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.
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   

2.
Five kinds of immobilized invertases (IMI)—covalently of porous glass and ion-exchange resins and ionically on ion-exchange resins—have been prepared and their kinetic characteristics for sucrose hydrolysis, such as Km, K, pH profile, and thermal stability were studied. Comparing the values of Km and activation energy and the entropy of IMI with those of native invertase, it was concluded that the immobilization influences not binding but kinetic specificity. The effects of the immobilization method on thermal stability were also discussed.  相似文献   

3.
In applied statistics it is customary to have to obtain a one‐ or two‐tail confidence interval for the difference d = p2p1 between two independent binomial proportions. Traditionally, one is looking for a classic and non‐symmetric interval (with respect to zero) of the type d ∈ [δLU], d ≤ δ0 or d ≥ δ0. However, in clinical trials, equivalence studies, vaccination efficacy studies, etc., and even after performing the classic homogeneity test, intervals of the type |d| ≤ Δ0 or |d| ≥ Δ0, where Δ0 > 0, may be necessary. In all these cases it is advisable to obtain the interval by inverting the appropriate test. The advantage of this procedure is that the conclusions obtained using the test are compatible with those obtained using the interval. The article shows how this is done using the new exact and asymptotic unconditional tests published. The programs for performing these tests may be obtained at URL http://www.ugr.es/~bioest/software.htm.  相似文献   

4.
Helix–coil transition of poly(γ-methyl-L -glutamate), poly(γ-ethyl-L -glutamate), and poly(γ-benzyl-L -glutamate) has been studied in mixed solvents by calorimetry, polarimetry, and viscometry. The experimental data have allowed the evaluation of solvation enthalpy Δhb, equilibrium constant K for hydrogen bond formation between the active solvent component and CO and NH groups, and the cooperativity parameter σ. The conformational transition of polypeptides in solution in a mixed solvent containing enough active solvent to maintain the coiled conformation has been produced by dilution with the helix-supporting solvent for the measurements of enthalpy of transition Δhs. The average value for Δhs is 3550 ± 300 J/mol and is practically independent of the nature of the side chain for the dichloroacetic acid-ethylene dichloride solvent pair at 25°C. A noticeable concentration effect exists in the case of poly(γ-benzyl-L -glutamate). The helical conformation is less stable for poly(γ-ethyl-L -glutamate), and this is explained by a steric effect hindering the access of dichloroacetic acid to side chains. Constant K has been calculated using polarimetric data and also from values of Δhs obtained at different temperatures using the Bixon and Lifson theory on the one hand and that of Sayama and coworkers on the other hand. Values of σ for poly(γ-ethyl-L -glutamate) have been calculated according to both theories mentioned, and the results show that the two sets of values are quite similar. The constant σ depends on the nature of the active solvent, on temperature, and on the binary-solvent composition. These conclusions are confirmed by viscometric results. Values of Δhb calculated from constant K are 5230 J/mol when Bixon and Lifson theory is used and 5569 J/mol when the theory at Sayama and coworkers is used. In both cases the value for Δhb is much lower than that of an intramolecular hydrogen bond. Experimental results suggest that the solvation mechanism would proceed in a manner so that mechanisms described in both theories are involved.  相似文献   

5.
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   

6.
We report here that Tyrophagus similis and Tyrophagus putrescentiae (Astigmata: Acaridae) have the ability to biosynthesize linoleic acid [(9Z, 12Z)-9, 12-octadecadienoic acid] via a Δ12-desaturation step, although animals in general and vertebrates in particular appear to lack this ability. When the mites were fed on dried yeast enriched with d31-hexadecanoic acid (16:0), d27-octadecadienoic acid (18:2), produced from d31-hexadecanoic acid through elongation and desaturation reactions, was identified as a major fatty acid component of phosphatidylcholines (PCs) and phosphatidylethanolamines (PEs) in the mites. The double bond position of d27-octadecadienoic acid (18:2) of PCs and PEs was determined to be 9 and 12, respectively by dimethyldisulfide (DMDS) derivatization. Furthermore, the GC/MS retention time of methyl 9, 12-octadecadienoate obtained from mite extracts agreed well with those of authentic linoleic acid methyl ester. It is still unclear whether the mites themselves or symbiotic microorganisms are responsible for inserting a double bond into the Δ12 position of octadecanoic acid. However, we present here the unique metabolism of fatty acids in the mites.  相似文献   

7.
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   

8.
Helical conformations of infinite polymer chains may be described by the helical parameters, d and θ (the translation along the helix axis and the angle of rotation about the axis per repeat unit), pi (the distance of the ith atom from the axis), dij, and dij (the translation along the axis and the angle of rotation, respectively, on passing from the ith atom to the jth). A general method has been worked out for calculating all those helical parameters from the bond lengths, bond angles, and internal-rotation angles. The positions of the main chain and side chain atoms with respect to the axis may also be calculated. All the equations are applicable to any helical polymer chain and are readily programmed for electronic computers. A method is also presented for calculating the partial derivatives of helical parameters with respect to molecular parameters.  相似文献   

9.
The low resolution structure of subunit d (Vma6p) of the Saccharomyces cerevisiae V-ATPase was determined from solution X-ray scattering data. The protein is a boxing glove-shaped molecule consisting of two distinct domains, with a width of about 6.5 nm and 3.5 nm, respectively. To understand the importance of the N- and C-termini inside the protein, four truncated forms of subunit d (d 11–345, d 38–345, d 1–328 and d 1–298) and mutant subunit d, with a substitution of Cys329 against Ser, were expressed, and only d 11–345, containing all six cysteine residues was soluble. The structural properties of d depends strongly on the presence of a disulfide bond. Changes in response to disulfide formation have been studied by fluorescence- and CD spectroscopy, and biochemical approaches. Cysteins, involved in disulfide bridges, were analyzed by MALDI-TOF mass spectrometry. Finally, the solution structure of subunit d will be discussed in terms of the topological arrangement of the V1VO ATPase.  相似文献   

10.
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   

11.
Photosynthetic temperature acclimation could strongly affect coupled vegetation–atmosphere feedbacks in the global carbon cycle, especially as the climate warms. Thermal acclimation of photosynthesis can be modelled as changes in the parameters describing the direct effect of temperature on photosynthetic capacity (i.e., activation energy, Ea; deactivation energy, Hd; entropy parameter, ΔS) or the basal value of photosynthetic capacity (i.e., photosynthetic capacity measured at 25°C). However, the impact of acclimating these parameters (individually or in combination) on vegetative carbon gain is relatively unexplored. Here we compare the ability of 66 photosynthetic temperature acclimation scenarios to improve the ability of a spatially explicit canopy carbon flux model, MAESTRA, to predict eddy covariance data from a loblolly pine forest. We show that: (1) incorporating seasonal temperature acclimation of basal photosynthetic capacity improves the model's ability to capture seasonal changes in carbon fluxes and outperforms acclimation of other single factors (i.e., Ea or ΔS alone); (2) multifactor scenarios of photosynthetic temperature acclimation provide minimal (if any) improvement in model performance over single factor acclimation scenarios; (3) acclimation of Ea should be restricted to the temperature ranges of the data from which the equations are derived; and (4) model performance is strongly affected by the Hd parameter. We suggest that a renewed effort be made into understanding whether basal photosynthetic capacity, Ea, Hd and ΔS co‐acclimate across broad temperature ranges to determine whether and how multifactor thermal acclimation of photosynthesis occurs.  相似文献   

12.
Temperature dependence of CD spectra of DNA from various sources   总被引:1,自引:0,他引:1  
The CD spectra of DNA from various sources (T2; T4; Cd; Escherichia coli; calf thymus; Streptomyces chrysomalis) were investigated. A new band Δε210 in the CD spectra of glucosylated DNA of the T even phages was found. The temperature dependence of the CD spectra of DNA was obtained over a wide range of temperatures, including those of the helix–coil transition. The band Δε275 for all DNAs does not appreciably change in the range of the helix–coil transition. The monotonic increase of this band before melting, and its decrease after melting is observed with an increase in temperature. The amplitude of the CD band Δε245 for all the DNAs studied and Δε210 (glucosylated DNA) parallels the change of E260 absorbance.  相似文献   

13.
Cyclic hexapeptides of the type cyclo(L -Xxx-L -Pro-D -Yyy)2 or cyclo(L -Xxx-L -Pro-Gly)2 exist in solution predominantly in two forms of C2 average symmetry, one with all-trans peptide bonds and generally well-established conformation, and another with both Xxx-Pro peptide bonds cis. We have been measuring the thermodynamic parameters of this equilibrium using carbon and proton nmr spectroscopy. Data have been obtained for peptides in which Yyy = Gly, D -Ala, or D -Phe, and Xxx = Gly, L -Ala, L -Leu, and L -Val. In a given solvent, stability of the all-trans form decreases (ΔG0 increases) as Xxx is changed through the series Gly, L -Ala-, L -Leu, and L -Val, consistent with expected increasing repulsion between the Xxx side chain and the proline δ methylene across the trnas Xxx-Pro bond. Also, for a given set of side chains, the stability of the all-trnas form increases as the polarity of the solvent decreases, consistent with models in which all C?O and N? H groups are accessible for solvation in the two-cis form, but two C?O and two N? H groups are somewhat sequestered in the all-trans form. With the available data it is not possible to identify pure intramolecular (solvent-independent) or pure peptide-bond solvation (side chain-independent) terms in ΔH° or ΔS°, although trends are discernible.  相似文献   

14.
Conformational studies of poly-L-alanine in water   总被引:5,自引:0,他引:5  
The conformational properties of poly-L -alanine have been examined in aqueous solutions in order to investigate the influence of hydrophobic interactions on the helix–random coil transition. Since water is a poor solvent for poly-L -alanine, water-soluble copolymers of the type (D , L -lysine)m–(L alanine)n-(D , L -lysine)m, having 10, 160, 450, and 1000 alanyl residues, respectively, in the central block, were synthezised. The optical rotatory dispersion of the samples was investigated in the range 190–500 mμ, and the rotation at 231 mμ was related to the α-helix content, θH, of the alanine section. In salt-free solutions, at neutral pH, the three large polymers show high θH values, which are greatly reduced when the temperature is increased from 5 to 80°C. No helicity was observed for the small (n = 10) polymer. By applying the Lifson-Roig theory, the following parameters were obtained for the transition of a residue from a coil to a helical state: ν = 0.012; ΔH = ?190 ± 40 cal./mole; ΔS = ?0.55 ± 0.12 e.u. Since ΔH and ΔS differ from the values expected for a process involving only the formation of a hydrogen bond, and in a manner predicted by theories for the influence of hydrophobic bonding on helix stability, it is concluded that a hydrophobic interaction is also involved. In the presence of salt (0.2M NaCl), or when the ε-amino groups of the lysyl residues are not protonated (pH = 12), the helical form of the two large polymers (n = 450 and n = 1000) is more stable than in water. Since the electrostatic repulsion between the lysine end blocks is greatly reduced under these conditions, the alanine helical sections fold back on themselves, and this conformation is stabilized by interchain hydrophobia bonds. This structure was predicted by the theory for the equilibrium between such interacting helices, non-interacting helices, and the random coil.  相似文献   

15.
The conformational properties of block copolymers of poly-L -leucine in water have been examined. The degree of polymerization of the poly-L -leucine block was 11 and 21, respectively, for samples prepared by the Merrifield procedure, and 56 for a sample prepared by the polymerization of leucine N-carboxyanhydride. The optical rotatory dispersion parameter b0 was used to obtain the helix content θh at various temperatures. Application of the Lifson-Roig theory gave the following parameters for the transition of a residue from a coil to a helical state: v = 0.05–0.011, ΔH = +100 cal/mole, ΔS = +0.70–1.00 e. u. These parameters, as well as those for other polyamino acids, are accounted for by hydrophobic bonds involving the nonpolar side chains in the helical and randomly coiled forms. From the data for poly-L -alanine and theoretical values of the thermodynamic parameters for hydrophobic bond formation, the parameters for formation of a polyglycine helix are computed. By separating the contributions of the backbone, it is possible to obtain a set of thermodynamic parameters for the side-chain contributions of a number of polyamino acids. Increased size of the nonpolar side chain (with a larger contribution from hydrophobic bonding) makes a larger contribution to the stability of the α-helix which is reflected, among other ways, in a higher helix content at given temperature.  相似文献   

16.
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   

17.
A sequence alignment of yeast cytochrome-c (y-cyt-c) with mammalian cyts-c shows that the yeast protein has a five residue long N-terminal extension. A question arises: Does this N-terminal extension play any roles in the stability, structure, and folding of the yeast protein? To answer this question, in silico and in vitro studies were carried out on the wild type (WT) protein and its five deletants (Δ(?5/?5), Δ(?5/?4), Δ(?5/?3), Δ(?5/?2), and Δ(?5/?1) where Δ denotes the deletion and the numbers refer to the residues deleted, e.g. Δ(?5/?1) denotes the deletion of residues numbered from ?5 to ?1 (TEFKA), while Δ(?5/?2) denotes the deletion of resides numbered from ?5 to ?2 (TEFK) and so on). The main conclusion of the in silico study is that the order of stability of deletants and WT protein is Δ(?5/?4) > WT > Δ(?5/?3) > Δ(?5/?5) > Δ(?5/?1) ~ Δ(?5/?2). In vitro studies involved (i) measurements of thermodynamic stability of all proteins by differential scanning calorimetry and from sigmoidal curves of two different structural properties ([θ]222, a probe for detecting change in secondary structure, and Δε405, a probe for detecting alteration in the heme environment), and (ii) characterization of all proteins by various spectral properties. The main conclusions of the in vitro studies are as follows: (i) The order of thermodynamic stability of all proteins is in excellent agreement with that predicted by in silico studies, and (ii) A sequential deletion of the N-terminal extension has no effects on protein structure and folding.  相似文献   

18.
An aminopeptidase that has peptide bond formation activity was identified in the cell-free extract of carpophore of Pleurotus eryngii. The enzyme, redesignated as eryngase, was purified for homogeneity and characterized. Eryngase had a molecular mass of approximately 79 kDa. It showed somewhat high stability with respect to temperature and pH; it was inhibited by iodoacetate. Among hydrolytic activities toward aminoacyl-p-nitroanilides (aminoacyl-pNAs), eryngase mainly hydrolyzed hydrophobic l-aminoacyl-pNAs and exhibited little activity toward d-Ala-pNA and d-Leu-pNA. In terms of peptide bond formation activity, eryngase used various aminoacyl derivatives as acyl donors and acceptors. The products were all dipeptidyl derivatives. Investigation of time dependence on peptide synthesis revealed that some peptides that are not recognized as substrates for hydrolytic activity of eryngase could become good targets for synthesis. Furthermore, eryngase has produced opioid dipeptides––l-kyotorphin (l-Tyr-l-Arg) and d-kyotorphin (l-Tyr-d-Arg)––using l-Tyr-NH2 and d- and l-Arg-methyl ester respectively as an acyl donor and acceptor. Yield evaluation of kyotorphin synthesis indicated that the conversion ratio of substrate to kyotorphin was moderate: the value was estimated as greater than 20%.  相似文献   

19.
Consider the model yijk=u ± ai ± bi ± cij ± eijk i=1, 2,…, t; j=1, 2,…b; k=1, 2,…,nij where μ is a constant and ai, bi, cij are distributed independently and normally with zero means and variances Δ2 Δ2/bdij and δ2 respectively. It is assumed that di's, and dij's are known (positive) constants (for all i and j). In this paper procedures for estimating the variance components (Δ2, Δ2b and Δ2a) and for testing the hypothesis Hoc2c2 = y3 and Hoa2b2 = y4 (where y2, y3, and y4, are specified constants) are presented. A generalization for the mixed model case is discussed in the last section.  相似文献   

20.
Hydrostatic pressure has a vital role in the biological adaptation of the piezophiles, organisms that live under high hydrostatic pressure. However, the mechanisms by which piezophiles are able to adapt their proteins to high hydrostatic pressure is not well understood. One proposed hypothesis is that the volume changes of unfolding (ΔVTot) for proteins from piezophiles is distinct from those of nonpiezophilic organisms. Since ΔVTot defines pressure dependence of stability, we performed a comprehensive computational analysis of this property for proteins from piezophilic and nonpiezophilic organisms. In addition, we experimentally measured the ΔVTot of acylphosphatases and thioredoxins belonging to piezophilic and nonpiezophilic organisms. Based on this analysis we concluded that there is no difference in ΔVTot for proteins from piezophilic and nonpiezophilic organisms. Finally, we put forward the hypothesis that increased concentrations of osmolytes can provide a systemic increase in pressure stability of proteins from piezophilic organisms and provide experimental thermodynamic evidence in support of this hypothesis.  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