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 activity coefficients of glycylglycine in four aqueous electrolyte solutions (+NaCl, +NaBr, +KCl and +KBr) were obtained at 298.2 K. The mean ionic activity coefficient of the electrolyte in aqueous solutions containing the peptide was determined from measurements of the potential differences of a cation and an anion ion-selective-electrode, each vs. a double junction reference electrode. The results show that the nature of the anion has a major effect on the activity coefficients of glycylglycine. Comparison of activity coefficient data for glycylglycine with literature data for glycine, both in aqueous NaCl solutions, indicates that the effect of the electrolyte is larger for the peptide than for the amino acid. For the peptide, in all cases, the effect of the electrolyte is more important at low molalities of the electrolyte. The Wilson equation was used to correlate the activity coefficient data obtained. The correlation results were satisfactory for the region of concentrated electrolyte.  相似文献   

2.
C C Chen  Y Zhu  J A King  L B Evans 《Biopolymers》1992,32(10):1375-1392
Under physiological conditions, many polypeptide chains spontaneously fold into discrete and tightly packed three-dimensional structures. The folded polypeptide chain conformation is believed to represent a minimum Gibbs energy of the system, governed by the weak interactions that operate between the amino acid residues and between the residues and the solvent. A semiempirical molecular thermodynamic model is proposed to represent the Gibbs energy of folding of aqueous homopolypeptide systems. The model takes into consideration both the entropy contribution and the enthalpy contribution of folding homopolypeptide chains in aqueous solutions. The entropy contribution is derived from the Flory-Huggins expression for the entropy of mixing. It accounts for the entropy loss in folding a random-coiled polypeptide chain into a specific polypeptide conformation. The enthalpy contribution is derived from a molecular segment-based Non-Random Two Liquid (NRTL) local composition model [H. Renon and J. M. Prausnitz (1968) AIChE J., Vol. 14, pp. 135-142; C.-C. Chen and L. B. Evans (1986) AIChE J., Vol. 32, pp. 444-454], which takes into consideration of the residue-residue, residue-solvent, and solvent-solvent binary physical interactions along with the local compositions of amino acid residues in aqueous homopolypeptides. The UNIFAC group contribution method [A. Fredenslund, R. L. Jones, and J. M. Prausnitz (1975) AIChE J., 21, 1086-1099; A. Fredenslund, J. Gmehling, and P. Rasmussen (1977) Vapor-Liquid Equilibrium Using UNIFAC, Elsevier Scientific Publishing Company, Amsterdam], developed originally to estimate the excess Gibbs energy of solutions of small molecules, was used to estimate the NRTL binary interaction parameters. The model yields a hydrophobicity scale for the 20 amino acid side chains, which compares favorably with established scales [Y. Nozaki and C. Tanford (1971) Journal of Biological Chemistry, Vol. 46, pp. 2211-2217; E. B. Leodidis and T. A. Hatton (1990) Journal of Physical Chemistry, Vol. 94, pp. 6411-6420]. In addition, the model generates qualitatively correct thermodynamic constants and it accurately predicts thermodynamically favorable folding of a number of aqueous homopolypeptides from random-coiled states into alpha-helices. The model further facilitates estimation of the Zimm-Bragg helix growth parameter s and the nucleation parameter sigma for amino acid residues [B. H. Zimm and J. K. Bragg (1959) Journal of Chemical Physics, Vol. 31, pp. 526-535]. The calculated values of the two parameters fall into the ranges suggested by Zimm and Bragg.  相似文献   

3.
Transfer free energies (ΔGtr) of amino acids from water to aqueous electrolyte solutions have been determined from the solubility measurements, as a function of salt concentration at 298.15 K under atmospheric pressure. The investigated aqueous systems contain amino acids of zwitterionic glycine peptides: glycine (Gly), diglycine (Gly2), triglycine (Gly3), and tetraglycine (Gly4) and cyclic glycylglycine (c(GG)) with an electrolyte compound of potassium chloride (KCl), potassium bromide (KBr) or potassium acetate (KAc). The solubilities of glycine and diglycine in aqueous solution decrease with increasing the concentration of salts (salting-out effect), whereas those of triglycine and tetraglycine increase with increasing the concentration of salts (salting-in effect). Furthermore, salting-in effect was found in aqueous c(GG)/KBr system, while salting-out effect was observed in aqueous c(GG)/KCl or c(GG)/KAc system. The experimental results were used to estimate the transfer free energies (Δgtr) of the peptide backbone unit (–CH2CONH–) from water to the aqueous electrolyte solutions. We developed a new trail to determine the activity coefficients (γ) for aqueous and aqueous electrolyte solutions using an activity coefficient model, with which the total contribution of transfer free energy between solute and the solvent was calculated. We compared the difference between neglecting and using the activity coefficients term in predicting ΔGtr. Since the transfer free energy contribution is negative, interactions between the ionic salts and the peptide backbone unit of zwitterionic glycine peptides are favorable and thus the ionic salts destabilize these amino acids. It was also found that KBr stabilizes c(GG), whereas KCl and KAc destabilize c(GG). These results provide evidence for the existence of interactions between the amide unit and ionic salts, in aqueous solution, which may be of importance in maintaining protein structure as well as in protein–solute and protein–solvent interactions.  相似文献   

4.
The association constants of various amino acids or their derivatives (methyl esters, amides, etc.) with mononucleosides in aqueous solutions have been measured by using precise ultraviolet difference absorbance photometry. Some of the results are in agreement with those of the previous solubility experiments. The superiority of this ultraviolet absorbance method over the solubility experiments is that it can discriminate between stacking and hydrogen-bonding interactions. New types of specific interactions of some amino acids with nucleic acid bases by using a peptidyl carboxylate ion and another donor or acceptor in their side chains have been found using this technique.  相似文献   

5.
Absorption spectra and ESR of aqueous and aqueous/glyceric solutions of oxyhemoglobin exposed to UV radiation (250-400 nm) at 293 and 77 K in the presence of ascorbic acid have been analyzed. Vitamin C (5 x 10(-5) M) has been shown to exert a photoprotective effect with regard to oxyhemoglobin (2 x 10(-6) M) UV-irradiated with a dose of 0.86 x 10(5) J/m2 at 293 K. The photoprotective effect of ascorbic acid is also displayed after UV irradiation of frozen (77 K) aqueous/glyceric oxyhemoglobin solutions (2.53 x 10(-5) M). It is concluded that ascorbic acid can be a scavenger with respect to active UV-induced particles in protein systems, including O2-. and OH. Proposed is a mode of processes leading to UV inactivation of hemoprotein molecules.  相似文献   

6.
Water activity, pH and density of aqueous amino acids solutions   总被引:2,自引:0,他引:2  
The water activity, pH and density of some aqueous amino acid solutions were determined at 25 degrees C in three different types of solvents. Previous published experimental data on water activity and solubility of amino acids in aqueous solutions were used together with data from this work to test the applicability of a group contribution model. The activity coefficients were estimated by the UNIFAC-Larsen model combined with the Debye-Hückel equation, taking also into account the partial dissociation phenomena of species in solution. Interaction energies between the charged species Na(+) and Cl(-) and the specific groups of amino acids (COOH and NH(2)) were adjusted using experimental solubility data.  相似文献   

7.
Homology modeling of gastric H,K-ATPase based on the E2 model of sarcoplasmic reticulum Ca2+-ATPase (Toyoshima, C., and Nomura, H. (2002) Nature 392, 835-839) revealed the presence of a single high-affinity binding site for K+ and an E2 form-specific salt bridge between Glu820 (M6) and Lys791 (M5). In the E820Q mutant this salt bridge is no longer possible, and the head group of Lys791, together with a water molecule, fills the position of the K+ ion and apparently mimics the K+-filled cation binding pocket. This gives an explanation for the K+-independent ATPase activity and dephosphorylation step of the E820Q mutant (Swarts, H. G. P., Hermsen, H. P. H., Koenderink, J. B., Schuurmans Stekhoven, F. M. A. H., and De Pont, J. J. H. H. M. (1998) EMBO J. 17, 3029-3035) and, indirectly, for its E1 preference. The model is strongly supported by a series of reported mutagenesis studies on charged and polar amino acid residues in the membrane domain. To further test this model, Lys791 was mutated alone and in combination with other crucial residues. In the K791A mutant, the K+ affinity was markedly reduced without altering the E2 preference of the enzyme. The K791A mutation prevented, in contrast to the K791R mutation, the spontaneous dephosphorylation of the E820Q mutant as well as its conformational equilibrium change toward E1. This indicates that the salt bridge is essential for high-affinity K+ binding and the E2 preference of H,K-ATPase. Moreover, its breakage (E820Q) can generate a K+-insensitive activity and an E1 preference. In addition, the study gives a molecular explanation for the electroneutrality of H,K-ATPases.  相似文献   

8.
The screening of natural deep eutectic solvents (NADES) to identify those with the ability to strongly solvate rutin was conducted using the COSMO-RS methodology. A NADES model was constructed that took into account the possible ionic and neutral forms of its constituents. The distributions of all forms were computed based on the equilibrium constants of neutralization reactions between amino and carboxylic acids. The proposed model was validated against the experimental solubilities of 15 NADES. A linear relationship between these data and the estimated activity coefficient values was found. The screening encompassed 126 different NADES. It was found that ten of them outperformed the best reference system. The most effective two-component solvent comprised proline combined with 2,3-diaminosuccinic acid, and the solubility of rutin in this solvent was found to be 130% greater than its solubility in the best reference system. The amino acids associated with the highest rutin solubilities were all cyclic, and the use of carboxylic acids with two carboxyl groups and a main chain consisting of two methylene groups with two amino substituents was observed to yield the best rutin solubilities. Because of the acidic properties of rutin, the presence of basic sites on the components of the NADES generally leads to enhanced solubility.  相似文献   

9.
The activity coefficients at 298.15 K of glycine in water + NaNO(3) + glycine system and dl-methionine in water + NaCl + dl-methionine system are reported. The measurements were performed in an electrochemical cell with two ion selective electrodes, a cation and an anion ion selective electrode, each versus a double junction reference electrode. The concentrations of the electrolytes and the amino acids studied covered up to 1.0 molality electrolyte, 2.4 molality glycine and 0.2 molality dl-methionine. The results of the activity coefficients of glycine are compared with the activity coefficients of glycine in water + NaCl + glycine and water + KCl + glycine systems, obtained from the previous studies. The results show that the nature of both the cation and the anion of an electrolyte have significant effects on the activity coefficient of glycine in aqueous electrolyte solutions. The results also show that there are attractive interactions between the molecules of glycine and NaNO(3) and repulsive interactions between the molecules of dl-methionine and NaCl.  相似文献   

10.
Partial molar heat capacities (CP2 degrees) and volumes (V2 degrees) for some amino acids and peptides were measured in 1 M aqueous calcium chloride solutions at 298.15 degrees K using a Picker flow microcalorimeter and an oscillating-tube digital density meter. Using the data for these amino acids and peptides in water, the corresponding partial molar heat capacities of transfer (CP2,tr degree) and volumes (V2,tr degree) from water to 1 M aqueous calcium chloride were deduced. These thermodynamic parameters are significantly positive, indicating that strong interactions occur between the ions of calcium chloride and the charged centres of these amino acids and peptides. A comparison has been made with a similar transfer of these compounds to sodium chloride solutions. The thermodynamic parameters of the transfer of peptide group (-CONH) are much more positive in calcium chloride than in sodium chloride solutions. The implication of this result for the ability of calcium chloride to act as a stronger destabilizer of protein conformation is discussed.  相似文献   

11.
12.
The thermodynamics of the hydrolysis of lactose to glucose and galactose have been investigated using both high pressure liquid chromatography and heat-conduction microcalorimetry. The reaction was carried out over the temperature range 282-316 K and in 0.1 M sodium acetate buffer at a pH of 5.65 using the enzyme beta-galactosidase to catalyze the reaction. For the process lactose(aq) + H2O(liq) = glucose(aq) + galactose(aq), delta G0 = -8.72 +/- 0.20 kJ.mol-1, K0 = 34 +/- 3, delta H0 = 0.44 +/- 0.11 kJ.mol-1, delta S0 = 30.7 +/- 0.8 J.mol-1.K-1, and delta Cop = 9 +/- 20 J.mol-1.K-1 at 298.15 K. The standard state is the hypothetical ideal solution of unit molality. Thermochemical cycle calculations using enthalpies of combustion and solution, entropies, solubilities, activity coefficients, and apparent molar heat capacities have also been performed. These calculations indicate large discrepancies which are attributable primarily to errors in literature data on the enthalpies of combustion and/or third law entropies of the crystalline forms of the substrates.  相似文献   

13.
This mini-review summarizes my postdoctoral research in the labs of T. Wydrzynski/C.B. Osmond, J.H.A. Nugent/M.C.W. Evans and V.K. Yachandra/K. Sauer/M.P. Klein. The results are reported in the context of selected data from the literature. Special emphasis is given to the mode of substrate water binding, Mn oxidation states and the structures of the Mn cluster in the four (meta)stable redox states of the oxygen evolving complex. The paper concludes with a working model for the mechanism of photosynthetic water oxidation that combines mu-oxo bridge oxidation in the S(3) state (V.K. Yachandra, K. Sauer, M.P. Klein, Chem. Rev. 96 (1996) 2927-2950) with O-O bond formation between two terminal Mn-hydroxo ligands during the S(3)-->(S(4))-->S(0) transition.  相似文献   

14.
Amino acid-amino acid interaction energies have been derived from crystal structure data for a number of years. Here is reported the first derivation of normalized relative interaction from binding data for each of the four bases interacting with a specific amino acid, utilizing data from combinatorial multiplex DNA binding of zinc finger domains [Desjarlais, J. R. and Berg, J. M. (1994) Proc. Natl. Acad. Sci. USA, 91, 11099-11103]. The five strongest interactions are observed for lysine-guanine, lysine-thymine, arginine-guanine, aspartic acid-cytosine and asparagine-adenine. These rankings for interactions with the four bases appear to be related to base-amino acid partial charges. Also, similar normalized relative interaction energies are derived by using DNA binding data for Cro and lambda repressors and the R2R3 c-Myb protein domain [Takeda, Y., Sarai, A. and Rivera, V. M. (1989) Proc. Natl. Acad. Sci. USA, 86, 439-443; Sarai, A. and Takeda, Y. (1989) Proc. Natl. Acad. Sci. USA, 86, 6513-6517; Ogata, K. et al. (1995) submitted]. These energies correlate well with the combinatorial multiplex energies, and the strongest cases are similar between the two sets. They also correlate well with similar relative interaction energies derived directly from frequencies of bases in the bacteriophage lambda operator sequences. These results suggest that such potentials are general and that extensive combinatorial binding studies can be used to derive potential energies for DNA-protein interactions.  相似文献   

15.
Ongoing debate in the ribosome field has focused on the role of bound E-site tRNA and the Shine-Dalgarno-anti-Shine-Dalgarno (SD-aSD) interaction on A-site tRNA interactions and the fidelity of tRNA selection. Here we use an in vitro reconstituted Escherichia coli translation system to explore the reported effects of E-site-bound tRNA and SD-aSD interactions on tRNA selection events and find no evidence for allosteric coupling. A large set of experiments exploring the role of the E-site tRNA in miscoding failed to recapitulate the observations of earlier studies (Di Giacco, V., Márquez, V., Qin, Y., Pech, M., Triana-Alonso, F. J., Wilson, D. N., and Nierhaus, K. H. (2008) Proc. Natl. Acad. Sci. U.S.A. 105, 10715-10720 and Geigenmüller, U., and Nierhaus, K. H. (1990) EMBO J. 9, 4527-4533); the frequency of miscoding was unaffected by the presence of E-site-bound cognate tRNA. Moreover, our data provide clear evidence that the reported effects of the SD-aSD interaction on fidelity can be attributed to the binding of ribosomes to an unanticipated site on the mRNA (in the absence of the SD sequence) that provides a cognate pairing codon leading naturally to incorporation of the purported "noncognate" amino acid.  相似文献   

16.
An analysis of genes for the major two subunits of the membrane-associated ATPase from an acidothermophilic archaebacterium, Sulfolobus acidocaldarius, suggested that it belongs to a different ATPase family from the F1-ATPase (Denda, K., Konishi, J., Oshima, T., Date, T., and Yoshida, M. (1988) J. Biol. Chem. 263, 17251-17254). In the same operon of the above two genes we found a gene encoding a very hydrophobic protein of 101 amino acids (Mr = 10,362). A proteolipid was purified from the membranes of this bacteria in which partial amino acid sequences matched with the sequence deduced from the gene. Significant amino acid sequence homology and a similar hydropathy profile appeared when the sequence was compared with the 8-kDa proteolipid subunit of F0F1-ATPases. It is about 30 amino acids larger than the 8-kDa proteolipid and has a small (11-amino acid) repeat sequence. However, it is distinct from the 16-kDa proteolipid subunit of an eukaryotic vacuolar H+-ATPase (Mandel, M., Moriyama, Y., Hulmes, J.D., Pan, Y.-E., Nelson, H., and Nelson, N. (1988) Proc. Natl. Acad. Sci. U.S.A. 85,5521-5524).  相似文献   

17.
kappa-Conotoxin PVIIA (kappa-PVIIA), a 27-amino acid peptide with three disulfide cross-links, isolated from the venom of Conus purpurascens, is the first conopeptide shown to inhibit the Shaker K(+) channel (Terlau, H., Shon, K., Grilley, M., Stocker, M., Stühmer, W., and Olivera, B. M. (1996) Nature 381, 148-151). Recently, two groups independently determined the solution structure for kappa-PVIIA using NMR; although the structures reported were similar, two mutually exclusive models for the interaction of the peptide with the Shaker channel were proposed. We carried out a structure/function analysis of kappa-PVIIA, with alanine substitutions for all amino acids postulated to be key residues by both groups. Our data are consistent with the critical dyad model developed by Ménez and co-workers (Dauplais, M., Lecoq, A., Song, J. , Cotton, J., Jamin, N., Gilquin, B., Roumestand, C., Vita, C., de Medeiros, C., Rowan, E. G., Harvey, A. L., and Ménez, A. (1997) J. Biol. Chem. 272, 4802-4809) for polypeptide antagonists of K(+) channels. In the case of kappa-PVIIA, Lys(7) and Phe(9) are essential for activity as predicted by Savarin et al. (Savarin, P., Guenneugues, M., Gilquin, B., Lamthanh, H., Gasparini, S., Zinn-Justin, S., and Ménez, A. (1998) Biochemistry 37, 5407-5416); these workers also correctly predicted an important role for Lys(25). Thus, although kappa-conotoxin PVIIA has no obvious sequence homology to polypeptide toxins from other venomous animals that interact with voltage-gated K(+) channels, there may be convergent functional features in diverse K(+) channel polypeptide antagonists.  相似文献   

18.
In a previous work, the parameters of the statistical associating fluid theory (SAFT) equation of state for amino acids were determined by using the method developed. The solubility of amino acids in water was modeled. In this work, the SAFT equation of state has been applied to describe the solubility of amino acids in aqueous alcohol solutions. The systems include dl-alanine/ethanol/water, glycine/ethanol/water, dl-valine/ethanol/water, dl-serine/ethanol/water, glycine/1-propanol/water, glycine/2-propanol/water, l-alanine/2-propanol/water, l-leucine/ethanol/water. Binary interaction parameters between amino acid and alcohol are needed by the SAFT model to get good modeling results.  相似文献   

19.
Oligopeptide-mediated helix stabilization of peptides in hydrophobic solutions was previously found by NMR and CD spectroscopic studies. The oligopeptide included the hydrophobic amino acids found in its parent peptide and were interposed by relevant basic oracidic amino acids. The strength of the interactions depended on the amino acid sequences. However, no helix-stabilizing effect was seen for the peptides in phosphate buffer solution, because the peptides assumed a random-coil structure. In order to ascertain whether the helix-stabilizing effect of an oligopeptide on its parent peptide could operate in aqueous solution, model peptides EK17 (Ac-AEAAAAEAAAKAAAAKA-NH2) and IFM17 (Ac-AEAAAAEIFMKAAAAKA-NH2) that may assume an alpha-helix in aqueous solutions were synthesized. Interactions were examined between various oligopeptides (EAAAK, KAAAE, EIFMK, KIFME, KIFMK and EYYEE) and EK17 or IFM17 in phosphate buffer and in 80% trifluoroethanol (TFE)-20% H2O solutions by CD spectra. EAAAK had little effect on the secondary structures of EK17 in both buffer and TFE solutions, while KAAAE, which has the reverse amino acid sequence of EAAAK, had a marked helix-destabilizing effect on EK17 in TFE. EIFMK and KIFME were found to stabilize the alpha-helical structure of EK17 in phosphate buffer solutions, whereas KIFMK and EYYEE destabilized the alpha-helical structure of EK17. EIFMK and KIFME had no effect on IFM17, because unexpectedly, IFM17 had appreciable amounts of beta-sheet structure in buffer solution. It was concluded that in order for the helix-stabilizing (1) the model peptide, the alpha-helical conformation of which is to be stabilized, should essentially assume an alpha-helical structure by nature, and (2) the hydrophobicity of the side-chains of the oligopeptide should be high enough for the oligopeptide to perform stable specific side chain-side chain intermolecular hydrophobic interactions with the model peptide.  相似文献   

20.
Buffered aqueous two-phase systems are effective extraction systems for separating amphoteric hydrocarbons like, for example, polypeptides from aqueous phases. The design and basic engineering of such processes requires the knowledge of the liquid-liquid equilibrium. The study presented here aims to contribute to the development of methods to predict the partitioning of peptides in aqueous two-phase systems. Experimental results are reported for the partitioning of small amounts ( approximately 0.001 g solute per gram of solution) of low molecular combination peptides of glycine, L-glutamic acid, L-phenylalanine, and L-lysine (9 dipeptides, gly-glu, gly-phe, gly-lys, glu-gly, phe-gly, phe-glu, lys-gly, lys-glu, lys-phe; 7 tripeptides, gly-gly-phe, gly-phe-gly, glu-gly-phe, phe-gly-gly, lys-gly-lys, lys-glu-gly, lys-phe-lys) in aqueous two-phase systems of high molecular weight dextran (molecular weight about 500,000) and poly(ethylene glycol) (molecular weight about 6,000 and 35,000, respectively) in the presence of small amounts (about 0.05 mol/kg) of K2HPO4/KH2PO4 buffer at about 293 K. The new data are compared to predictions. Partition coefficients are predicted applying a group contribution excess Gibbs energy model. The model is an osmotic virial equation. It uses surface fractions to encounter for the probability of interactions between solutes. All model parameters were taken from the literature. They were determined exclusively from experimental data for the phase forming systems and for the partitioning of amino acids and their di- and tripeptides (containing only a single amino acid), but no experimental data for the partitioning of combinations peptides were used. In most cases predicted partition coefficients agree favourably with the experimental data.  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