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.
We present heteronuclear two-dimensional methods for the analysis of the geometry of exchangeable protons on a protein-bound carbohydrate. By using a water-selective NOESY-HSQC, we observed cross-relaxation between carbohydrate hydroxyl protons and non-exchangeable ring protons in the complex of [13C6]--methyl-D-mannopyranoside with recombinant rat mannose binding protein. Using a simple kinetic model, we were able to explain the differences in the initial slopes of the resulting cross-relaxation buildup curves in terms of the geometry of the hydroxyl protons in the bound state. The hydroxyl rotamers consistent with our cross-relaxation data fit very well with predictions based on the crystal structure of MBP bound to a mannose-rich oligosaccharide. These methods should be applicable to other systems where both ligand exchange and water exchange are fast relative to the rate of cross-relaxation.  相似文献   

2.
E Mukidjam  G A Elgavish  S Barnes 《Biochemistry》1987,26(21):6785-6792
A paramagnetic NMR study of the structure of the calcium-glycocholate complex in submicellar solution, utilizing dysprosium as an isomorphous lanthanide replacement of calcium, is presented. The dysprosium-induced relaxation rate (1/T1) enhancements of certain glycocholate protons have been used to estimate internuclear distances between these protons and the metal ion. An approximation to calculate the intrinsic relaxation rate (1/T1) enhancements for a nondilute paramagnetic solution is given in the Appendix. From these data, and analysis based on conformation averaging and minimum energy conformations, a molecular model of the dysprosium-glycocholate complex in submicellar aqueous solution has been constructed. In this model the metal ion has a unidentate, first-sphere interaction with the proximal oxygen atom of the glycine carboxyl. The metal ion has second-sphere interactions with the peptide bond carbonyl oxygen (3.6 A) and the distal carboxyl oxygen (4.4 A). The metal ion to hydroxyl oxygen distances (8.4-12.4 A) are not compatible with any metal ion to hydroxyl coordination. The side chain appears to exist in one predominant conformation. All six oxygen atoms of glycocholate, the peptide bond carbonyl, the carboxyl group, and the hydroxyl groups are on the alpha face of the bile salt molecule. On the basis of these features we conclude that in the submicellar state the solution structure of the dysprosium-glycocholate complex displays a metal ion enhanced segregation of polar versus nonpolar groups to the two separate faces of the molecule, which may result in a facilitated hydrophobic interaction of different complex units.  相似文献   

3.
Water T2 relaxation in sugar solutions   总被引:2,自引:0,他引:2  
1H spin-spin relaxation times of water were measured with the CPMG sequence in dilute aqueous solutions of glucitol, mannitol, glycerol, glycol, the methyl D-pyranosides of alpha-glucose, beta-glucose, alpha-galactose, beta-galactose, alpha-xylose, beta-xylose, beta-arabinose and sucrose, alpha,alpha-trehalose, beta-maltose, maltotriose and maltoheptaose. The relaxation-time dispersion was measured by varying the CPMG pulse spacing, tau. These data were interpreted by means of the Carver-Richards model in which exchange between water protons and labile solute hydroxyl protons provides a significant contribution to the relaxation. From the dependences on temperature and tau, parameters characteristic of the pool of hydroxyls belonging to a given solute were extracted by nonlinear regression, including: the fraction of exchangeable protons, P, the chemical-shift difference between water protons and hydroxyl protons, deltaomega, the intrinsic spin-spin relaxation time, T2, and the chemical exchange rate, k. These solute-specific parameters are related, respectively, to the concentration, identity, mobility and exchange life-time of the hydroxyl site. At 298 K, values of deltaomega, T2 and k were found to be of the order of 1 ppm, 100 ms and 1000 s(-1), respectively. Effects of molecular size, conformation and solute concentration were investigated. The exchange mechanism was characterised by Eyring activation enthalpies and entropies with values in the ranges 50-70 kJ mol(-1) and -10 to 60 J K(-1)mol(-1), respectively.  相似文献   

4.
Understanding the properties of interfacial water at solid–liquid interfaces is important in a wide range of applications. Molecular dynamics is becoming a widespread tool for this purpose. Unfortunately, however, the results of such studies are known to strongly depend on the selection of force fields. It is, therefore, of interest to assess the extent by which the implemented force fields can affect the predicted properties of interfacial water. Two silica surfaces, with low and high surface hydroxyl density, respectively, were simulated implementing four force fields. These force fields yield different orientation and flexibility of surface hydrogen atoms, and also different interaction potentials with water molecules. The properties for interfacial water were quantified by calculating contact angles, atomic density profiles, surface density distributions, hydrogen bond density profiles and residence times for water near the solid substrates. We found that at low surface density of hydroxyl groups, the force field strongly affects the predicted contact angle, while at high density of hydroxyl groups, water wets all surfaces considered. From a molecular-level point of view, our results show that the position and intensity of peaks observed from oxygen and hydrogen atomic density profiles are quite different when different force fields are implemented, even when the simulated contact angles are similar. Particularly, the surfaces simulated by the CLAYFF force field appear to attract water more strongly than those simulated by the Bródka and Zerda force field. It was found that the surface density distributions for water strongly depend on the orientation of surface hydrogen atoms. In all cases, we found an elevated number of hydrogen bonds formed between interfacial water molecules. The hydrogen bond density profile does not depend strongly on the force field implemented to simulate the substrate, suggesting that interfacial water assumes the necessary orientation to maximise the number of water–water hydrogen bonds irrespectively of surface properties. Conversely, the residence time for water molecules near the interface strongly depends on the force field and on the flexibility of surface hydroxyl groups. Specifically, water molecules reside for longer times at contact with rigid substrates with high density of hydroxyl groups. These results should be considered when comparisons between simulated and experimental data are attempted.  相似文献   

5.
Khare D  Alexander P  Orban J 《Biochemistry》1999,38(13):3918-3925
Protium-deuterium fractionation factors (phi) were determined for more than 85% of the backbone amide protons in the IgG binding domains of protein G, GB1 and GB2, from NMR spectra recorded over a range of H2O/D2O solvent ratios. Previous studies suggest a correlation between phi and hydrogen bond strength; amide and hydroxyl groups in strong hydrogen bonds accumulate protium (phi < 1), while weak hydrogen bonds accumulate deuterium (phi > 1). Our results show that the alpha-helical residues have slightly lower phi values (1.03 +/- 0.05) than beta-sheet residues (1.12 +/- 0.07), on average. The lowest phi value obtained (0.65) does not involve a backbone amide but rather is for the interaction between two side chains, Y45 and D47. Fractionation factors for solvent-exposed residues are between the alpha-helix and beta-sheet values, on average, and are close to those for random coil peptides. Further, the difference in phiav between alpha-helix and solvent-exposed residues is small, suggesting that differences in hydrogen bond strength for intrachain hydrogen bonds and amide...water hydrogen bonds are also small. Overall, the enrichment for deuterium suggests that most backbone...backbone hydrogen bonds are weak.  相似文献   

6.
J Andrasko 《Biophysical journal》1975,15(12):1235-1243
The dependence of the spin-lattice relaxation time in the rotating frame (T1rho) on radio frequency (RF) field strength and temperature has been studied for agarose gels in order to investigate molecular motion. The results indicate the presence of slow motions with a correlation time of ca. 5-10(-6) s at room temperature. This interaction is responsible for the short spin-spin relaxation times (T2) for water protons in agarose gels and is ascribed to firmly bound water. The fraction of bound water is estimated to about 0.003 for a 7.3% agarose gel. The motion of the more mobile protons in agarose-water systems can not be characterized by single correlation time. This fraction is presumably composed of water in different motional states and some of the agarose hydroxyl protons. Higher mobilities are the most common.  相似文献   

7.
The hydrogen bonding arrangement in anhydrous β-chitin, a homopolymer of N-acetylglucosamine, was directly determined by neutron fiber diffraction. Data were collected from a sample prepared from the bathophilous tubeworm Lamellibrachia satsuma in which all labile hydrogen atoms had been replaced by deuterium. Initial positions of deuterium atoms on hydroxyl and acetamide groups were directly located in Fourier maps synthesized using phases calculated from the X-ray structure and amplitudes measured from the neutron data. The hydrogen bond arrangement in the refined structure is in general agreement with predictions based on the X-ray structure: O3 donates a hydrogen bond to the O5 ring oxygen atom of a neighboring residue in the same chain; N2 and O6 donate hydrogen bonds to the same carbonyl oxygen O7 of an adjacent chain. The intramolecular O3···O5 hydrogen bond has the most energetically favorable geometry with a hydrogen to acceptor distance of 1.77 ? and a hydrogen bond angle of 171°.  相似文献   

8.
In large molecular structures, the magnetization of all hydrogen atoms in the solute is strongly coupled to the water magnetization through chemical exchange between solvent water and labile protons of macromolecular components, and through dipole–dipole interactions and the associated “spin diffusion” due to slow molecular tumbling. In NMR experiments with such systems, the extent of the water polarization is thus of utmost importance. This paper presents a formalism that describes the propagation of the water polarization during the course of different NMR experiments, and then compares the results of model calculations for optimized water polarization with experimental data. It thus demonstrates that NMR spectra of large molecular structures can be improved with the use of paramagnetic spin relaxation agents which selectively enhance the relaxation of water protons, so that a substantial gain in signal-to-noise can be achieved. The presently proposed use of a relaxation agent can also replace the water flip-back pulses when working with structures larger than about 30 kDa. This may be a valid alternative in situations where flip-back pulses are difficult to introduce into the overall experimental scheme, or where they would interfere with other requirements of the NMR experiment.Supplementary material to this paper is available in electronic form at http://dx.doi.org/10.1007/s10858-005-3070-8  相似文献   

9.
The number of protons available for hydrogen-deuterium exchange was predicted for ten globular proteins using a method described elsewhere by the authors. The average number of protons replaced by deuterium was also determined by mass spectrometry of the intact proteins in their native conformations. Based on these data, we find that two models proposed earlier agree with each other in estimation of the number of protons replaced by deuterium. Using a model with a probability scale for hydrogen bond formation, we estimated a number of protons replaced by deuterium that is close to the experimental data for long-term incubation in D2O (24 h). Using a model based on estimations with a scale of the expected number of contacts in globular proteins there is better agreement with the experimental data obtained for a short period of incubation in D2O (15 min). Therefore, the former model determines weakly fluctuating parts of a protein that are in contact with solvent only for a small fraction of the time. The latter model (based on the scale of expected number of contacts) predicts either flexible parts of a protein chain exposed to interactions with solvent or disordered parts of the protein.  相似文献   

10.
M Eisenstadt 《Biochemistry》1985,24(14):3407-3421
We have measured T1 and T2 of protein and water protons in hemoglobin solutions using broad-line pulse techniques; selective excitation and detection methods enabled the intrinsic protein and water relaxation rates, as well as the spin-transfer rate between them, to be obtained at 5, 10, and 20 MHz. Water and protein T1 data were also obtained at 100 and 200 MHz for hemoglobin in H2O/D2O mixtures by using commercial Fourier-transform instruments. The T1 data conform to a simple model of two well-mixed spin systems with single intrinsic relaxation times and an average spin-transfer rate, with each phase recovering from a radio-frequency excitation with a biexponential time dependence. At low frequencies, protein T1 and T2 agree reasonably with a model of dipolar relaxation of an array of fixed protons tumbling in solution, explicitly calculating methyl and methylene relaxation and using a continuum approximation for the others. Differing values in H2O and D2O are mainly ascribed to solvent viscosity. For water-proton relaxation, T1, T2, and spin transfer were measured for H2O and HDO, which enabled a separation of inter-and intramolecular contributions to relaxation. Despite such detail, few firm conclusions could be reached about hydration water. But it seems clear that few long-lived hydration sites are needed to explain T1 and T2, and the spin-transfer value mandates fewer than five sites with a lifetime longer than 10(-8) s.  相似文献   

11.
H Kessler  J W Bats  K Wagner  M Will 《Biopolymers》1989,28(1):385-395
The strategy and tactics of conformational analysis of cyclic peptides in solution is demonstrated by the example of cyclo(-D-Pro-Phe-Thr-Phe-Trp-Phe-). Spin-locked experiments like rotating frame nuclear Overhauser enhancement spectroscopy (ROESY), ROTO, and TOCSY are successfully applied to assign all proton signals and to obtain distance information. A crude conformational model was built using the nmr data. This starting model was refined by restrained molecular dynamics (MD) calculations using ROE derived distances and fixed bond angles as determined from homo- and heteronuclear coupling constants. To mimic the solvent and to reduce artifacts in an in vacuo calculation the charges of the solvent-exposed NH protons were gradually reduced according to the temperature gradients. The thus obtained "conformation" (mean of a 40 ps MD trajectory) shows very close similarity to x-ray structures in an orthorhombic and in two monoclinic crystal modifications of the same compound. The main difference is the breaking of an intermolecular hydrogen bond of the threonine hydroxyl group on dissolution of the crystal and forming an intramolecular hydrogen bond in solution.  相似文献   

12.
Nuclear magnetic resonance (nmr) relaxation times are measured for water protons in cross-linked lysozyme crystals below the freezing event as a function of the mole fraction of protons in the water phase. Proton longitudinal nmr relaxation in these samples is nonexponential and the slow longitudinal relaxation component becomes slower linearly with decreasing proton mole fraction in the water. The data are analyzed using a cross relaxation model that eliminates the necessity of postulating long residence times for water molecules in the domain of the protein. The observed isotope dilution behavior is consistent with the cross relaxation model. The deuterium nmr relaxation is also reported for deuterium oxide in the cross-linked protein crystal sample below the freezing event and the relaxation is shown to be accurately exponential.  相似文献   

13.
Park S  Saven JG 《Proteins》2005,60(3):450-463
Buried solvent molecules are common in the core of globular proteins and contribute to structural stability. Folding necessitates the burial of polar backbone atoms in the protein core, whose hydrogen-bonding capacities should be satisfied on average. Whereas the residues in alpha-helices and beta-sheets form systematic main-chain hydrogen bonds, the residues in turns, coils and loops often contain polar atoms that fail to form intramolecular hydrogen bonds. The statistical analysis of 842 high resolution protein structures shows that well-resolved, internal water molecules preferentially reside near residues without alpha-helical and beta-sheet secondary structures. These buried waters most often form primary hydrogen bonds to main-chain atoms not involved in intramolecular hydrogen bonds, providing strong evidence that hydrating main-chain atoms is a key structural role of buried water molecules. Additionally, the average B-factor of protein atoms hydrogen-bonded to waters is smaller than that of protein atoms forming intramolecular hydrogen bonds, and the average B-factor of water molecules involved in primary hydrogen bonds with main-chain atoms is smaller than the average B-factor of water molecules involved in secondary hydrogen bonds to protein atoms that form concurrent intramolecular hydrogen bonds. To study the structural coupling between internal waters and buried polar atoms in detail we simulated the dynamics of wild-type FKBP12, in which a buried water, Wat137, forms one side-chain and multiple main-chain hydrogen bonds. We mutated E60, whose side-chain hydrogen bonds with Wat137, to Q, N, S or A, to modulate the multiplicity and geometry of hydrogen bonds to the water. Mutating E60 to a residue that is unable to form a hydrogen bond with Wat137 results in reorientation of the water molecule and leads to a structural readjustment of residues that are both near and distant to the water. We predict that the E60A mutation will result in a significantly reduced affinity of FKBP12 for its ligand FK506. The propensity of internal waters to hydrogen bond to buried polar atoms suggests that ordered water molecules may constitute fundamental structural components of proteins, particularly in regions where alpha-helical or beta-sheet secondary structure is not present.  相似文献   

14.
The D-gluconate ion is found to have the planar, extended carbon-chain conformation in the crystal structure of potassium D-gluconate monohydrate, with an intramolecular hydrogen-bond between 0-2 and 0-4. The D-gluconate ions and water molecules are linked in puckered sheets by a series of intermolecular hydrogen-bonds that involve the water molecules, the carboxylate groups, and pairs of hydroxyl groups. One hydroxyl group in the ion does not form a hydrogen bond. The potassium ions lie between the puckered sheets, with an eight-fold coordination of six D-gluconate groups and two water oxygen atoms. The crystal structure was determined from three-dimensional, CuKα, X-ray diffraction data taken on an automatic diffractometer.  相似文献   

15.
Exchangeable protons in the immediate neighborhood of the semiquinone (SQ) at the Qi-site of the bc1 complex (ubihydroquinone:cytochrome c oxidoreductase (EC 1.10.2.2)) from Rhodobacter sphaeroides have been characterized using electron spin echo envelope modulation (ESEEM) and hyperfine sublevel correlation spectroscopy (HYSCORE) and visualized by substitution of H2O by 2H2O. Three exchangeable protons interact with the electron spin of the SQ. They possess different isotropic and anisotropic hyperfine couplings that allow a clear distinction between them. The strength of interactions indicates that the protons are involved in hydrogen bonds with SQ. The hyperfine couplings differ from values typical for in-plane hydrogen bonds previously observed in model experiments. It is suggested that the two stronger couplings involve formation of hydrogen bonds with carbonyl oxygens, which have a significant out-of-plane character due to the combined influence of bulky substituents and the protein environment. These two hydrogen bonds are most probably to side chains suggested from crystallographic structures (His-217 and Asp-252 in R. sphaeroides). Assignment of the third hydrogen bond is more ambiguous but may involve either a bond between Asn-221 and a methoxy O-atom or a bond to water. The structural and catalytic roles of the exchangeable protons are discussed in the context of three high resolution crystallographic structures for mitochondrial bc1 complexes. Potential H-bonds, including those to water molecules, form a network connecting the quinone (ubiquinone) occupant and its ligands to the propionates of heme bH and the external aqueous phase. They provide pathways for exchange of protons within the site and with the exteriors, needed to accommodate the different hydrogen bonding requirements of different quinone species during catalysis.  相似文献   

16.
The results of molecular dynamics simulations of three lignin-water systems are presented. Static and dynamic properties of each system are compared to a benchmark system consisting entirely of water molecules. The significantly reduced mobility of water molecules local to lignin hydroxyl regions is attributed to hydrogen bond formation, while the slightly reduced mobility of water molecules in the vicinity of lignin methoxyl groups results from a hydrophobic effect that causes water molecules to structure themselves around these groups. The average diffusion of water in each system correlates with the number of methoxyl groups present in the system. As the number of methoxyls in the system increases, so too does the average diffusion constant of water in that system. The bulky methoxyl groups obstruct water from accessing lignin hydroxyl regions where hydrogen bond formation is anticipated and the hydrogen-bonded water lowers the average diffusion constant.  相似文献   

17.
The crystal structure of galactinol dihydrate has been determined by X-ray diffraction. The crystal belongs to the orthorhombic system, space group P2(1)2(1)2, a = 15.898(6), b = 19.357(5), c = 5.104(4) A, and Z = 4. The structure was refined to R = 0.044 for 1818 observed structure amplitudes. The primary hydroxyl group exhibits twofold orientational disorder. The linkage conformation is close to those of alpha-(1 --> 4) linkages in methyl alpha-maltotrioside tetrahydrate and erlose trihydrate. Although there is no interring hydrogen bond in galactinol, an indirect interring hydrogen bond including a water molecule is present. The observed conformation is additionally stabilized by the indirect interring hydrogen bond. The global minimum in the relaxed-residue energy map based on the MM3(92) force-field is close to the observed conformation in the crystal structure. All hydroxyl, ring and water oxygen atoms are involved in a complex three-dimensional hydrogen-bonding network.  相似文献   

18.
The depsipeptide DNA-intercalating antibiotic luzopeptin was studied in solution by n.m.r. methods. Two-dimensional 1H double-quantum-filtered correlation spectroscopy (DQF-COSY) and nuclear-Overhauser-effect spectroscopy (NOESY) confirm the primary structure and twofold symmetry of luzopeptin and provide details of its three-dimensional conformation in solution. Trans-annular hydrogen bonds between the glycine NH groups and carbonyl oxygen atoms have been identified in the crystalline state [Arnold & Clardy (1981) J. Am. Chem. Soc. 103, 1243-1244], and are important in maintaining an antiparallel beta-sheet conformation. The n.m.r. data indicate that the glycine NH protons are appreciably shielded from the solvent molecules, which suggests that these hydrogen bonds are maintained in solution. The orientation of the quinoline chromophores is defined by two-dimensional NOE cross-peaks that position the N-methyl group of the L-beta-hydroxyvaline residue close in space to both the quinoline H-8 and serine NH proton. This pattern of NOEs is in accord both with the chromophore configuration found in the crystal and one where the quinoline rings are aligned in a parallel manner at right-angles to the depsipeptide ring. The n.m.r. data are consistent with a hydrogen bond between the quinoline hydroxy groups and the quinoline carbonyl oxygen atoms. The pyridazine acetylmethyl groups give NOEs to the C(alpha)H groups of the beta-hydroxy-N-methylvaline residues, showing that the acetyl groups, for at least some of the time, stretch over the depsipeptide ring, occluding one face of the molecule. Both of the latter features are also found in the crystal structure. Resonances in the 13C-n.m.r. spectrum of luzopeptin have been assigned by transferring 1H assignments to their covalently bonded carbon atoms via a heteronuclear shift-correlation experiment (HETCOR). The measurement of spin-lattice relaxation times and 1H-13C NOEs at specific sites in the molecule has led us to conclude that segmental motions within the depsipeptide ring are restricted and that the 13C relaxation data for luzopeptin's protonated carbon atoms are adequately described by isotropic tumbling in solution. Furthermore, relaxation data for the carbon atoms of the quinoline chromophores show that these rings exhibit similar motion to the depsipeptide ring and are not rotating rapidly with respect to it. Taken together all the data imply that luzopeptin is fairly rigid in solution, on the time scale of molecular tumbling, and has, or can readily attain, a staple-like structure suitable for bisintercalation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Refined crystal structure of carboxypeptidase A at 1.54 A resolution   总被引:19,自引:0,他引:19  
The crystal structure of bovine carboxypeptidase A (Cox) has been refined at 1.54 A resolution using the restrained least-squares algorithm of Hendrickson & Konnert (1981). The crystallographic R factor (formula; see text) for structure factors calculated from the final model is 0.190. Bond lengths and bond angles in the carboxypeptidase A model have root-mean-square deviations from ideal values of 0.025 A and 3.6 degrees, respectively. Four examples of a reverse turn like structure (the "Asx" turn) requiring an aspartic acid or asparagine residue are observed in this structure. The Asx turn has the same number of atoms as a reverse turn, but only one peptide bond, and the hydrogen bond that closes the turn is between the Asx side-chain CO group and a main-chain NH group. The distributions of CO-N and NH-O hydrogen bond angles in the alpha-helices and beta-sheet structures of carboxypeptidase A are centered about 156 degrees. A total of 192 water molecules per molecule of enzyme are included in the final model. Unlike the hydrogen bonding geometry observed in the secondary structure of the enzyme, the CO-O(wat) hydrogen bond angle is distributed about 131 degrees, indicating the role of the lone pair electrons of the carbonyl oxygen in the hydrogen bond interaction. Twenty four solvent molecules are observed buried within the protein. Several of these waters are organized into hydrogen-bonded chains containing up to five waters. The average temperature factor for atoms in carboxypeptidase A is 8 A2, and varies from 5 A2 in the center of the protein, to over 30 A2 at the surface.  相似文献   

20.
The nuclear magnetic spin-lattice relaxation rates of water protons are reported for solutions of manganese(II), copper(II), and chromium(III) cage complexes of the sarcophagine type. As simple aqueous solutions, the complexes are only modest magnetic relaxation agents, presumably because they lack protons on atoms in the first-coordination-sphere protons that are sufficiently labile to mix the large relaxation rate at the metal complex with that of the bulk solvent. The relaxation is approximately modeled using spectral density functions derived for translational diffusion of the interacting dipole moments with the modification that the electron spin relaxation rate is directly included as a contribution to the correlation time. In all cases studied, the electron spin relaxation rate is sufficiently large that it contributes directly to the water-proton spin relaxation process. The poor relaxation efficiency of the cage compound may, however, be improved dramatically by binding the complex to a protein. The efficiency is improved even further if the rotational motion of the protein is reduced drastically by an intermolecular cross-linking reaction. The relaxation efficiency of the cross-linked protein-cage complexes rivals that of the best first-coordination-sphere relaxation agents like [Gd(DTPA)(H2O)]2- and [Gd(DOTA)(H2O)]-.  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