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 15 毫秒
1.
The partial molar heat capacities of the tripeptides of the sequence glycyl-X-glycine, where X is one of the amino acids leucine, threonine, glutamine, phenylalanine, histidine, cysteine, proline, glutamic acid or arginine, and of the two tetrapeptides tetraglycine and glycyltryptophanylglycylglycine in aqueous solution over the temperature range 10-100 degrees C have been determined using high sensitivity scanning microcalorimetry. These results were used to derive the partial molar heat capacities of the various amino acid side-chains. This report completes our programme to derive reliable side-chain heat capacities for all 20 amino acids of proteins over a wide temperature range using the tripeptides Gly-X-Gly as realistic model compounds. Included in the study is a summary of the partial molar heat capacities of all 20 amino acid side-chains. These results, along with the heat capacity of the peptide backbone group, were used to calculate the partial molar heat capacities of some oligopeptides and of the random coil form of some unfolded proteins in water. The calculated heat capacities of the proteins obtained using this new set of heat capacities for the constituent groups are consistent with the heat capacities of the denatured state determined experimentally.  相似文献   

2.
The values of k and alpha in the Mark-Houwink equation have been determined for chitosans with different degrees of deacetylation (DD) (69, 84, 91 and 100% respectively), in 0.2 M CH3COOH/0.1 M CH3COONa aqueous solution at 30 degrees C by the light scattering method. It was shown that the values of alpha decreased from 1.12 to 0.81 and the values of k increased from 0.104 x 10(-3) to 16.80 x 10(-3) ml/g, when the DD varied from 69 to 100%. This is due to a reduction of rigidity of the molecular chain and an increase of the electrostatic repulsion force of the ionic groups along the polyelectrolyte chain in chitosan solution, when the DD of chitosan increases gradually.  相似文献   

3.
Integral enthalpies of solution at low concentrations of several amino acids and peptides in 2 and 6M urea solutions have been determined at 25 and 35°C. These data have been used to derive the enthalpies of transfer (at 25 and 35°C) and heat capacities of transfer (at 30°C) of these amino acids and peptides from water to aqueous urea solutions. Furthermore, the enthalpies of transfer and heat capacities of transfer per CH2 group and per peptide group ? CONH? have also been estimated. These results show that while the enthalpies and heat capacities of transfer per CH2 group are positive and negative, respectively, the reverse is true for ? CONH? group. The implications of these results in the mechanism of the denaturation of proteins by urea are discussed.  相似文献   

4.
The partial molar heat capacities have been determined for the series of peptides alanyl(glycyl)(x)glycine, x=1-3, and for the compounds N-acetylglycinamide and N-acetyl glycylglycinamide in aqueous solution over the temperature range 10-100 degrees C using high sensitivity scanning microcalorimetry. The partial molar volumes for these compounds have also been determined over the temperature range from 10 to 90 degrees C using a scanning densimetric method. The results were used to derive the partial molar heat capacities and volumes of the glycyl group at temperatures in the range 10-100 degrees C. The results obtained are critically compared with literature results derived using heat capacity and volume data for some oligoglycines.  相似文献   

5.
Partial molar heat capacities of five linear alcohols (methanol, ethanol, n-propanol, n-butanol, n-pentanol) and five N-substituted amides (n-propionamide, N-methylformamide, N-methylacetamide, N-methylpropionamide, N-ethylacetamide) in aqueous D(2)O solution have been measured at 25 degrees C. The heat capacities of transfer of these compounds from H(2)O to D(2)O were calculated using previously reported (Makhatadze et al., Biophys. Chem. 64 (1997) 93) values of partial heat capacities of alcohols and amides in aqueous H(2)O solutions. It is shown that the sign and magnitude of the heat capacity change upon transfer from H(2)O to D(2)O depends on the relative amount of polar and non-polar solvent accessible surface areas of solute. Analysis shows that transfer of non-polar surface from H(2)O to D(2)O is accompanied by a positive heat capacity change. In contrast, transfer of polar surface from H(2)O to D(2)O occurs with negative heat capacity change. Estimates show that the solvent isotope effect on the heat capacity changes upon protein unfolding can be predicted using the changes of the polar and non-polar surface area changes upon protein unfolding and the transfer data of model compounds. Analysis of the thermodynamic functions of transfer of non-polar compounds from H(2)O to D(2)O shows puzzling behavior which contradicts current definitions of the hydrophobic effect.  相似文献   

6.
Using the heat capacity values for amino acid side-chains and the peptide unit determined in the accompanying paper, we calculated the partial heat capacities of the unfolded state for four proteins (apomyoglobin, apocytochrome c, ribonuclease A, lysozyme) in aqueous solution in the temperature range from 5 to 125 degrees C, with an assumption that the constituent amino acid residues contribute additively to the integral heat capacity of a polypeptide chain. These ideal heat capacity functions of the extended polypeptide chains were compared with the calorimetrically determined heat capacity functions of the heat and acid-denatured proteins. The average deviation of the experimental functions from the calculated ideal ones in the whole studied temperature range does not exceed the experimental error (5%). Therefore, the heat-denatured state of a protein, in solutions with acidic pH preventing aggregation, approximates well the completely unfolded state of this macromolecule. The heat capacity change caused by hydration of amino acid residues upon protein unfolding was also determined and it was shown that this is the major contributor to the observed heat capacity effect of unfolding. Its value is different for different proteins and correlates well with the surface area of non-polar groups exposed upon unfolding. The heat capacity effect due to the configurational freedom gain by the polypeptide chain was found to contribute only a small part of the overall heat capacity change on unfolding.  相似文献   

7.
It is shown that in 0,5 M NaCl 8 M CH3COOH heat absorption and the second structure alteration in a heated solution proceed between two stages following one another, and besides, salts not only decrease the macromolecule denaturation temperature in total, but produce different destabilization effect on different regions. The presence of the thermostable domain in the macromolecule helical part permits investigation of the folding mechanism of the triple collagen helix under partial denaturation. The localization and biological role of the stable domain in the triple helix formation are suggested.  相似文献   

8.
The partial molar isentropic compressibilities at infinite dilution, K(S,2)(o), have been determined for the peptides serylglycine, serylglycylglycine and serylglycylglycylglycine in aqueous solution at 25 degrees C. The partial molar volumes at infinite dilution, V(2)(o), have also been determined for these peptides in aqueous solution at the temperatures 15, 30 and 40 degrees C. These results, along with those obtained previously at 25 degrees C, were used to derive the partial molar exansibilities, E(2)(o), of the peptides at 25 degrees C, which in turn were used to convert the isentropic compressibilities into the partial molar isothermal compressibilities at infinite dilution, K(T,2)(o). These K(S,2)(o) and K(T,2)(o) results were used to obtain the partial molar compressibilities of the glycyl group CH(2)CONH at 25 degrees C. The results are compared with those obtained using data for other series of peptides of sequence ala(gly)(n), n=1-4, and (gly)(n), n=2-5.  相似文献   

9.
固态乙酸钠氮离子注入的辐照效应研究   总被引:1,自引:0,他引:1  
用 25ke V 氮离子辐照固体乙酸钠样品,测定离子注入前后样品的紫外光谱,发现辐照后的样品中产生了新的具有紫外吸收的化学物质,给出了新物质的紫外吸收值的剂量效应关系式。通过高效液相色谱的分析测定,发现经氮离子束辐照后的乙酸钠样品中还产生了另一种新的化学基团一氨基,再结合茚三酮反应的检测,得到了离子束辐照剂量与新产物中氨基产生量的曲线关系。  相似文献   

10.
Y V Griko  P L Privalov 《Biochemistry》1992,31(37):8810-8815
Temperature-induced changes of the states of beta-lactoglobulin have been studied calorimetrically. In the presence of a high concentration of urea this protein shows not only heat but also cold denaturation. Its heat denaturation is approximated very closely by a two-state transition, while the cold denaturation deviates considerably from the two-state transition and this deviation increases as the temperature decreases. The heat effect of cold denaturation is opposite in sign to that of heat denaturation and is noticeably larger in magnitude. This difference in magnitude is caused by the temperature-dependent negative heat effect of additional binding of urea to the polypeptide chain of the protein upon its unfolding, which decreases the positive enthalpy of heat denaturation and increases the negative enthalpy of cold denaturation. The binding of urea considerably increases the partial heat capacity of the protein, especially in the denatured state. However, when corrected for the heat capacity effect of urea binding, the partial heat capacity of the denatured protein is close in magnitude to that expected for the unfolded polypeptide chain in aqueous solution without urea but only for temperatures below 10 degrees C. At higher temperatures, the heat capacity of the denatured protein is lower than that expected for the unfolded polypeptide chain. It appears that at temperatures above 10 degrees C not all the surface of the beta-lactoglobulin polypeptide chain is exposed to the solvent, even in the presence of 6 M urea; i.e., the denatured protein is not completely unfolded and unfolds only at temperatures lower than 10 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Heat capacity and conformation of proteins in the denatured state   总被引:30,自引:0,他引:30  
Heat capacity, intrinsic viscosity and ellipticity of a number of globular proteins (pancreatic ribonuclease A, staphylococcal nuclease, hen egg-white lysozyme, myoglobin and cytochrome c) and a fibrillar protein (collagen) in various states (native, denatured, with and without disulfide crosslinks or a heme) have been studied experimentally over a broad range of temperatures. It is shown that the partial heat capacity of denatured protein significantly exceeds the heat capacity of native protein, especially in the case of globular proteins, and is close to the value calculated for an extended polypeptide chain from the known heat capacities of individual amino acid residues. The significant residual structure that appears at room temperature in the denatured states of some globular proteins (e.g. myoglobin and lysozyme) at neutral pH results in a slight decrease of the heat capacity, probably due to partial screening of the protein non-polar groups from water. The heat capacity of the unfolded state increases asymptotically, approaching a constant value at about 100 degrees C. The temperature dependence of the heat capacity of the native state, which can be determined over a much shorter range of temperature than that of the denatured state and, correspondingly, is less certain, appears to be linear up to 80 degrees C. Therefore, the denaturational heat capacity increment seems to be temperature-dependent and is likely to decrease to zero at about 140 degrees C.  相似文献   

12.
Integral enthalpies of solution of several dipeptides and tripeptides in water at low concentrations have been determined at 25 and 35°C. These data have been used to derive the changes in heat capacity on dissolution at infinite dilution ΔC at 30°C. Limiting partial molal heat capacities ΔC have been determined by combining ΔC with Cp2 (heat capacity of pure solid peptides). Using the data on ω-amino acids and these peptides, the partial molal heat capacity of a peptide group ? CONH? was semiquantitatively estimated.  相似文献   

13.
Partial molar heat capacities of the three tripeptides glycylglycylglycine, glycyl-L-alanylglycine and glycyl-DL-threonylglycine in aqueous solution at the temperatures 50, 75, 100 and 125 degrees C have been determined using differential flow calorimetry. The results have been used to estimate the contributions to the partial molar heat capacities of peptides of the alanyl and threonyl side-chains. These side-chain contributions are compared with those reported in the literature.  相似文献   

14.
The heat capacity change upon protein unfolding has been analysed using the heat capacity data for the model compounds' transfer into water, corrected for volume effects. It has been shown that in the unfolding, the heat capacity increment is contributed to by the effect of hydration of the non-polar groups, which is positive and decreases with temperature increase, and by the effect of hydration of the polar groups, which is negative and decreases in magnitude as temperature increases. The sum of these two effects is very close to the total heat capacity increment of protein unfolding at room temperature but is likely to deviate from it at higher temperatures. Therefore, the expected heat capacity effect caused by the increase of configurational freedom of the polypeptide chain upon unfolding seems to be compensated for by some other effect, perhaps associated with fluctuation of the native protein structure.  相似文献   

15.
Photoanlogues of the initiation substrates of the RNA polymerase II, N3ArNH(CH2)(n)NHpppA where N3Ar is 5-azido-2-nitrobenzoyl group (n = 2 or 4) were synthesized, allowing the preparation of photoreactive oligonucleotides in situ by RNA polymerase II for application as photolabels. Photolysis of p-nitro-substituted aromatic azide in aqueous medium was investigated. Using the azoxy-coupling reaction it was possible to determine whether a nitrene or p-nitrophenyl hydroxylamine azoxy compound is the trappable intermediate that is generated at ambient temperature in aqueous solution.  相似文献   

16.
Long aliphatic hydrocarbon chains aggregate in aqueous solution due to the hydrophobic effect, forming structures such as micelles and membranes, while amino groups titrate at basic pH. These two biologically important behaviors are linked in alkylamines, in which the pK(a) of the amino group is shifted downward by aggregation. In this paper we study the thermodynamics of these coupled processes, following aggregation by observing alkylamine pH titration behavior. The magnitude of the shift depended on the aliphatic chain length and on the concentration of alkylamine: longer chains and higher concentrations lowered the pK(a) to a greater extent. Gibbs free energies of protonation and aggregation were calculated from the pK(a) shifts. Enthalpies, entropies, and heat capacities were estimated by van't Hoff analysis from the pK(a) shift dependencies on temperature. However, the results were less precise than the calorimetrically measured values, as described in the following article. A model to calculate titration curves, pK(a) shifts, and aggregation of uncharged alkylamines as a function of aliphatic chain length, concentration, and temperature is presented.  相似文献   

17.
Physiological processes such as protein folding and molecular recognition are intricately linked to their dynamic signature, which is reflected in their thermal coefficient. In addition, the local conformational entropy is directly related to the degrees of freedom, which each residue possesses within its conformational space. Therefore, the temperature dependence of the local conformational entropy may provide insight into understanding how local dynamics may affect the stability of proteins. Here, we analyze the temperature dependence of internal methyl group dynamics derived from the cross-correlated relaxation between dipolar couplings of two CH bonds within ubiquitin. Spanning a temperature range from 275 to 308 K, internal methyl group dynamics tend to increase with increasing temperature, which translates to a general increase in local conformational entropy. With this data measured over multiple temperatures, the thermal coefficient of the methyl group order parameter, the characteristic thermal coefficient, and the local heat capacity were obtained. By analyzing the distribution of methyl group thermal coefficients within ubiquitin, we found that the N-terminal region has relatively high thermostability. These results indicate that methyl groups contribute quite appreciably to the total heat capacity of ubiquitin through the regulation of local conformational entropy.  相似文献   

18.
The heat capacity plays a major role in the determination of the energetics of protein folding and molecular recognition. As such, a better understanding of this thermodynamic parameter and its structural origin will provide new insights for the development of better molecular design strategies. In this paper we have analyzed the absolute heat capacity of proteins in different conformations. The results of these studies indicate that three major terms account for the absolute heat capacity of a protein: (1) one term that depends only on the primary or covalent structure of a protein and contains contributions from vibrational frequencies arising from the stretching and bending modes of each valence bond and internal rotations; (2) a term that contains the contributions of noncovalent interactions arising from secondary and tertiary structure; and (3) a term that contains the contributions of hydration. For a typical globular protein in solution the bulk of the heat capacity at 25°C is given by the covalent structure term (close to 85% of the total). The hydration term contributes about 15 and 40% to the total heat capacity of the native and unfolded states, respectively. The contribution of non-covalent structure to the total heat capacity of the native state is positive but very small and does not amount to more than 3% at 25°C. The change in heat capacity upon unfolding is primarily given by the increase in the hydration term (about 95%) and to a much lesser extent by the loss of noncovalent interactions (up to ~5%). It is demonstrated that a single universal mathematical function can be used to represent the partial molar heat capacity of the native and unfolded states of proteins in solution. This function can be experimentally written in terms of the molecular weight, the polar and apolar solvent accessible surface areas, and the total area buried from the solvent. This unique function accurately predicts the different magnitude and temperature dependences of the heat capacity of both the native and unfolded states, and therefore of the heat capacity changes associated with folding/unfolding transitions. © 1995 Wiley-Liss, Inc.  相似文献   

19.
The organometallic precursor (NEt(4))(2)[ReBr(3)(CO)(3)] was reacted with bidendate dithioethers (L) of the general formula H(3)C-S-CH(2)CH(2)-S-R (R = -CH(2)CH(2)COOH, CH(2)-C&tbd1;CH) and R'-S-CH(2)CH(2)-S-R' (R' = CH(3)CH(2)-, CH(3)CH(2)-OH, and CH(2)COOH) in methanol to form stable rhenium(I) tricarbonyl complexes of the general composition [ReBr(CO)(3)L]. Under these conditions, the functional groups do not participate in the coordination. As a prototypic representative of this type of Re compounds, the propargylic group bearing complex [ReBr(CO(3))(H(3)C-S-CH(2)CH(2)-S-CH(2)C&tbd1;CH)] Re2 was studied by X-ray diffraction analysis. Its molecular structure exhibits a slightly distorted octahedron with facial coordination of the carbonyl ligands. The potentially tetradentate ligand HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH was reacted with the trinitrato precursor [Re(NO(3))(3)(CO)(3)](2-) to yield a cationic complex [Re(CO)(3)(HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH)]NO(3) Re8 which shows the coordination of one hydroxy group. Re8 has been characterized by correct elemental analysis, infrared spectroscopy, capillary electrophoresis, and X-ray diffraction analysis. Ligand exchange reaction of the carboxylic group bearing ligands H(3)C-S-CH(2)CH(2)-S-CH(2)CH(2)-COOH and HOOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH with (NEt(4))(2)[ReBr(3)(CO)(3)] in water and with equimolar amounts of NaOH led to complexes in which the bromide is replaced by the carboxylic group. The X-ray structure analysis of the complex [Re(CO)(3)(OOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH)] Re6 shows the second carboxylic group noncoordinated offering an ideal site for functionalization or coupling a biomolecule. The no-carrier-added preparation of the analogous (99m)Tc(I) carbonyl thioether complexes could be performed using the precursor fac-[(99m)Tc(H(2)O)(3)(CO)(3)](+), with yields up to 90%. The behavior of the chlorine containing (99m)Tc complex [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 in aqueous solution at physiological pH value was investigated. In saline, the chromatographically separated compound was stable for at least 120 min. However, in chloride-free aqueous solution, a water-coordinated cationic species Tc1a of the proposed composition [(99m)Tc(H(2)O)(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))](+) occurred. The cationic charge of the conversion product was confirmed by capillary electrophoresis. By the introduction of a carboxylic group into the thioether ligand as a third donor group, the conversion could be suppressed and thus the neutrality of the complex preserved. Biodistribution studies in the rat demonstrated for the neutral complexes [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 and [(99m)TcCl(CO)(3)(CH(2)-S-CH(2)CH(2)-S-CH(2)-C&tbd1;CH)] Tc2 a significant initial brain uptake (1.03 +/- 0.25% and 0.78 +/- 0.08% ID/organ at 5 min. p.i.). Challenge experiments with glutathione clearly indicated that no transchelation reaction occurs in vivo.  相似文献   

20.
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