antibody complexes. Thanks to the development of highly sensitive microcalorimeters, the thermodynamic parameters of antigen–antibody interactions can now be measured with precision and using only nanomole quantities of protein. The method of choice is isothermal titration calorimetry, in which a solution of the antibody (or antigen) is titrated with small aliquots of the antigen (or antibody) and the heat change accompanying the formation of the antigen–antibody complex is measured with a sensitivity as high as 0.1 μcal s−1. The free energy of binding (ΔG), the binding enthalpy (ΔH), and the binding entropy (ΔS) are usually obtained from a single experiment, and no spectroscopic or radioactive label must be introduced into the antigen or antibody. The often large and negative change in heat capacity (ΔCp) accompanying the formation of an antigen–antibody complex is obtained from ΔHmeasured at different temperatures. The basic theory and the principle of the measurements are reviewed and illustrated by examples. The thermodynamic parameters relate to the dynamic physical forces that govern the association of the freely moving antigen and antibody into a well-structured and unique complex. This information complements the static picture of the antigen–antibody complex that results from X-ray diffraction analysis. Attempts to correlate dynamic and static aspects are discussed briefly.  相似文献   

7.
Values for the thermodynamic quantities, ΔH° = 11.8 ± 2.0 Kcal/mole and ΔS° = 43.6 ± 6.0 e.u., of the 3-13 helix–coil equilibrium of isolated S-peptide (19 residue N-terminal fragment of ribonuclease A) in aqueous solution (3 m M, 1M NaCl, pD 5.4) have been determined from a joint analysis of the Thr 3γ, Ala 6β, Phe 8meta, and Phe 8para 1H chemical shift vs temperature curves (?7 to 80°C) in several aqueous–trifluorethanol mixtures. Chemical shifts in the coil and in the helix have been determined for up to 16 protons belonging to the 3-13 fragment. Thermodynamic parameters have also been determined for C-peptide (13 residue fragment) and a number of S-peptide derivatives. From the variation of the values of the thermodynamic parameters at pD 2.5, 5.4, and 8.0, a quantitation of the two helix-stabilizing side-chain interactions can be made: (1) Δ(ΔH°) ? 5 Kcal/mole and Δ(ΔS°) ? 18 e.u. for the salt bridge Glu 2? … Arg 10+ and (2) Δ(ΔH°) ? 3 Kcal/mole and Δ(ΔS°) = 9 e.u. for the one in which the His 12+ imidazolium group is involved, presumably a partial stacking with the Phe 8 side chain.  相似文献   

8.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

9.
Differential scanning microcalorimetry was used to investigate the enthalpy (ΔHd) and the temperature (td) of thermal denaturation of normal (nondeuterated) (H-PC) and deuterated (D-PC) phycocyanins in D2O solvent. Values of td in D-PC are about 5–7°C lower than those in H-PC. The magnitudes of ΔHd in D-PC are only 21–32% of those in H-PC. During the protein unfolding, the heat-capacity changes (ΔCp) in D-PC are also lower than those in H-PC. CD was employed to evaluate the secondary structure and the urea denaturation of these proteins in D2O solvent. These proteins have about the same α-helix content. D-PC is less resistant to the denaturant urea than is H-PC. In general, the apparent free-energy change in the process of protein unfolding at zero denaturant concentration is higher in H-PC than in D-PC. Comparisons of the present results for D2O solvent with those previously reported for H2O reveal that solvent isotope effect essentially does not change the α-helix content in H-PC and D-PC. However, D-PC or H-PC has a higher random-coil content in its secondary structure in D2O than in H2O. Substitution of H2O with D2O as the solvent increases td in both D-PC and H-PC, lowers ΔHd in H-PC, and greatly lowers ΔHd in D-PC. The deuterium solvent isotope effect does not change ΔCp in H-PC but lowers ΔCp in D-PC. In the urea denaturation, the magnitudes of (Cu)1/2 in H-PC and D-PC are not affected by such a solvent effect, whereas those of ΔG are greatly increased. These results are correlated with the structure and stability of the proteins.  相似文献   

10.
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   

11.
R A Haworth  J A Chapman 《Biopolymers》1977,16(9):1895-1906
Electron microscopic autoradiographic observations on collagen fibrils grown in vitro allow growth rates in the N- and C-terminal directions to be measured on individual fibrils. Such observations, made on normal and iodinated collagen, show that normal fibrils grow at both ends (although rather more rapidly at the N-terminal end), whereas fully-iodinated collagen fibrils grow only at the N-terminal end. Measurements of growth rates at different temperatures provide estimates of the activation enthalpy (ΔH) and entropy (ΔS) of precipitation for the two types of collagen. Solubility measurements have also yielded values for the thermodynamic enthalpy (ΔH) and entropy (ΔS) of precipitation. Results show that the activated (rate-limiting) state is characterized by a large positive ΔH and ΔS similar in magnitude to the ΔH and ΔS of transition from solution to fibril. It is also concluded that the different rates of precipitation of normal and iodinated collagen cannot be explained in terms of fibril formation requiring ionization of the tyrosine residues.  相似文献   

12.
Combined effects of flunitrazepam (FNZ) and lidocaine (LDC) were studied on the thermotropic equilibrium of dipalmitoyl phosphatidylcholine (dpPC) bilayers. This adds a thermodynamic dimension to previously reported geometric analysis in the erythrocyte model. LDC decreased the enthalpy and temperature for dpPC pre- and main-transitions (ΔH p, ΔH m, T p, T m) and decreased the cooperativity of the main-transition (ΔT 1/2,m). FNZ decreased ΔH m and, at least up to 59 μM, also decreased ΔH p. In conjunction with LDC, FNZ induced a recovery of ?T 1/2,m control values and increased ΔH m even above the control level. The deconvolution of the main-transition peak at high LDC concentrations revealed three components possibly represented by: a self-segregated fraction of pure dpPC, a dpPC–LDC mixture and a phase with a lipid structure of intermediate stability associated with LDC self-aggregation within the lipid phase. Some LDC effects on thermodynamic parameters were reverted at proper LDC/FNZ molar ratios, suggesting that FNZ restricts the maximal availability of the LDC partitioned into the lipid phase. Thus, beyond its complexity, the lipid–LDC mixture can be rationalized as an equilibrium of coexisting phases which gains homogeneity in the presence of FNZ. This work stresses the relevance of nonspecific drug–membrane binding on LDC–FNZ pharmacological interactions and would have pharmaceutical applications in liposomal multidrug-delivery.  相似文献   

13.
Histamine was immobilized on Sepharose CL‐6B (Sepharose) for use as a ligand of hydrophobic charge induction chromatography (HCIC) of proteins. Lysozyme adsorption onto Histamine‐Sepharose (HA‐S) was studied by adsorption equilibrium and calorimetry to uncover the thermodynamic mechanism of the protein binding. In both the experiments, the influence of salt (ammonium sulfate and sodium sulfate) was examined. Adsorption isotherms showed that HA‐S exhibited a high salt tolerance in lysozyme adsorption. This property was well explained by the combined contributions of hydrophobic interaction and aromatic stacking. The isotherms were well fitted to the Langmuir equation, and the equilibrium parameters for lysozyme adsorption were obtained. In addition, thermodynamic parameters (ΔHads, ΔSads, and ΔGads) for the adsorption were obtained by isothermal titration calorimetry by titrating lysozyme solutions into the adsorbent suspension. Furthermore, free histamine was titrated into lysozyme solution in the same salt‐buffers. Compared with the binding of lysozyme to free histamine, lysozyme adsorption onto HA‐S was characterized by a less favorable ΔGads and an unfavorable ΔSads because histamine was covalently attached to Sepharose via a three‐carbon‐chain spacer. Consequently, the immobilized histamine could only associate with the residues on the protein surface rather than those in the hydrophobic pocket, causing a less favorable orientation between histamine and lysozyme. Further comparison of thermodynamic parameters indicated that the unfavorable ΔSads was offset by a favorable ΔHads, thus exhibiting typical enthalpy‐entropy compensation. Moreover, thermodynamic analyses indicated the importance of the dehydration of lysozyme molecule and HA‐S during the adsorption and a substantial conformational change of the protein during adsorption. The results have provided clear insights into the adsorption mechanisms of lysozyme onto the new HCIC material. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

14.
In the present study a novel biomass, derived from the pulp of Saccharum bengalense, was used as an adsorbent material for the removal of Pb (II) ions from aqueous solution. After 50 minutes contact time, almost 92% lead removal was possible at pH 6.0 under batch test conditions. The experimental data was analyzed using Langmuir, Freundlich, Timken and Dubinin-Radushkevich two parameters isotherm model, three parameters Redlich—Peterson, Sip and Toth models and four parameters Fritz Schlunder isotherm models. Langmuir, Redlich—Peterson and Fritz-Schlunder models were found to be the best fit models. Kinetic studies revealed that the sorption process was well explained with pseudo second-order kinetic model. Thermodynamic parameters including free energy change (ΔG°), enthalpy change (ΔH°) and entropy change (ΔS°) have been calculated and reveal the spontaneous, endothermic and feasible nature of the adsorption process. The thermodynamic parameters of activation (ΔG #, ΔH #and ΔS #) were calculated from the pseudo-second order rate constant by using the Eyring equation. Results showed that Pb (II) adsorption onto SB is an associated mechanism and the reorientation step is entropy controlled.  相似文献   

15.
One of the most powerful techniques that are currently available to measure thermodynamic parameters such as enthalpy (ΔH), Gibbs free energy (ΔG), entropy changes (ΔS), and binding affinity in chemical reactions is isothermal titration calorimetry (ITC). Recent advances in instrumentation have facilitated the development of ITC as a very essential analytical tool in biology and chemistry. In this article, we will focus on a review of the literature on the application of ITC for the study of chiral systems and chiral interactions. We present studies in which the ITC technique is used to study chiral interactions, for instance in chiral solutions, chiral organometallic complexes, guest‐host chiral binding interactions, and biological macromolecules. Finally, we put strong emphasis on the most recent application of ITC for the study of chirality in nanosystems and at the nanoscale.  相似文献   

16.
Jong Jin Lim 《Biopolymers》1976,15(12):2371-2383
The transition temperatures tt and enthalpy changes ΔH in the helix–coil transition of solid tendon collagen soaked in a solution containing one of the following stabilizing or destabilizing agents, HCHO, NaF, NaCl, NaI, NaBr, NaOH, NH2CONH2, CaCl2, MgCl2, were measured as a function of molar concentration by a calorimetric method. The temperature and the enthalpy changes accompanying the transition behaved in a similar manner: when the tt was depressed by the presence of ions, similar behaviour was observed in ΔH. Both parameters (tt and ΔH) increased for HCHO, and decreased for NaF and NaCl at concentrations lower than 0.2 M. Above 0.2 M they increased for NaF and NaCl, and decreased in the presence of the other reagents listed above. The average tt and the ΔH observed in collagen soaked in water were 63.5°C and 12.3 cal/g, respectively. In addition to the parameters mentioned above, the molar effectiveness of the various reagents was obtained for the cases where there was a linear relationship between the tt and molar concentration of the reagent in the solution. Since both the tt and the ΔH were observed to vary, the entropy change (ΔS) accompanying the transition was calculated using thermodynamic relations. In order to explain the ΔS observed as a function of ionic concentration, the thermodynamic relationships have been obtained from a partition function under suitable assumptions. Since the partition function is dependent on the number of hydrogen bonds responsible for collagen stability, the result obtained has been compared with the values predicted by the two most quoted models for collagen. The present study is in accordance with the Ramachandran model for collagen structure, which predicts more than one hydrogen bond per three residues.  相似文献   

17.
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   

18.
The interactions between the HIV Rev-responsive element (RRE) RNA and the HIV regulatory protein Rev, are crucial for the HIV life-cycle. Earlier, we showed that single C2H2 zinc fingers (znfs) have the same binding site as the Rev peptide and exhibit nanomolar affinity. In this study, the specific role of amino acid side chains and molecular processes involved with complex formation were investigated by perturbation of the binding energetics via changes in temperature, pH, buffers, and salt concentrations, as well as znf and RNA mutations, by isothermal titration calorimetry. Interestingly, despite the large cationic charge on the znfs, the number of interactions with the RNA phosphate backbone was lower than intuitively expected. The presence of binding induced protonation was established by ITC and localized by NMR to a histidine on the znf β-sheet. The ΔCp of znf-RNA binding was observed to be substantially negative and could not be accounted for by conventional solvent-accessible surface area models. An alternative model, based on the extent of hydrogen bond changes as a result of differences in ligand-induced water displacement at the binding site, provided reasonable explanation of the trends in ΔCp, as well as ΔH and ΔS. Our studies show that incorporation of favorable interactions at the solvent-excluded binding interface can be used to alleviate the unfavorable enthalpic penalties of displacing water molecules from the hydrated RNA surface.  相似文献   

19.
To get an idea about the pharmacokinetics and pharmacodynamics, it is important to study the drug‐protein interaction. Therefore, herein, we studied the interaction of diclofenac sodium (DIC) with human hemoglobin. The binding study of nonsteroidal antiinflammatory drug, DIC with human hemoglobin (HHB) was done by utilizing fluorescence, UV–visible, time‐resolved fluorescence and far‐UV circular dichroism spectroscopy (CD). Various thermodynamic parameters such as enthalpy change (ΔH), entropy change (ΔS), and Gibbs free energy change (ΔG) were also calculated. CD results showed that DIC induces secondary structure change in HHB. Fluorescence resonance energy transfer was also performed. Additionally, it was also observed that DIC inhibits the esterase‐like enzymatic activity of HHB via competitive inhibition.  相似文献   

20.
Differential scanning microcalorimetry was used to investigate the enthalpy (ΔHd) and the temperature (td) of thermal denaturation of normal and deuterated phycocyanins isolated from two blue-green algae, Plectonema calothricoides and Phormidium luridum. Values of td in deuterated proteins are about 5°C lower than those in normal proteins. The magnitudes of ΔHd in deuterated proteins are 18–36% lower than in normal proteins. The heatcapacity change (ΔCp) in protein unfolding is essentially the same (2 kcal/mol/K) for deuterated and normal proteins within the experimental error. At close to physiological temperature (27°C), the differences in thermodynamic functions in the native and denatured states are much higher in normal proteins than in deuterated proteins. CD was employed to evaluate both the secondary structures and urea denaturation of these two types of proteins. In P. luridum, deuterated protein is about 8% higher in α-helix content; in P. calothricoides it is not significantly higher. Deuterated proteins are less resistant to the denaturant urea than are normal proteins: the denaturant concentration at the midpoint of the denaturation curve is 0.6–1.2 mol/L lower in the deuterated proteins. The apparent free energies of unfolding of deuterated proteins at zero denaturant concentration are 1.1–1.5 kcal/mol less than for normal proteins.  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