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.
In general, proteins fold with hydrophobic residues buried, away from water. Reversible protein folding due to hydrophobic interactions results from inverse temperature transitions where folding occurs on raising the temperature. Because homoiothermic animals constitute an infinite heat reservoir, it is the transition temperature, Tt, not the endothermic heat of the transition, that determines the hydrophobically folded state of polypeptides at body temperature. Reported here is a new hydrophobicity scale based on the values of Tt for each amino acid residue as a guest in a natural repeating peptide sequence, the high polymers of which exhibit reversible inverse temperature transitions. Significantly, a number of ways have been demonstrated for changing Tt such that reversibly lowering Tt from above to below physiological temperature becomes a means of isothermally and reversibly driving hydrophobic folding. Accordingly, controlling Tt becomes a mechanism whereby proteins can be induced to carry out isothermal free energy transduction.  相似文献   

2.
Metabolism is the conversion of available energy sources to those energy forms required for sustaining and propagating living organisms; this is simply biological energy conversion. Proteins are the machines of metabolism; they are the engines of motility and the other machines that interconvert energy forms not involving motion. Accordingly, metabolic engineering becomes the use of natural protein-based machines for the good of society. In addition, metabolic engineering can utilize the principles, whereby proteins function, to design new protein-based machines to fulfill roles for society that proteins have never been called upon throughout evolution to fulfill. This article presents arguments for a universal mechanism whereby proteins perform their diverse energy conversions; it begins with background information, and then asserts a set of five axioms for protein folding, assembly, and function and for protein engineering. The key process is the hydrophobic folding and assembly transition exhibited by properly balanced amphiphilic protein sequences. The fundamental molecular process is the competition for hydration between hydrophobic and polar, e.g., charged, residues. This competition determines Tt, the onset temperature for the hydrophobic folding and assembly transition, Nhh, the numbers of waters of hydrophobic hydration, and the pKa of ionizable functions. Reported acid-base titrations and pH dependence of microwave dielectric relaxation data simultaneously demonstrate the interdependence of Tt, Nhh and the pKa using a series of microbially prepared protein-based poly(30mers) with one glutamic acid residue per 30mer and with an increasing number of more hydrophobic phenylalanine residues replacing valine residues. Also, reduction of nicotinamides and flavins is shown to lower Tt, i.e., to increase hydrophobicity. Furthermore, the argument is presented, and related to an extended Henderson-Hasselbalch equation, wherein reduction of nicotinamides represents an increase in hydrophobicity and resulting hydrophobic-induced pKa shifts become the basis for understanding a primary energy conversion (proton transport) process of mitochondria. Copyright 1998 John Wiley & Sons, Inc.  相似文献   

3.
D W Urry  S Q Peng  T M Parker 《Biopolymers》1992,32(4):373-379
Three polypentapeptides--poly[0.8(GVGVP), 0.2(GEGVP)], poly[0.8(GVGIP), 0.2(GEGIP)], and poly[0.75(GFGVP), 0.25(GEGVP)]--all analogues of the polypentapeptide of elastin--(Val1-Pro2-Gly3-Val4-Gly5)n or poly(VPGVG)--have been prepared to determine the effect of changing the hydrophobicity, i.e., Val1----Ile1 (I) and Val4----Phe4 (F), on the pKa and the temperature dependence of pKa of the Glu (E) residue. Shifts in pKa as large as 1.7 units are observed and the temperature dependence is much steeper for the structure-dependent proximity of the more hydrophobic Ile1 residues to the Glu4 residue. Even though this system is dominated by the inverse temperature transition of hydrophobically driven folding on raising the temperature, the effect of adding 0.15 N NaCl is to suppress the hydrophobicity-induced pKa shift.  相似文献   

4.
TMDSC data have been employed to observe the effect of NaCl on the inverse temperature transition of the model elastin-like polymer (GVGVP)251. NaCl causes a decrease in Tt and an increase in DeltaH. The increase in enthalpy appears both in the enthalpy related with the folding of the polymer and in the contribution associated with disruption of the structured water of hydrophobic hydration. It has been suggested that the presence of NaCl may cause a better formation of water structures surrounding the apolar polymer chains.  相似文献   

5.
Poly[15(IPGVG),(RGYSLG)], where RGYSLG is a protein kinase site, was synthesized. On raising the temperature of a 5 mg/ml solution, this polypeptide undergoes an inverse temperature transition at 18 degrees C in which it folds into a contracted state by optimizing intramolecular hydrophobic interactions. Averaging the data of five experiments, phosphorylation by means of a 3':5' cyclic AMP dependent protein kinase to the extent of one phosphate in 360 residues raises the temperature of the folding transition to 32 degrees C. The shift is completely reversed on dephosphorylation by alkaline phosphatase. Phosphorylation is hereby shown to be the most potent chemical perturbation known for shifting the temperature of an inverse temperature transition, which has been shown to be an efficient mechanism for achieving chemomechanical transduction (mechanochemical coupling).  相似文献   

6.
Dimer-tetramer assembly reactions of the 10 CN-met ligation microstates of hemoglobin (Hb) were analyzed as a function of NaCl concentration while maintaining constant water activity by the addition of compensating sucrose. The assembly free energy for fully ligated cyanomet Hb and for fully oxygenated Hb becomes less favorable by 1.8 kcal when [NaCl] is increased from 0.08 to 0.7 M, whereas that of unligated Hb is practically insensitive to changes in [NaCl]. Values of 1.6 and 0.3 mol chloride release were found for the assembly of fully ligated and deoxy Hb, respectively; i.e., a net release of 1.3 mol chloride is coupled to the ligation of tetramers for both oxygen and cyanomet ligation. The ligation-linked salt component at constant water activity was evaluated to be 1.0 mol for the full oxygenation of the Hb tetramer in agreement with the overall value previously reported. When the detailed salt linkages accompanying all 16 stepwise cyanomet ligation reactions were experimentally resolved, only two "chloride" effects were found. The first chloride effect correlates with the ligation steps, which create tertiary constraint, and the second effect is coupled to the six switchpoints of quaternary T-->R transition. The distribution of these chloride effects agrees closely with predictions of the "symmetry rule mechanism." The total chloride release for CN-met ligation is in good agreement with that for oxygenation. Free energy contributions to assembly and cooperativity arising from the osmotic effects of chloride were found to be small for all ligation species.  相似文献   

7.
Guo W  Lampoudi S  Shea JE 《Proteins》2004,55(2):395-406
The temperature dependence of the free energy landscape of the src-SH3 protein domain is investigated through fully atomic simulations in explicit solvent. Simulations are performed above and below the folding transition temperature, enabling an analysis of both protein folding and unfolding. The transition state for folding and unfolding, identified from the free energy surfaces, is found to be very similar, with structure in the central hydrophobic sheet and little structure throughout the rest of the protein. This is a result of a polarized folding (unfolding) mechanism involving early formation (late loss) of the central hydrophobic sheet at the transition state. Unfolding simulations map qualitatively well onto low-temperature free energy surfaces but appear, however, to miss important features observed in folding simulations. In particular, details of the folding mechanism involving the opening and closing of the hydrophobic core are not captured by unfolding simulations performed under strongly denaturing conditions. In addition, free energy surfaces at high temperatures do not display a desolvation barrier found at lower temperatures, involving the expulsion of water molecules from the hydrophobic core.  相似文献   

8.
Elucidating the physiochemical principles that govern molecular self-assembly is of great importance for understanding biological systems and may provide insight into the emergence of the earliest macromolecules of life, an important challenge facing the RNA World hypothesis. Self-assembly results from a delicate balance between multiple noncovalent interactions and solvent effects, but achieving efficient self-assembly in aqueous solution with synthetic molecules has proven particularly challenging. Here, we demonstrate how two physical properties – monomer solubility and large hydrophobic surfaces of intermediate structures – are key elements to achieving supramolecular polymers in aqueous solution (Cafferty et al., 2013). Applying these two principles, we report the highly cooperative self-assembly of two weakly interacting, low molecular weight monomers [cyanuric acid and a modified triaminopyrimidine] into a water-soluble supramolecular assembly (see scheme below). The observed equilibrium between only two appreciably populated states – free monomers and supramolecular assemblies – is in excellent agreement with the values previously determined for the free energy of hydrogen bonding (Klostermeier & Millar, 2002), π???π stacking (Frier et al., 1985), and the calculated free energy penalty for the solvation of hydrophobic structures in water (Chandler, 2005). The similarity of the molecules used in this study for the nucleobases found in contemporary nucleic acids and the demonstration that these monomers assemble while the natural nucleobases do not, suggests that the first informational polymers may have emerged from a similar self-assembly process, if the nucleobases were different then they are today (Hud et al., 2013).  相似文献   

9.
M L Johnson 《Biochemistry》1988,27(2):833-837
Studies of the linkage between ligand binding and subunit assembly of oligomeric proteins have extensively used the concept of free energy coupling. The "order" of these free energy couplings was introduced [Weber, G. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 7098-7102] as the number of subunits that must be liganded to alter specific intersubunit interactions. This concept dictates that the ligation of fewer subunits has no effect, but once the order number of subunits becomes ligated, the specific intersubunit interaction energy between those particular subunits is completely eliminated. Weber's report claims that the free energy coupling between oxygen binding and the dimer-tetramer subunit assembly in stripped human hemoglobin A is "first order". This conclusion is based on the analysis of a set of previously published equilibrium constants [Mills, F. C., Johnson, M. L., & Ackers, G. K. (1976) Biochemistry 15, 5350-5362]. I subsequently reported that the original experimental data, from which the equilibrium constants were derived, are consistent with both the first-order and "second-order" free energy coupling concepts [Johnson, M. L. (1986) Biochemistry 25, 791-797]. I also demonstrated that more precise recent experimental data [Chu, A. H., Turner, B. W., & Ackers, G. K. (1984) Biochemistry, 23, 604-617] are consistent with both the first-order and second-order free energy coupling concepts. A recent article [Weber, G. (1987) Biochemistry 26, 331-332] disagrees that the oxygen-binding data for human hemoglobin A are consistent with a second-order model.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
In this study, we synthesized a polypeptide from its pentapeptide unit using microwave irradiation. Effective methods for polypeptide synthesis from unit peptides have not been reported. Here, we used a key elastin peptide, H‐GlyValGlyValPro‐OH (GVGVP), as the monomer peptide. It is difficult to obtain poly(Gly‐Val‐Gly‐Val‐Pro) (poly(GVGVP)) from the pentapeptide unit of elastin, GVGVP, via polycondensation. Poly(GVGVP) prepared from genetically recombinant Escherichia coli is a well‐known temperature‐sensitive polypeptide, and this temperature sensitivity is known as the lower critical solution temperature. When microwave irradiation was performed in the presence of various additives, the pentapeptide (GVGVP) polycondensation reaction proceeded smoothly, resulting in a product with a high molecular weight in a relatively good yield. The reaction conditions, like microwave irradiation, coupling agents, and solvents, were optimized to increase the reaction efficiency. The product exhibited a molecular weight greater than Mr 7000. Further, the product could be synthesized on a gram scale. The synthesized polypeptide exhibited a temperature sensitivity that was similar to that of poly(GVGVP) prepared from genetically recombinant E. coli. Therefore, this technique offers a facile and quick approach to prepare polypeptides in large amounts. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The driving force for folding of polypeptide chains into their threedimensional compact units has been designated as being hydrophobic and a measure of the hydrophobic character of the constituent amino acids has been determined by relative solubility measurements. It has been found however that the hydrophobic character of a protein is not sufficient to account for the complete stabilization of the tertiary structure of proteins. It is suggested that if the free energy of mixing of the hydrophobic side chains in the interior of the protein is added to the free energy of desolvation, i.e. the hydrophobic free energy, then the total free energy of mixing and desolvation can account for the known stability of the tertiary structure of proteins.  相似文献   

12.
The secondary structure of the alternating polydeoxynucleotide sequence poly[d(C-T)] was studied as a function of pH by ultraviolet absorbance and circular dichroism spectroscopy and by the analysis of UV-induced photoproducts. As the pH was lowered, poly[d(C-T)] underwent a conformational transition that was characterized by changes in the long-wavelength region (280-320 nm) of the CD spectrum. These changes have previously been interpreted as evidence for the formation of a core of stacked, protonated C X C+ base pairs in a double-helical complex of poly[d(C-T)], with the thymidyl residues being looped out into the solvent [Gray, D. M., Vaughan, M., Ratliff, R. L., & Hayes, F. N. (1980) Nucleic Acids Res. 8, 3695-3707]. In the present work, poly[d(C-T)] was labeled with [U-14C]cytosine and [methyl-3H]thymine and irradiated at pH values both above and below the conformational transition point (monitored by CD spectroscopy). The distribution of radioactivity in uracil means value of uracil dimers, uracil means value of thymine dimers (the deamination products of cytosine means value of cytosine and cytosine means value of thymine dimers, respectively), and thymine-means value of thymine dimers was then determined. As the pH was decreased, we found an increase in the yield of uracil means value of uracil dimers and a decrease in the yield of uracil means value of thymine dimers, which occurred concomitantly with the change in the CD spectrum.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Shelton VM  Sosnick TR  Pan T 《Biochemistry》1999,38(51):16831-16839
The equilibrium folding of a series of self-complementary RNA duplexes and the unmodified yeast tRNA(Phe) is studied as a function of urea and Mg(2+) concentration with optical spectroscopies and chemical modification under isothermal conditions. Via application of standard methodologies from protein folding, the folding free energy and its dependence on urea concentration, the m value, are determined. The free energies of the RNA duplexes obtained from the urea titrations are in good agreement with those calculated from thermal melting studies [Freier, S. I., et al. (1986) Proc. Natl. Acad. Sci. U.S.A. 83, 9373]. The m value correlates with the length of the RNA duplex and is not sensitive to ionic conditions and temperature. The folding of the unmodified yeast tRNA(Phe) can be described by two Mg(2+)-dependent transitions, the second of which corresponds to the formation of the native tertiary structure as confirmed by hydroxyl radical protection and partial nuclease digestion. Both transitions are sensitive to urea and have m values of 0.94 and 1.70 kcal mol(-)(1) M(-)(1), respectively. Although the precise chemical basis of urea denaturation of RNA is uncertain, the m values for the duplexes and tRNA(Phe) are proportional to the amount of the surface area buried in the folding transition. This proportionality, 0.099 cal mol(-)(1) M(-)(1) A(-)(2), is very similar to that observed for proteins, 0.11 cal mol(-)(1) M(-)(1) A(-)(2) [Myers, J., Pace, N., and Scholtz, M. (1995) Protein Sci. 4, 2138]. These results indicate that urea titration can be used to measure both the free energy and the magnitude of an RNA folding transition.  相似文献   

14.
L Kelly  L A Holladay 《Biochemistry》1990,29(21):5062-5069
Differential scanning microcalorimetry (DSC) of horse, rat, opossum, raccoon, carp, and armadillo metmyoglobins at alkaline pH gave data that fit the two-state unfolding model well. Monte Carlo studies were used to assess the impact of truncating DSC scans on the reliability of the calculated results when aggregation exotherms overlapped the unfolding endotherm at the high-temperature end of the scan. The DSC estimates for the conformational free energy at pH 8 and 298 K are compared to earlier results from isothermal acid and guanidinium chloride unfolding. Stability estimates at pH 8 for these six metmyoglobins obtained by DSC experiments do not agree with free energy estimates at pH 8 from guanidinium chloride unfolding. This is true for all three models used to extrapolate the free energy change to 0 M guanidinium chloride. Among these six myoglobins, significant variation appears in the temperature at which the myoglobin is half-unfolded, in the change in heat capacity upon unfolding, and in the change in enthalpy at 310 K. Calculations made with the hydrophobic model for protein folding [Baldwin, R.L. (1986) Proc. Natl. Acad. Sci. U.S.A. 83, 8069] suggest that a sizable variation exists for that portion of the unfolding enthalpy change assigned to forces other than the hydrophobic effect.  相似文献   

15.
We recently introduced a physical model [T.X. Hoang, A. Trovato, F. Seno, J.R. Banavar, A. Maritan, Geometry and symmetry pre-sculpt the free energy landscape of proteins. Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 7960-7964, J.R. Banavar, T.X. Hoang, A. Maritan, F. Seno, A. Trovato, A unified perspective on proteins-a physics approach. Phys. Rev., E 70 (2004) 041905] for proteins which incorporates, in an approximate manner, several key features such as the inherent anisotropy of a chain molecule, the geometrical and energetic constraints placed by the hydrogen bonds and sterics, and the role played by hydrophobicity. Within this framework, marginally compact conformations resembling the native state folds of proteins emerge as broad competing minima in the free energy landscape even for a homopolymer. Here we show how the introduction of sequence heterogeneity using a simple scheme of just two types of amino acids, hydrophobic (H) and polar (P), and sequence design allows a selected putative native fold to become the free energy minimum at low temperature. The folding transition exhibits thermodynamic cooperativity, if one neglects the degeneracy between two different low energy conformations sharing the same fold topology.  相似文献   

16.
Understanding the factors influencing the folding rate of proteins is a challenging problem. In this work, we have analyzed the role of non-covalent interactions for the folding rate of two-state proteins by free-energy approach. We have computed the free-energy terms, hydrophobic, electrostatic, hydrogen-bonding and van der Waals free energies. The hydrophobic free energy has been divided into the contributions from different atoms, carbon, neutral nitrogen and oxygen, charged nitrogen and oxygen, and sulfur. All the free-energy terms have been related with the folding rates of 28 two-state proteins with single and multiple correlation coefficients. We found that the hydrophobic free energy due to carbon atoms and hydrogen-bonding free energy play important roles to determine the folding rate in combination with other free energies. The normalized energies with total number of residues showed better results than the total energy of the protein. The comparison of amino acid properties with free-energy terms indicates that the energetic terms explain better the folding rate than amino acid properties. Further, the combination of free energies with topological parameters yielded the correlation of 0.91. The present study demonstrates the importance of topology for determining the folding rate of two-state proteins.  相似文献   

17.
Under the free energy landscape theory, both the protein-folding and protein–ligand binding processes are driven by the decrease in total Gibbs free energy of the protein-solvent or protein–ligand-solvent system, which involves the non-complementary changes between the entropy and enthalpy, ultimately leading to a global free energy minimization of these thermodynamic systems (Ji & Liu, 2011; Liu et al., 2012; Yang, Ji & Liu, 2012). In the case of protein folding, the lowering of the system free energy coupled with the gradual reduction in conformational degree of freedom of the folding intermediates determines that the shape of the free energy landscape for protein folding must be funnel-like (Dill & Chan, 1997), rather than non-funneled shapes (Ben-Naim, 2012). In the funnel-like free energy landscape, protein folding can be viewed as going down the hill via multiple parallel routes from a vast majority of individual non-native states on surface outside the funnel to the native states located around the bottom of the funnel. The first stage of folding, i.e. the rapid hydrophobic collapse process, is driven by the solvent entropy maximization. Concretely, the water molecules squeeze and sequestrate the hydrophobic amino acid side chains within the interior of the folding intermediates while exposing the polar and electrostatically charged side chains on the intermediate surface so as to minimize the solvent-accessible surface area of the solute and thus, the minimal contacts between the folding intermediates and the water molecules. This will maximize the entropy of the solvent, thus contributing substantially to lowering of the system free energy due to an absolute advantage of the solvent in both quantity and mass (Yang, Ji & Liu, 2012). The resulting molten globule states (Ohgushi & Wada, 1983), within which a few transient secondary structural components and tertiary contacts have been formed but many native contacts or close residue–residue interactions has yet to form, need to be further sculptured into the native states. This is a relatively slow “bottleneck” process because the competitive interactions between protein residues within the folding intermediates and between residues and water molecules may repeat many rounds to accumulate a large enough number of stable noncovalent bonds capable of counteracting the conformational entropy loss of the intermediates, thus putting this bottleneck stage under the enthalpy control (i.e. negative enthalpy change), contributing further to the lowering of the system free energy. Although the protein–ligand association occurs around the rugged bottom of the free energy landscape, the exclusion of water from the binding interfaces and the formation of noncovalent bonds between the two partners can still lower the system free energy. In conjunction with the loss of the rotational and translational degrees of freedom of the two partners as well as the loss of the conformational entropy of the protein, these processes could merge, downwards expand, and further narrow the free energy wells within which the protein–ligand binding process takes place, thereby making them look like a funnel, which we term the binding funnel. In this funnel, the free energy downhill process follows a similar paradigm to the protein-folding process. For example, if the initial collisions/contacts occur between the properly complementary interfaces of the protein and ligand, a large amount of water molecules (which usually form a water network around the solute surface) will be displaced to suit the need for maximizing the solvent entropy. This process is similar to that of the hydrophobic collapse during protein folding, resulting in a loosely associated protein–ligand complex that needs also to be further adapted into a tight complex, i.e. the second step which is mainly driven by the negative enthalpy change through intermolecular competitive interactions to gradually accumulate the noncovalent bonds and ultimately, to stabilize the complex at a tightly bound state. Taken together, we conclude that whether in the protein-folding or in the protein–ligand binding process, both the entropy-driven first step and the enthalpy-driven second step contribute to the lowering of the system free energy, resulting in the funnel-like folding or binding free energy landscape.  相似文献   

18.
The folding kinetics of a three-stranded antiparallel beta-sheet (WW domain) have been measured by temperature jump relaxation. Folding and activation free energies were determined as a function of temperature for both the wild-type and the mutant domain, W39F, which modifies the beta(2)-beta(3) hydrophobic interface. The folding rate decreases at higher temperatures as a result of the increase in the activation free energy for folding. Phi-Values were obtained for thermal perturbations allowing the primary features of the folding free energy surface to be determined. The results of this analysis indicate a significant shift from an "early" (Phi(T)=0. 4) to a "late" (Phi(T)=0.8) transition state with increasing temperature. The temperature-dependent Phi-value analysis of the wild-type WW domain and of its more stable W39F hydrophobic cluster mutant reveals little participation of residue 39 in the transition state at lower temperature. As the temperature is raised, hydrophobic interactions at the beta(2)-beta(3) interface gain importance in the transition state and the barrier height of the wild-type, which contains the larger tryptophan residue, increases more slowly than the barrier height of the mutant.  相似文献   

19.
During the synthesis of integral membrane proteins (IMPs), the hydrophobic amino acids of the polypeptide sequence are partitioned mostly into the membrane interior and hydrophilic amino acids mostly into the aqueous exterior. Using a many-body statistical mechanics model, we analyze the minimum free energy state of polypeptide sequences partitioned into α-helical transmembrane (TM) segments and the role of thermal fluctuations. Results suggest that IMP TM segment partitioning shares important features with general theories of protein folding. For random polypeptide sequences, the minimum free energy state at room temperature is characterized by fluctuations in the number of TM segments with very long relaxation times. Moreover, simple assembly scenarios do not produce a unique number of TM segments due to jamming phenomena. On the other hand, for polypeptide sequences corresponding to actual IMPs, the minimum free energy structure with the wild-type number of segments is free of number fluctuations due to an anomalously large gap in the energy spectrum. Now, simple assembly scenarios do reproduce the minimum free energy state without jamming. Finally, we find a threshold number of random point mutations where the size of the anomalous gap is reduced to the point that the wild-type ground state is destabilized and number fluctuations reappear.  相似文献   

20.
Commonly a key element enabling proteins to function is an amino acid residue or residues with functional side chains having shifted pKa values. This article reports the results on a set of protein-based polymers (model proteins) that exhibit hydrophobic folding and assembly transitions, and that have been designed for the purpose of achieving large hydrophobic-induced pKa shifts by selectively replacing Val residues by Phe residues. The high molecular weight polypentapeptides, actually poly (tricosapeptides) with six varied pentamers in fixed sequence, were designed and synthesized to have the same amino acid compositions but different proximities between a single aspartic acid residue and 5 Phe residues per 30 residues. With the 5 Phe residues distal from the Asp residue, the observed pKa shift was 2.9 when compared to the Val-containing reference. With the 5 Phe residues within 1 nm of the Asp residue, the pKa shift was 6.2. This represents a free energy of interaction of 8 kcal/moles. To our knowledge, this is the largest pKa shift documented for an Asp residue in a polypeptide– or protein–water system. Data are reviewed that do not support the usual electrostatic arguments for pKa shifts of charge–charge repulsion and/or unfavorable ion self-energies arising from displacement of water by hydrophobic moieties, but rather the data are interpreted to indicate the presence of an apolar–polar repulsive free energy of hydration, which results from a potentially highly cooperative competition between apolar and polar species for hydration. © 1994 John Wiley & Sons, Inc.  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