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.
P L Poole  J L Finney 《Biopolymers》1984,23(9):1647-1666
Direct difference ir spectra are presented as a function of hydration for lysozyme and α-lactalbumin, and detailed sequential hydration molecular events identified. Despite the strong sequence homology between the two proteins, and their expected conformational similarity, the hydration behaviour of the polar groups is different for the two proteins. Using a Hill-type analysis, we conclude that the acid groups ionize and hydrate rapidly and noncooperatively in both proteins, consistent with the known (lysozyme) and postulated (α-lactalbumin) surface chemistry. The polar group hydration shows a clear cooperativity, which is quantitatively different in the two proteins. Complementary work suggests this cooperativity relates to a hydration-induced “loosening up” of the lysozyme conformation at about 55 mol water/mol protein. α-Lactalbumin appears to “open up” more easily for hydration than does lysozyme, consistent with its lower stability against thermal and acid denaturation.  相似文献   

2.
R P Rand  N Fuller  V A Parsegian  D C Rau 《Biochemistry》1988,27(20):7711-7722
It is now generally recognized that hydration forces dominate close interactions of lipid hydrophilic surfaces. The commonality of their characteristics has been reasonably established. However, differences in measured net repulsion, particularly evident when phosphatidylethanolamine (PE) and phosphatidylcholine (PC) bilayers are compared, suggest there exists a variety of behavior wider than expected from earlier models of hydration and fluctuation repulsion balanced by van der Waals attraction. To find a basis for this diverse behavior, we have looked more closely at measured structural parameters, degrees of hydration, and interbilayer repulsive forces for the lamellar phases of the following lipids: 1-palmitoyl-2-oleoyl-PE (POPE), egg PE, transphosphatidylated egg PE (egg PE-T), mono- and dimethylated egg PE-T (MMPE and DMPE), 1-stearoyl-2-oleoyl-PC (SOPC), and mixtures of POPE and SOPC. POPE and SOPC bilayers differ not only in their maximum degrees of hydration but also in the empirical hydration force coefficients and decay lengths that characterize their interaction. When mixed with POPE, SOPC effects sudden and disproportionate increases in hydration. POPE, egg PE, and egg PE-T differ in their degree of hydration, molecular area, and hydration repulsion. A single methylation of egg PE-T almost completely converts its hydration and bilayer repulsive properties to those of egg PC; little progression of hydration is seen with successive methylations. In order to reconcile these observations with the conventional scheme of balancing interbilayer hydration and fluctuation-enhanced repulsion with van der Waals attraction, it is necessary to relinquish the fundamental idea that the decay of hydration forces is a constant determined by the properties of the aqueous medium. Alternatively, one can retain that fundamental idea if one recognizes the possibility that polar group hydration has an attractive component to it. In the latter view, that attractive component originates from interbilayer hydrogen-bonded water bridges between apposing bilayer surfaces, arising from correlation of zwitterionic or other complementary polar groups or from factors that affect polar group solubility. The same Marcelja and Radic formalism that accounts so well for the repulsive component also leads to an estimate of the attractive one. We suggest that the full range of degrees of hydration and of interbilayer spacings observed for different neutral bilayers results in part from variable contributions of the attractive and repulsive hydration components.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
The hydration of polar and apolar groups can be explained quantitatively, via the random network model of water, in terms of differential distortions in first hydration shell water-water hydrogen bonding angle. This method of analyzing solute induced structural distortions of water is applied to study the ice-binding type III thermal hysteresis protein. The analysis reveals subtle but significant differences in solvent structuring of the ice-binding surface, compared to non-ice binding protein surface. The major differences in hydration in the ice-binding region are (i). polar groups have a very apolar-like hydration. (ii). there is more uniform hydration structure. Overall, this surface strongly enhances the tetrahedral, or ice-like, hydration within the primary hydration shell. It is concluded that these two specific features of the hydration structure are important for this surface to recognize, and preferentially interact with nascent ice crystals forming in liquid water.  相似文献   

4.
Yang C  Sharp KA 《Proteins》2005,59(2):266-274
The random network model of water quantitatively describes the different hydration heat capacities of polar and apolar solutes in terms of distortions of the water-water hydrogen bonding angle in the first hydration shell (Gallagher and Sharp, JACS 2003;125:9853). The distribution of this angle in pure water is bimodal, with a low-angle population and high-angle population. Polar solutes increase the high-angle population while apolar solutes increase the low-angle population. The ratio of the two populations quantifies the hydrophobicity of the solute and provides a sensitive measure of water structural distortions. This method of analysis is applied to study hydration of type I thermal hysteresis protein (THP) from winter flounder and three quadruple mutants of four threonine residues at positions 2, 13, 24, and 35. Wild-type and two mutants (VVVV and AAAA) have antifreeze (thermal hysteresis) activity, while the other mutant (SSSS) has no activity. The analysis reveals significant differences in the hydration structure of the ice-binding site. For the SSSS mutant, polar groups have a typical polar-like hydration, that is, more high-angle H-bonds than bulk water. For the wild-type and active mutants, polar groups have unusual, very apolar-like hydration, that is, more low-angle H-bonds than bulk water. This pattern of hydration was seen previously in the structurally distinct type III THPs (Yang & Sharp Biophys Chem 2004;109:137), suggesting for the first time a general mechanism for different THP classes. The specific shape, residue size, and clustering of both polar and apoler groups are essential for an active ice binding surface.  相似文献   

5.
Aggregation is an irreversible form of protein complexation and often toxic to cells. The process entails partial or major unfolding that is largely driven by hydration. We model the role of hydration in aggregation using “Dehydrons.” “Dehydrons” are unsatisfied backbone hydrogen bonds in proteins that seek shielding from water molecules by associating with ligands or proteins. We find that the residues at aggregation interfaces have hydrated backbones, and in contrast to other forms of protein–protein interactions, are under less evolutionary pressure to be conserved. Combining evolutionary conservation of residues and extent of backbone hydration allows us to distinguish regions on proteins associated with aggregation (non‐conserved dehydron‐residues) from other interaction interfaces (conserved dehydron‐residues). This novel feature can complement the existing strategies used to investigate protein aggregation/complexation. Proteins 2016; 84:254–266. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
H-bonding in protein hydration revisited   总被引:1,自引:0,他引:1  
H-bonding between protein surface polar/charged groups and water is one of the key factors of protein hydration. Here, we introduce an Accessible Surface Area (ASA) model for computationally efficient estimation of a free energy of water-protein H-bonding at any given protein conformation. The free energy of water-protein H-bonds is estimated using empirical formulas describing probabilities of hydrogen bond formation that were derived from molecular dynamics simulations of water molecules at the surface of a small protein, Crambin, from the Abyssinian cabbage (Crambe abyssinica) seed. The results suggest that atomic solvation parameters (ASP) widely used in continuum hydration models might be dependent on ASA for polar/charged atoms under consideration. The predictions of the model are found to be in qualitative agreement with the available experimental data on model compounds. This model combines the computational speed of ASA potential, with the high resolution of more sophisticated solvation methods.  相似文献   

7.
The random network model of water quantitatively describes the different hydration heat capacities of polar and apolar solutes in terms of differential distortions of the water-water hydrogen bonding angle in the first hydration shell. This method of hydration analysis is applied here to study the hydration of the wild type III thermal hysteresis protein from eel pout and three mutations at residue 16. Wild type and one mutant have full activity, the other two mutants have little or no anti-freeze (thermal hysteresis) activity. The analysis reveals significant differences in the hydration structure of the ice-binding site (centered on residue 16) among four proteins. For the A16T and A16Y mutants with reduced activity, polar groups have a typical polar-like hydration. For the wild type and mutant A16C with 100% of the wild type activity, polar groups have unusual, very apolar-like hydration. In the latter case, hydrating water molecules form a more ice-like pattern of hydrogen bonding on the ice-binding face, while in the former case water-water H-bonds are more distorted and more heterogenous. Overall, the binding surface of active protein strongly enhances the water tetrahedral structure, i.e. promotes ice-like hydration. It is concluded that the specific shape, residue size and clustering of both polar/apolar groups are essential for the binding surface to recognize, and preferentially interact with nascent ice crystals forming in liquid water.  相似文献   

8.
As a first step in the determination of the hydration scheme of small peptides, the hydration sites of the cyclic dipeptide c(l-Thr-l-His) have been determined by two empirical potential treatments. In the first approach the energy is calculated by using the “Caillet-Claverie's” potentials, including electrostatic, dispersion-repulsion and polarization contributions. In the second approach (EMPWI method), the energy is calculated by simplified treatment, taking into account the electrostatic interactions of a suitable charge distribution and the dispersion-repulsion contributions. In this study, only the crystalline conformation of the cyclic dipeptide is considered. The hydration sites determined can be classified in three groups: (a) bridging sites, in which water interacts with both side chains, (b) bridging sites in which water interacts with one side chain and the DKP ring, (c) individual sites in which water interacts only with one polar group. The agreement between the results obtained by the two calculations is sufficiently satisfactory. This allows us to use EMPWI potential for calculations of more complex systems.  相似文献   

9.
We have determined the partial molar volumes, expansibilities, and adiabatic compressibilities of glycine, diglycine, triglycine, tetraglycine, and pentaglycine over the temperature range 18–55°C. These data were analyzed and interpreted in terms of the hydration of these short oligoglycines and their constituent groups. From our results, we have estimated the contributions of the peptide group to the partial molar volume and the partial molar adiabatic compressibility of these oligoglycines. Based on these data, we propose that each of the polar atomic groups of the peptide bond forms approximately two hydrogen bonds with adjacent water molecules. Furthermore, the temperature dependence of the partial molar volume suggests that water that solvates the polar groups of a peptide linkage behaves more like a “normal” liquid than does bulk water, which exhibits its well-known anomalous liquid properties. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
《MABS-AUSTIN》2013,5(1):231-242
Bispecific antibodies have emerged in recent years as a promising field of research for therapies in oncology, inflammable diseases, and infectious diseases. Their capability of dual target recognition allows for novel therapeutic hypothesis to be tested, where traditional mono-specific antibodies would lack the needed mode of target engagement. Among extremely diverse architectures of bispecific antibodies, knobs-into-holes (KIHs) technology, which involves engineering CH3 domains to create either a “knob” or a “hole” in each heavy chain to promote heterodimerization, has been widely applied. Here, we describe the use of a cell-free expression system (Xpress CF) to produce KIH bispecific antibodies in multiple scaffolds, including 2-armed heterodimeric scFv-KIH and one-armed asymmetric BiTE-KIH with tandem scFv. Efficient KIH production can be achieved by manipulating the plasmid ratio between knob and hole, and further improved by addition of prefabricated knob or hole. These studies demonstrate the versatility of Xpress CF in KIH production and provide valuable insights into KIH construct design for better assembly and expression titer.  相似文献   

11.
Self‐assembly of artificially designed proteins is extremely desirable for nanomaterials. Here we show a novel strategy for the creation of self‐assembling proteins, named “Nanolego.” Nanolego consists of “structural elements” of a structurally stable symmetrical homo‐oligomeric protein and “binding elements,” which are multiple heterointeraction proteins with relatively weak affinity. We have established two key technologies for Nanolego, a stabilization method and a method for terminating the self‐assembly process. The stabilization method is mediated by disulfide bonds between Cysteine‐residues incorporated into the binding elements, and the termination method uses “capping Nanolegos,” in which some of the binding elements in the Nanolego are absent for the self‐assembled ends. With these technologies, we successfully constructed timing‐controlled and size‐regulated filament‐shape complexes via Nanolego self‐assembly. The Nanolego concept and these technologies should pave the way for regulated nanoarchitecture using designed proteins.  相似文献   

12.
The concept of “static accessibility” to water has been used to determine the accessible surface area of a cyclic dipeptide: c(l-Thr-l-His). Different calculated and experimental conformations of this model molecule have been examined, which allows us to analyse the variations of accessibility of the hydration sites localized on the peptide backbone and on the polar side chains. The maximum solvation criterion involves a large destabilization of conformations governed by intramolecular interactions. The variations of the amphiphilic character with the conformations are relatively small. Nevertheless, the experimental conformation seems to reflect such a behaviour, especially in the crystal, in which the amphiphilic character is compatible with intermolecular interactions. The accessibility studies must be regarded only as a preliminary step to a more quantitative analysis of peptide hydration.  相似文献   

13.
The heat capacity, enthalpy, entropy, and Gibbs energy changes for the temperature-induced unfolding of 11 globular proteins of known three-dimensional structure have been obtained by microcalorimetric measurements. Their experimental values are compared to those we calculate from the change in solvent-accessible surface area between the native proteins and the extended polypeptide chain. We use proportionality coefficients for the transfer (hydration) of aliphatic, aromatic, and polar groups from gas phase to aqueous solution, we estimate vibrational effects, and we discuss the temperature dependence of each constituent of the thermodynamic functions. At 25 degrees C, stabilization of the native state of a globular protein is largely due to two favorable terms: the entropy of non-polar group hydration and the enthalpy of interactions within the protein. They compensate the unfavorable entropy change associated with these interactions (conformational entropy) and with vibrational effects. Due to the large heat capacity of nonpolar group hydration, its stabilizing contribution decreases quickly at higher temperatures, and the two unfavorable entropy terms take over, leading to temperature-induced unfolding.  相似文献   

14.
Optical second-harmonic scattering in rat-tail tendon   总被引:1,自引:0,他引:1  
Shmuel Roth  Isaac Freund 《Biopolymers》1981,20(6):1271-1290
The angular dependence of optical second-harmonic generation by native, wet, rat-tail tendon is found to display a sharp, intense, forward peak superimposed on a broad background. The sharp peak is shown to imply long-range polar order, while the broad background corresponds to that predicted for the random “up”/“down” array of collagen fibrils seen with the electron microscope. The degree of polar order is determined, and the dependence of the fibril diameter distribution on age and state of hydration is measured. The coherence length of tendon for harmonic generation and the absolute magnitude of the nonlinear susceptibility of the collagen fibril are also determined. The biological significance of these various findings is discussed.  相似文献   

15.
Y K Cheng  P J Rossky 《Biopolymers》1999,50(7):742-750
The use of a linear relationship between free energy of hydrophobic hydration and solvent-accessible apolar surface area has been helpful in interpreting the thermodynamics of biological macromolecules. However, a recent study (Y.-K. Cheng, P. J. Rossky, Nature 1998, Vol. 392, pp. 696-699) has established a substantial enthalpic dependence on biomolecular surface topography, originating from solvent hydrogen-bonding loss in a restrictive geometry. In this study, we use molecular dynamics simulations of 2-Zn insulin in water solvent to explore the further effect of vicinal polar or charged groups on hydrophobic hydration at a biomolecular surface. In contrast to the case for solvent more isolated from such polar solute influences, the binding energies of the water that is proximal to the hydrophobic dimeric interface of insulin and vicinal to polar and charged groups are comparable to the bulk solvent value, a result of compensating interaction primarily with the solute counterions. The results suggest a special importance for such polar/charged groups in biological processes involving hydrophobic surface regions of restricted geometry and also suggest a general route for tuning the hydrophobicity of interfaces.  相似文献   

16.
The heat capacity plays a major role in the determination of the energetics of protein folding and molecular recognition. As such, a better understanding of this thermodynamic parameter and its structural origin will provide new insights for the development of better molecular design strategies. In this paper we have analyzed the absolute heat capacity of proteins in different conformations. The results of these studies indicate that three major terms account for the absolute heat capacity of a protein: (1) one term that depends only on the primary or covalent structure of a protein and contains contributions from vibrational frequencies arising from the stretching and bending modes of each valence bond and internal rotations; (2) a term that contains the contributions of noncovalent interactions arising from secondary and tertiary structure; and (3) a term that contains the contributions of hydration. For a typical globular protein in solution the bulk of the heat capacity at 25°C is given by the covalent structure term (close to 85% of the total). The hydration term contributes about 15 and 40% to the total heat capacity of the native and unfolded states, respectively. The contribution of non-covalent structure to the total heat capacity of the native state is positive but very small and does not amount to more than 3% at 25°C. The change in heat capacity upon unfolding is primarily given by the increase in the hydration term (about 95%) and to a much lesser extent by the loss of noncovalent interactions (up to ~5%). It is demonstrated that a single universal mathematical function can be used to represent the partial molar heat capacity of the native and unfolded states of proteins in solution. This function can be experimentally written in terms of the molecular weight, the polar and apolar solvent accessible surface areas, and the total area buried from the solvent. This unique function accurately predicts the different magnitude and temperature dependences of the heat capacity of both the native and unfolded states, and therefore of the heat capacity changes associated with folding/unfolding transitions. © 1995 Wiley-Liss, Inc.  相似文献   

17.
Finding why protein–protein interactions (PPIs) are so specific can provide a valuable tool in a variety of fields. Statistical surveys of so‐called transient complexes (like those relevant for signal transduction mechanisms) have shown a tendency of polar residues to participate in the interaction region. Following this scheme, residues in the unbound partners have to compete between interacting with water or interacting with other residues of the protein. On the other hand, several works have shown that the notion of active site electrostatic preorganization can be used to interpret the high efficiency in enzyme reactions. This preorganization can be related to the instability of the residues important for catalysis. In some enzymes, in addition, conformational changes upon binding to other proteins lead to an increase in the activity of the enzymatic partner. In this article the linear response approximation version of the semimacroscopic protein dipoles Langevin dipoles (PDLD/S‐LRA) model is used to evaluate the stability of several residues in two phosphate hydrolysis enzymes upon complexation with their activating partners. In particular, the residues relevant for PPI and for phosphate hydrolysis in the CDK2/Cyclin A and Ras/GAP complexes are analyzed. We find that the evaluation of the stability of residues in these systems can be used to identify not only active site regions but it can also be used as a guide to locate “hot spots” for PPIs. We also show that conformational changes play a major role in positioning interfacing residues in a proper “energetic” orientation, ready to interact with the residues in the partner protein surface. Thus, we extend the preorganization theory to PPIs, extrapolating the results we obtained from the above‐mentioned complexes to a more general case. We conclude that the correlation between stability of a residue in the surface and the likelihood that it participates in the interaction can be a general fact for transient PPIs. Proteins 2006. © 2005 Wiley‐Liss, Inc.  相似文献   

18.
Solvent-accessible surfaces of nucleic acids   总被引:14,自引:0,他引:14  
Static solvent-accessible surface areas were calculated for DNA and RNA double helices of varied conformation, composition and sequence, for the single helix of poly(rC), and for a transfer RNA. The results show that for DNA and RNA double helices, two thirds of the water-accessible surface area become buried on double helix formation; phosphate oxygens retain near maximal exposure while the bases are 80% buried. Transfer RNA exposes slightly less surface per residue than does double-helical RNA, despite the presence of several additional “modified” groups, all of which are exposed significantly.When a probe corresponding to a single water molecule is used, both the total and atom type exposures are very similar for A-DNA and B-DNA, although marked differences appear in the major and minor groove exposures between the two conformations. For a given base-pair, the accessible surface area buried upon double-helical stacking is nearly constant (within 5%) for different sequences of neighboring base-pairs.For probes larger than single water molecules, there exist considerable differences in the total and atom type exposures of A-DNA and B-DNA. Conformational transitions between the A-DNA and B-DNA helical forms can thus be related to differences in the accessible areas for “structured” water, or a secondary hydration shell, rather than to interactions with individual water molecules of the primary hydration shell. The base-composition dependence of DNA helical conformation can be explained in terms of the opposing effects of thymine methyl groups of A · T base-pairs and the amino groups of G · C base-pairs upon the solvent within the grooves.The area calculations show that primarily the major groove of B-DNA and the minor groove of A-DNA have sufficient accessible surface area to be recognized by a probe size corresponding to the side-chains of amino acids.  相似文献   

19.
Bush J  Makhatadze GI 《Proteins》2011,79(7):2027-2032
It is well known that nonpolar residues are largely buried in the interior of proteins, whereas polar and ionizable residues tend to be more localized on the protein surface where they are solvent exposed. Such a distribution of residues between surface and interior is well understood from a thermodynamic point: nonpolar side chains are excluded from the contact with the solvent water, whereas polar and ionizable groups have favorable interactions with the water and thus are preferred at the protein surface. However, there is an increasing amount of information suggesting that polar and ionizable residues do occur in the protein core, including at positions that have no known functional importance. This is inconsistent with the observations that dehydration of polar and in particular ionizable groups is very energetically unfavorable. To resolve this, we performed a detailed analysis of the distribution of fractional burial of polar and ionizable residues using a large set of ?2600 nonhomologous protein structures. We show that when ionizable residues are fully buried, the vast majority of them form hydrogen bonds and/or salt bridges with other polar/ionizable groups. This observation resolves an apparent contradiction: the energetic penalty of dehydration of polar/ionizable groups is paid off by favorable energy of hydrogen bonding and/or salt bridge formation in the protein interior. Our conclusion agrees well with the previous findings based on the continuum models for electrostatic interactions in proteins. Proteins 2011; © 2011 Wiley‐Liss, Inc.  相似文献   

20.
The cooperative nature of the protein folding process is independent of the characteristic fold and the specific secondary structure attributes of a globular protein. A general folding/unfolding model should, therefore, be based upon structural features that transcend the peculiarities of α-helices, β-sheets, and other structural motifs found in proteins. The studies presented in this paper suggest that a single structural characteristic common to all globular proteins is essential for cooperative folding. The formation of a partly folded state from the native state results in the exposure to solvent of two distinct regions: (1) the portions of the protein that are unfolded; and (2) the “complementary surfaces,” located in the regions of the protein that remain folded. The cooperative character of the folding/unfolding transition is determined largely by the energetics of exposing complementary surface regions to the solvent. By definition, complementary regions are present only in partly folded states; they are absent from the native and unfolded states. An unfavorable free energy lowers the probability of partly folded states and increases the cooperativity of the transition. In this paper we present a mathematical formulation of this behavior and develop a general cooperative folding/unfolding model, termed the “complementary region” (CORE) model. This model successfully reproduces the main properties of folding/unfolding transitions without limiting the number of partly folded states accessible to the protein, thereby permitting a systematic examination of the structural and solvent conditions under which intermediates become populated. It is shown that the CORE model predicts two-state folding/unfolding behavior, even though the two-state character is not assumed in the model. © 1993 Wiley-Liss, 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