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.
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   

2.
The thermal perturbation difference spectra of phenolic and indolic chromophores in water resemble the isothermal D2O and H2O spectra of these chromophores. For phenols approximately equal Δ? values are obtained in both types of spectra, but for their methyl ethers Δ? values of D2O vs H2O spectra are about half of those of the thermal perturbation spectra. Phenols and their methyl ethers were studied in deuterated ethylene glycol and glycerol vs the corresponding protiated solvent, and in nonprotic solvents containing 0.25–4% D2O or H2O. For phenols in D2O vs H2O, about one-third to one-half of the difference spectrum is attributed to solvent structure difference, and the remainder to the effects of replacing OH by OD and to differences in accepting hydrogen bonds from D2O and H2O. The refractive index difference between D2O and H2O was shown to be a minor contribution by means of experiments in which D2O was at 5 dgC and H2O at 47 dgC, conditions of equal refractive index (NaD). D2O vs H2O and glycerol-d vs glycerol-h difference spectra of ribonuclease are about twice as large as expected from the known number of exposed tyrosyl side chains. Possible sources of error in D2O vs H2O spectra of proteins are discussed.  相似文献   

3.
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   

4.
P Appel  W D Brown 《Biopolymers》1971,10(11):2309-2322
Purified sperm whale myoglobin was deuterated by being exposed to pD 3.5 in D2O buffer for 1 hr, then raised to pD 10.6 for an additional hour, and finally brought to neutrality in a D2O environment. Control myoglobin was similarly treated in H2O. The specific rotation at 233 mμ and/or the absorbance in the Soret region were used to follow the helix-coil transition of myoglobin subjected to denaturation by acid, alkali, urea and guanidine. Deuterated and control myoglobin had similar 50% transition points in the four denaturing media studied (acid: pH 4.4, pD 4.9; alkali: pH 9.4, pD 10.0; urea, 7.2M; guanidine, 2.5M). The shift toward the alkaline side of 0.5 or 0.6 units of the transition induced by either acid or alkaline denaturation reflects only the weakened acidity of ionizable groups in deuterium systems. Deuterated myoglobin in 3.25M guanidine had a 20% faster denaturation rate than that of control. Acid, urea, and guanidine denaturation curves showed fairly steep transitions, taken as indicative of a one-step process involving only two states (native and denatured molecules). Supporting this conclusion was the fact that plots of absorption and polarimetry measurements of the helix-coil transition induced by either acid or guanidine could be superimposed.  相似文献   

5.
D A Torchia  C M Deber 《Biopolymers》1972,11(3):653-659
NMR spectra of cyclo(tri-L -prolyl), c-(P)3, show large shifts of the Hα resonance on adding C6D6 to a solution of c-(P)3 in CD2Cl2. CPK models and observed coupling constants indicate a rigid c-(P)3 conformation, independent of solvent composition, suggesting that these shifts result from formation of stereospecific C6D6–c-(P)3 collision complexes in which the c-(P)3 Hα lie near the face of the aromatic ring. The temperature dependence of the Hα shifts and the solvent dependent shifts observed on adding toluene-d8 or nitrobenzene-d5 to the c-(P)3 solution suggest that preferred C6D6–c-(P)3 orientations result from attractive interactions between the electron-rich aromatic ring and the electropositive Hα's and/or δ+ nitrogen atoms in the peptide backbone. Reports of such interactions in increasingly diverse peptide model systems suggest that they may play a role in stabilizing protein structures.  相似文献   

6.
The endotherm enthalpy changes ΔHD and temperatures TD of thermal denaturation of tropocollagen fibers were measured by DSC calorimetry as functions of water content. The denaturation temperatures decrease with increasing water content. The enthalpy change values increase sharply in the range 0–28% of water content, where a maximum of 14.3 cal g?1 is reached. The effect of water uptake on the enthalpy term is explained by water bridge formation within the collagen triple helix. Evidence is given for the existence of approximately three intercatenary water bridges per triplet at the enthalpy maximum, their H-bond energy amounting to approximately 4000 kcal/mol of protein. In the 30–60% range of water content, ΔHD decreases by 2 cal?1 probably due to interactions between secondary water structures and the stabilizing intrahelical water bonds. The influence of two neutral potassium salts, with a structure-stabilizing and a structure-breaking anion (F? and I?), on the hydration dependence of ΔHD and TD was also studied. It was shown that the primary hydration is not influenced by these ions, but that TD and ΔHD are altered in an ion specific way in the presence of interface and bulk water. Hydrophobic interactions do not explain the experimental results. A reaction mechanism of the effects of ions upon the structural stability of collagen is proposed and discussed in terms of interactions of the medium water molecules with the intrahelical water bonds, and in terms of proton-donor/proton-acceptor equilibria between peptide groups, hydrated ions, and intrahelical water molecules.  相似文献   

7.
The interpretation of ΔG (the free energy change for the reaction, globular conformation ? randomly coiled conformation, in the absence of denaturant), in terms of the free energies of transfer of various parts of the protein molecule from water to denaturant solution, is unsatisfactory because the latter are assumed to be identical to the transfer-free energies of similar groups attached to smaller model compounds. We have made empirical adjustments to transfer-free energy theory that make possible linear extrapolation of the free energy of denaturation of a protein from transition region to zero denaturant concentration. The modified theory, used to analyze the denaturation of proteins by guanidine hydrochloride and urea, allowed us to calculate reasonable values for Δα, the average change in accessibility to solvent of the component groups of protein.  相似文献   

8.
The interaction between puerarin and β-cyclodextrin (CD) has been studied in D2O, H2O/acetone-d 6, acetone-d 6 and DMSO-d 6 solutions by 1H NMR spectroscopy. The NMR data obtained from hydroxy protons indicate that the formation of the inclusion complex between the two molecules is not stabilized by strong hydrogen bond interactions. The sugar part of puerarin as well as the A ring are outside the β-CD cavity while the B and C rings are located inside the cavity and the interaction is mainly stabilized by hydrophobic interactions. In DMSO at 30°C and in acetone-d 6/H2O at temperature below −5°C, doubling of some signals indicated that, in these solvent systems, free rotation of the C-glycosyl bond was restricted due to the steric hindrance between the phenolic hydroxy group at C-7 and the bulky sugar moiety at C-8. In acetone, fast exchange of phenolic protons on the NMR timescale was observed, showing the effect of the solvent on the hindered rotation.  相似文献   

9.
10.
11.
12.
The compatible osmolyte glycine betaine (GB) is the most efficient osmoprotectant and best excluder from the protein surface. It can reverse protein aggregation and correct mutant protein defects and counter the harmful effects of urea and salts in vivo and in vitro. In this study we have investigated the pH dependence of the stabilizing effect of GB on three different proteins, namely, α-lactalbumin (α-LA), lysozyme and ribonuclease-A (RNase-A). We show here that (a) GB stabilizes RNase-A at all pH values, and (b) GB has opposite effects on two proteins at high pH and low pH values, namely, α-LA and lysozyme. This conclusion was reached by determining Tm (midpoint of denaturation), ΔHm (denaturational enthalpy change at Tm), ΔCp (constant-pressure heat capacity change) and ΔGDo (denaturational Gibbs energy change at 25 °C) of proteins in the presence of different GB concentrations. Another conclusion of this study is that ΔHm and ΔCp are not significantly changed in the presence of GB. This study suggests that other methylated glycine osmolytes may also behave in the same manner.  相似文献   

13.
Mixing characteristics of a laboratory scale internal loop air-lift fermenter has been investigated. The effects of different draft tube dimensions and positions as well as varying levels of liquid height over the draft tube, on mixing time were determined. The results indicate the existance of an optimum liquid height and thus liquid volume with respect to mixing performance especially for the taller draft tubes.List of Symbols A mm distance between draft tube and reactor base - A D mm2 area of the downcomer region - A R mm2 area of the riser region - B mm width of annulus - D d mm draft tube diameter - D t mm fermenter diameter - H d mm draft tube height - H l mm liquid height in the fermenter - H t mm fermenter height - V d m3 draft tube volume - V t m3 fermenter volume - D d /D l - B H d /H l - F H l /D t   相似文献   

14.
Thermal denaturation (Tm) data are easy to obtain; it is a technique that is used by both small labs and large‐scale industrial organizations. The link between ligand affinity (K D) and ΔTm is understood for reversible denaturation; however, there is a gap in our understanding of how to quantitatively interpret ΔTm for the many proteins that irreversibly denature. To better understand the origin, and extent of applicability, of a K D to ΔTm correlate, we define equations relating K D and ΔTm for irreversible protein unfolding, which we test with computational models and experimental data. These results suggest a general relationship exists between K D and ΔTm for irreversible denaturation.  相似文献   

15.
Raman spectra were measured for poly(L -histidine) in H2O, poly(L -histidine-d2 and -d3) in D2O, L -histidine in H2O, L -histidine-d3 (and d4) in D2O, and 4-methylimidazole in H2O with various pH (or pD) values. The Raman scattering peaks observed for these samples were ascribed to the neutral and positively charged imidazole groups on the basis of the spectral changes due to the pH variation and to the deuterium substitution of the imino protons. The vibrational modes of these peaks were deduced from the normal coordinate analysis made on the positively charged and neutral 4-ethylimidazoles. The Raman scattering peaks from the imidazole groups in the neutral form clearly indicate that these imidazole groups exist in the equilibrium between the two tautomeric forms, the 1-N protonated from (tautomer I) and the 3-N protonated one (tautomer II). For example, the breathing vibration of the 1-N protonated form is observed at 1282 cm?1 for L -histidine and at 1304 cm?1 for 4-methylimidazole, while the breathing vibration of the 3-N protonated form is observed at 1260 cm?1 for L -histidine and 4-methylimidazole. From the temperature dependence of the relative intensities of the tautomer I peak to that of the tautomer II, it was concluded that the tautomer I is energetically more stable than the tautomer II, and the ΔH value is 1.0 ± 0.3 kcal/mol for L -histidine and 0.4 ± 0.1 kcal/mol for 4-methylimidazole. Poly(L -histidine) with the neutral imidazole side chains shows the amide I peak at 1672 cm?1, indicating that the sample assumes the antiparallel pleated-sheet structure. Poly(L -Ala75L -His25) and poly(L -Ala50L -His50) were found to take the α-helical and β-form conformations, respectively.  相似文献   

16.
The stabilizing role of Trehalose, Sucrose, Sorbitol and Mannitol as sugar osmolytes and polyols on beta-lactoglobulin A (β-lgA) against its chemical denaturation at pH 2.0 and 25 °C has been explored by means of UV–vis spectroscopy. It has been observed that ΔG D o of β-lgA in the presence of 10% (w/v) Trehalose, Sucrose, Sorbitol and Mannitol is increased. We report that the functional dependence of ΔG D of proteins in the absence and the presence of sugar osmolytes on denaturant concentration is linear. Trehalose is found to induce remarkable stability of β-lgA against chemical denaturation. Furthermore, we assumed sugar osmolytes do not affect the secondary and tertiary structures of the native and GdnHCl-denatured states.  相似文献   

17.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  相似文献   

18.
A successful method for the purification of NADP isocitrate dehydrogenase from a plant source, Zea mays, is reported. Two mitochondrial isoenzymes were found and purified to homogeneity by a course of acetone fractionation, bulk exchange on DEAE-cellulose, cellulose hydroxylapatite column chromatography, and continuous elution electrophoresis. The mitochondrial isoenzymes are very similar with respect to kinetic properties, response to solvent perturbation, and temperature dependence of the pH/V relationship of isocitrate dehydrogenation. The Michaelis constant for isocitrate is identical for both isoenzymes. The enzymes have a molecular weight of 81,000 as estimated by permeation chromatography and an isoelectric point of 5.5 as extrapolated from gel-electrophoretic mobilities. Detectable differences are confined to differences in electrophoretic mobilities and heat denaturation. In D2O the rate of the overall reaction from isocitrate to α-ketoglutarate and CO2 was about 3.6 times slower than the same reaction in H2O. Both the forward and reverse reactions, in which isocitrate is dehydrogenated or generated from oxalosuccinate, were observed to decrease by this amount in D2O. The decarboxylation of oxalosuccinate was found to decrease by only about 25% in D2O relative to the velocity of the reaction in H2O. Thus the slow step in the overall reaction must be the initial dehydrogenation step rather than the decarboxylation of oxalosuccinate. The pK of the overall reaction did not change in D2O as compared to H2O.  相似文献   

19.
20.
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   

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