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 46 毫秒
1.
A reduced point charge distribution is used to model Ubiquitin and two complexes, Vps27 UIM-1–Ubiquitin and Barnase–Barstar. It is designed from local extrema in charge density distributions obtained from the Poisson equation applied to smoothed molecular electrostatic potentials. A variant distribution is built by locating point charges on atoms. Various charge fitting conditions are selected, i.e. from either electrostatic Amber99 (Assisted Model Building with Energy Refinement) Coulomb potential or forces, considering reference grid points located within various distances from the protein atoms, with or without separate treatment of main and side chain charges. The program GROMACS (Groningen Machine for Chemical Simulations) is used to generate Amber99SB molecular dynamics (MD) trajectories of the solvated proteins modelled using the various reduced point charge models (RPCMs) so obtained. Point charges that are not located on atoms are considered as virtual sites. Some RPCMs lead to stable MD trajectories. They, however, involve a partial loss in the protein secondary structure and lead to a less-structured solute solvation shell. The model built by fitting charges on Coulomb forces calculated at grid points ranging between 1.4 and 2.0 times the van der Waals radius of the atoms, with a separate treatment of main chain and side chain charges, appears to best approximate all-atom MD trajectories.  相似文献   

2.
The charge density (CD) distribution of an atom is the difference per unit volume between the positive charge of its nucleus and the distribution of the negative charges carried by the electrons that are associated with it. The CDs of the atoms in macromolecules are responsible for their electrostatic potential (ESP) distributions, which can now be visualized using cryo‐electron microscopy at high resolution. CD maps can be recovered from experimental ESP density maps using the negative Laplacian operation. CD maps are easier to interpret than ESP maps because they are less sensitive to long‐range electrostatic effects. An ESP‐to‐CD conversion involves multiplication of amplitudes of structure factors as Fourier transforms of these maps in reciprocal space by 1/d2, where d is the resolution of reflections. In principle, it should be possible to determine the charges carried by the individual atoms in macromolecules by comparing experimental CD maps with experimental ESP maps.  相似文献   

3.
The preparation and crystal and molecular structure of the osmium tetraoxide bispyridine ester of 1-methylthymine are reported. The complex crystallizes in the triclinic system, space group P1, with a = 11.493(6)A, b = 16.655(7)A, c = 6.082(2)A, alpha = 92.07(3) degrees, beta = 90.58(3) degrees, gamma = 71.36(4) degrees, V = 1102.4 A3, Dm = 1.85(1) g cm-3, DC = 1.84 g cm-3. The unit cell contains 2 osmium tetraoxide bispyridine esters of 1-methylthymine, 2 waters of crystallization and 1 disordered pyridine of solvation. Intensities for 3814 independent reflections were collected by counter methods. The structure was solved by standard heavy-atom techniques and has been refined by full-matrix least squares, based on F, to a final R value of 0.065. The osmium complex binds as a cis osmate ester to the C(5)-C(6) bond of the methylated pyrimidine in a fashion which is expected to be similar to the binding of the complex to thymidine residues in nucleic acids. The conformation of the 1-methylthymine ester is that of a half chair with C(6) showing a substantial deviation, 0.55 A, from the best mean plane of the thymine moiety. The primary coordination sphere about the Os(VI) atom is completed by 2 axial Os=O bonds and the binding of the 2 pyridine ligands in cis positions in the equatorial plane containing the ester linkages. The O=Os=O group is substantially nonlinear, 164.0(5) degrees, and this nonlinearity is attributed to intracomplex electronic effects.  相似文献   

4.
In our attempts to design crystalline alpha-helical peptides, we synthesized and crystallized GAI (C11H21N3O4) in two crystal forms, GAI1 and GAI2. Form 1 (GAI1) Gly-L-Ala-L-Ile (C11H21N3O4.3H2O) crystals are monoclinic, space group P2(1) with a = 8.171(2), b = 6.072(4), c = 16.443(4) A, beta = 101.24(2) degrees, V = 800 A3, Dc = 1.300 g cm-3 and Z = 2, R = 0.081 for 482 reflections. Form 2 (GAI2) Gly-L-Ala-L-Ile (C11H21N3O4.1/2H2O) is triclinic, space group P1 with a = 5.830(1), b = 8.832(2), c = 15.008(2) A, alpha = 102.88(1), beta = 101.16(2), gamma = 70.72(2) degrees, V = 705 A3, Z = 2, Dc = 1.264 g cm-3, R = 0.04 for 2582 reflections. GAI1 is isomorphous with GAV and forms a helix, whereas GAI2 does not. In GAI1, the tripeptide molecule is held in a near helical conformation by a water molecule that bridges the NH3+ and COO- groups, and acts as the fourth residue needed to complete the turn by forming two hydrogen bonds. Two other water molecules form intermolecular hydrogen bonds in stabilizing the helical structure so that the end result is a column of molecules that looks like an incipient alpha-helix. GAI2 imitates a cyclic peptide and traps a water molecule. The conformation angles chi 11 and chi 12 for the side chain are (-63.7 degrees, 171.1 degrees) for the helical GAI1, and (-65.1 degrees, 58.6 degrees) and (-65.0 degrees, 58.9 degrees) for the two independent nonhelical molecules in GAI2; in GAI1, both the C gamma atoms point away from the helix, whereas in GAI2 the C gamma atom with the g+ conformation points inward to the helix and causes sterical interaction with atoms in the adjacent peptide plane. From these results, it is clear that the helix-forming tendencies of amino acids correlate with the restrictions of side-chain rotamer conformations. Both the peptide units in GAI1 are trans and show significant deviation from planarity [omega 1 = -168(1) degrees; omega 2 = -171(1) degrees] whereas both the peptide units in both the molecules A and B in GAI2 do not show significant deviation from planarity [omega 1 = 179.3(3) degrees; omega 2 = -179.3(3) degrees for molecule A and omega 1 = 179.5(3) degrees; omega 2 = -179.4(3) degrees for molecule B], indicating that the peptide planes in these incipient alpha-helical peptides are considerably bent.  相似文献   

5.
We propose an approach for approximating electrostatic charge distributions with a small number of point charges to optimally represent the original charge distribution. By construction, the proposed optimal point charge approximation (OPCA) retains many of the useful properties of point multipole expansion, including the same far-field asymptotic behavior of the approximate potential. A general framework for numerically computing OPCA, for any given number of approximating charges, is described. We then derive a 2-charge practical point charge approximation, PPCA, which approximates the 2-charge OPCA via closed form analytical expressions, and test the PPCA on a set of charge distributions relevant to biomolecular modeling. We measure the accuracy of the new approximations as the RMS error in the electrostatic potential relative to that produced by the original charge distribution, at a distance the extent of the charge distribution–the mid-field. The error for the 2-charge PPCA is found to be on average 23% smaller than that of optimally placed point dipole approximation, and comparable to that of the point quadrupole approximation. The standard deviation in RMS error for the 2-charge PPCA is 53% lower than that of the optimal point dipole approximation, and comparable to that of the point quadrupole approximation. We also calculate the 3-charge OPCA for representing the gas phase quantum mechanical charge distribution of a water molecule. The electrostatic potential calculated by the 3-charge OPCA for water, in the mid-field (2.8 Å from the oxygen atom), is on average 33.3% more accurate than the potential due to the point multipole expansion up to the octupole order. Compared to a 3 point charge approximation in which the charges are placed on the atom centers, the 3-charge OPCA is seven times more accurate, by RMS error. The maximum error at the oxygen-Na distance (2.23 Å ) is half that of the point multipole expansion up to the octupole order.  相似文献   

6.
We describe an improved force field parameter set for the generalized AMBER force field (GAFF) for urea. Quantum chemical computations were used to obtain geometrical and energetic parameters of urea dimers and larger oligomers using AM1 semiempirical MO theory, density functional theory at the B3LYP/6-31G(d,p) level, MP2 and CCSD ab initio calculations with the 6-311++G(d,p), aug-cc-pVDZ, aug-cc-pVTZ, and aug-cc-pVQZ basis sets, and with the CBS-QB3 and CBS-APNO complete basis set methods. Seven different urea dimer structures were optimized at the MP2/aug-cc-pVDZ level to obtain accurate interaction energies. Atomic partial charges were calculated at the MP2/aug-cc-pVDZ level with the restrained electrostatic potential (RESP) fitting approach. The interaction energies computed with these new RESP charges in the force field are consistent with those obtained from CCSD and MP2 calculations. The linear dimer structure calculated using the force field with modified geometrical parameters and the new RESP charge set agrees well with available experimental data.  相似文献   

7.
Bis-Methyl N,N-diethylcarbamylmethylenephosphonato dysprosium thiocyanate, Dy[O2P(OCH3)CH2C(O)N(C2H5)2]2(NCS) was prepared from the combination of ethanolic solutions of Dy(NCS)3·xH2O and (CH3O)2P(O)CH2C(O)N(C2H5)2. The complex was characterized by infrared and NMR spectroscopy, and single crystal X-ray diffraction methods. The crystal structure was determined at 25 °C from 3727 independent reflections by using a standard automated diffractometer. The complex was found to crystallize in the monoclinic space group P21/c with a = 13.282(4) Å, b = 19.168(5) Å, c = 9.648(2) Å, β = 90.09(2)°, Z = 4, V = 2456.4 Å3 and ?cald = 1.72 g cm?3. The structure was solved by standard heavy atom techniques, and blocked least-squares refinement converged with Rf = 4.7% and RwF = 4.9%. The Dy atom is seven coordinate and bonded in a bidentate fashion to two anionic phosphonate ligands [O2P(OCH3)CH2C(O)N(C2H5)2?] through the carbonyl oxygen atoms and one of two phosphonate oxygen atoms. In addition, each Dy atom is coordinated to two phosphonate oxygen atoms from two neighboring complexes and to the nitrogen atom of a thiocyanate ion. This coordination scheme gives rise to a two-dimensional polymeric structure. Some important bond distances include DyNCS 2.433(8) Å, DyO(carbonyl)avg 2.39(2) Å, DyO(equat. phosphoryl)avg 2.303(8) Å, DyO(axial phosphoryl)avg 2.25(2), PO(phosphoryl)avg 1.493(3) Å and CO(carbonyl)avg 1.25(1) Å.  相似文献   

8.
The association and conformational structure of the molecule of erythromycin in solutions of CCl4, C2Cl4 and CHCl3 were studied by the IR spectra in the region of v OH and vC = O. The analysis of the concentration and temperature changes showed that the erythromycin association was accounted for by the hydrogen linkage of OH ... O = C to the ester group. In the monomer molecule of erythromycin, all hydroxyl groups participated in the intramolecular hydrogen linkage. Band 3513 cm-1 belonged to the OH group in the five-membered cycles of OH ... O. Components 3500, 3530 and 3560 cm-1 of the wide band vOH were assigned to the cycles with OH ... N and OH ... O linkages of a larger size. The association was due to a break in a part of the intramolecular hydrogen linkages. Addition of strong acceptors of proton-hexamethanol and trioctylphosphinoxide to the solution resulted in attenuation of these bands and appearance of a strong band vOH of the erythromycin-acceptor complexes. In the presence of monochloroacetic acid in the solution of CHCl3 stoichiometric protonization of erythromycin was observed. The total acid was in the form of anion (vaCO-2 1610 cm-1) up to a ratio of 1:1. The protonization proceeded according to the nitrogen atom since the antibiotic spectrum in the region of vC=O did not change. Propionic acid titrated erythromycin in methanol solution and in mixtures of water and methanol up to a ratio of 1:5 (v/v). However, in the solution of CHCl3 equilibrium between the neutral and ionized molecules of the acid was seen.  相似文献   

9.
Thomas A  Milon A  Brasseur R 《Proteins》2004,56(1):102-109
Using a semiempirical quantum mechanical procedure (FCPAC) we have calculated the partial atomic charges of amino acids from 494 high-resolution protein structures. To analyze the influence of the protein's environment, we considered each residue under two conditions: either as the center of a tripeptide with PDB structure geometry (free) or as the center of 13-16 amino acid clusters extracted from the PDB structure (buried). The partial atomic charges from residues in helices and in sheets were separated. The FCPAC partial atomic charges of the Cbeta and Calpha of most residues correlate with their helix propensity, positively for Cbeta and negatively for Calpha (r2 = 0.76 and 0.6, respectively). The main consequence of burying residues in proteins is the polarization of the backbone C=O bond, which is more pronounced in helices than in sheets. The average shift of the oxygen partial charges that results from burying is -0.120 in helix and -0.084 in sheet with the charge of the proton as unit. Linear correlations are found between the average NMR chemical shifts and the average FCPAC partial charges of Calpha (r2 = 0.8-0.85), N (r3 = 0.67-0.72), and Cbeta (r2 = 0.62) atoms. Correlations for helix and beta-sheet FCPAC partial charges show parallel regressions, suggesting that the charge variations due to burying in proteins differentiate between the dihedral angle effects and the polarization of backbone atoms.  相似文献   

10.
The crystal structure of cholestanyl n-octanoate (caprylate) (C35H62O2) is monoclinic with space group A2 and cell dimensions a = 10.103(7), b = 7.646(7), c = 87.63(7) A, beta = 90.51(6) degrees; Z = 8 [two molecules (A, B) in asymmetric unit], V = 6769 A3, Dc = 1.010 g cm-3. Integrated X-ray intensities for 3798 reflections with I greater than 2 sigma (I) were measured with a rotating anode diffractometer at room temperature. The structure was determined using direct methods. Block diagonal least squares refinement gave R = 0.111. Molecules A and B have almost fully extended conformations, but differ significantly in the rotation about the ester bond and in the C17 chains. The molecular packing in the crystal structure of cholestanyl caprylate consists of stacked bilayers each having d002 = 43.8 A in thickness and within each bilayer, cholestanols pack with cholestanols and caprylate chains pack with caprylate chains. The crystal structure is very similar to that of cholesteryl myristate but is quite different from that of cholesteryl caprylate. The phase equilibria of the cholestanyl caprylate/cholesteryl caprylate binary system have been shown to involve limited mutual solubility of the two components and to have a eutectic point at 73% cholestanyl caprylate. The cholesteric mesophase is monotropic at all compositions except for a narrow range near the eutectic point where it is enantiotropic.  相似文献   

11.
Molecular electrostatic potential(MEP) maps of azido thymidine (AZT), some of its analogs and derivatives and certain other 2′,3′-dideoxy nucleosides having different anti-HIV activities have been studied. The optimised hybridization displacement charges (HDC) combined with MNDO Löwdin charges, continuosly distributed in three dimension spherically symmetrically as a Slater cloud at each site were used to compute the MEP maps. The negative MEP region near the O5′ sites of these molecules appears to be of primary importance from the point of view of their anti-HIV activity. The roles of the azido group in AZT and fluorine atoms substituted at different positions in the sugar moiety have been evaluated. The azido group in AZT behaves as a strongly electronegative group.  相似文献   

12.
Boc-L-Asn-L-Pro-OBzl: C21H29O6N3.CH3OH, Mr = 419.48 + CH3 OH, monoclinic, P2(1), a = 10.049(1), b = 10.399(2), c = 11.702(1) A, beta = 92.50(1)degrees, V = 1221.7(3) A3, dx = 1.14 g.cm-3, Z = 2, CuK alpha (lambda = 1.54178 A), F(000) = 484 (with solvent), 23 degrees, unique reflections (I greater than 3 sigma(I)) = 1745, R = 0.043, Rw = 0.062, S = 1.66. Boc-beta-cyano-L-alanine-L-Pro-OBzl: C21H27O5N3, Mr = 401.46, orthorhombic, P2(1)2(1)2(1), a = 15.741(3), b = 21.060(3), c = 6.496(3) A, V = 2153(1) A3, dx = 1.24 g.cm-3, Z = 4, CuK alpha (lambda = 1.54178 A), F(000) = 856, 23 degrees, unique reflections (I greater than 3 sigma(I)) = 1573, R = 0.055, Rw = 0.078, S = 1.86. The tert.-butyloxycarbonyl (Boc) protected dipeptide benzyl ester (OBzl), Boc-L-Asn-L-Pro-OBzl, prepared from a mixed anhydride reaction using isobutylchloroformate, Boc-L-asparagine, and HCl.L-proline-OBzl, crystallized with one methanol per asymmetric unit in an extended conformation with the Asn-Pro peptide bond trans. Intermolecular hydrogen bonding occurs between the methanol and the Asn side chain and between the peptide backbone and the Asn side chain. A minor impurity due to the dehydration of the Asn side chain to a beta-CNala crystallized with a similar extended conformation and a single intermolecular hydrogen bond.  相似文献   

13.
The Tanford–Kirkwood theory for evaluating the electrostatic free energy of a discrete charge distribution in the presence of ion atmosphere is extended to concentric dielectric continua. The theory is applied to study the conformational preferences with respect to phosphodiester torsion angles in the dimethylphosphate anion (DMP?) and the sodium dimethylphosphate ion pair (Na+DMP?), in the absence and presence of ion atmosphere and at varying local dielectric constants. Results indicate that phosphodiester torsion angles in DMP? prefer the gauche-gauche conformation in aqueous solutions.  相似文献   

14.
Two quasi-multipole electrostatic models for molecular charge distributions are presented. They assign arrays of point charges to nonhydrogen atoms on the basis of hybrid orbitals or localised molecular orbitals. When used with common semiempirical MO-techniques, they reproduce natural atomic orbital derived point charge (NAO-PC) and ab initio molecular potentials well. The localised orbital technique (LMO-PC) is intuitively more attractive than the hybrid orbital-point charge (HO-PC) method, although the former is more CPU-intensive.Electronic Supplementary Material available.  相似文献   

15.
We used the nonlinear Poisson-Boltzmann equation to calculate electrostatic potentials in the aqueous phase adjacent to model phospholipid bilayers containing mixtures of zwitterionic lipids (phosphatidylcholine) and acidic lipids (phosphatidylserine or phosphatidylglycerol). The aqueous phase (relative permittivity, epsilon r = 80) contains 0.1 M monovalent salt. When the bilayers contain < 11% acidic lipid, the -25 mV equipotential surfaces are discrete domes centered over the negatively charged lipids and are approximately twice the value calculated using Debye-Hückel theory. When the bilayers contain > 25% acidic lipid, the -25 mV equipotential profiles are essentially flat and agree well with the values calculated using Gouy-Chapman theory. When the bilayers contain 100% acidic lipid, all of the equipotential surfaces are flat and agree with Gouy-Chapman predictions (including the -100 mV surface, which is located only 1 A from the outermost atoms). Even our model bilayers are not simple systems: the charge on each lipid is distributed over several atoms, these partial charges are non-coplanar, there is a 2 A ion-exclusion region (epsilon r = 80) adjacent to the polar headgroups, and the molecular surface is rough. We investigated the effect of these four factors using smooth (or bumpy) epsilon r = 2 slabs with embedded point charges: these factors had only minor effects on the potential in the aqueous phase.  相似文献   

16.
Partial charges of atoms in a molecule and electrostatic potential (ESP) density for that molecule are known to bear a strong correlation. In order to generate a set of point‐field force field parameters for molecular dynamics, Kollman and coworkers have extracted atomic partial charges for each of all 20 amino acids using restrained partial charge‐fitting procedures from theoretical ESP density obtained from condensed‐state quantum mechanics. The magnitude of atomic partial charges for neutral peptide backbone they have obtained is similar to that of partial atomic charges for ionized carboxylate side chain atoms. In this study, the effect of these known atomic partial charges on ESP is examined using computer simulations and compared with the experimental ESP density recently obtained for proteins using electron microscopy. It is found that the observed ESP density maps are most consistent with the simulations that include atomic partial charges of protein backbone. Therefore, atomic partial charges are integral part of atomic properties in protein molecules and should be included in model refinement.  相似文献   

17.
Phosphoenolpyruvate carboxykinase (PCK) catalyzes the conversion of oxaloacetate (OAA) to PEP and carbon dioxide with the subsequent conversion of nucleoside triphosphate to nucleoside diphosphate (NDP). The 1.9 A resolution structure of Escherichia coli PCK consisted of a 275-residue N-terminal domain and a 265-residue C-terminal domain with the active site located in a cleft between these domains. Each domain has an alpha/beta topology and the overall structure represents a new protein fold. Furthermore, PCK has a unique mononucleotide-binding fold. The 1.8 A resolution structure of the complex of ATP/Mg(2+)/oxalate with PCK revealed a 20 degrees hinge-like rotation of the N- and C-terminal domains, which closed the active site cleft. The ATP was found in the unusual syn conformation as a result of binding to the enzyme. Along with the side chain of Lys254, Mg(2+) neutralizes charges on the P beta and P gamma oxygen atoms of ATP and stabilizes an extended, eclipsed conformation of the P beta and P gamma phosphoryl groups. The sterically strained high-energy conformation likely lowers the free energy of activation for phosphoryl transfer. Additionally, the gamma-phosphoryl group becomes oriented in-line with the appropriate enolate oxygen atom, which strongly supports a direct S(N)2-type displacement of this gamma-phosphoryl group by the enolate anion. In the 2.0 A resolution structure of the complex of PCK/ADP/Mg(2+)/AlF(3), the AlF(3) moiety represents the phosphoryl group being transferred during catalysis. There are three positively charged groups that interact with the fluorine atoms, which are complementary to the three negative charges that would occur for an associative transition state.  相似文献   

18.
Given the electronic charge parameters obtained from a diffraction study of the charge density distribution in a crystal, a mathematical procedure is presented for deriving the electrostatic potential. The procedure allows the mapping of electrostatic potential for a molecule or group of molecules removed from the crystal structure but with each molecule retaining the effects of polarization owing to the original crystal environment. The method is applied for the neurotransmitter gamma-aminobutyric acid. The potential for a gamma-aminobutyric acid molecule is analyzed in terms of a simple model that is suitable for rapid computations concerned with Coulombic molecular interactions. Outside the molecular envelope at 1.2 A from the atomic nuclei, the total potential is well represented by a sum of spherical atom contributions with V(r) = (q/r) + exp(-beta r2). The most important aspherical component in the potential is the dipole contribution from the hydrogen atoms. This can be represented as V(r, phi) = (0.162 cos phi)/(r2 + 0.615). Here, V is in e/A, r is the distance from each nucleus in A, q is the experimentally determined net atomic charge in electron units, and phi is the angle between r and the bond X-H. We obtain beta = 1.47, 1.66, 1.83 A-2 for C, N, and O respectively. For H, no term in beta is needed.  相似文献   

19.
Sun Y  Soloway RD  Han YZ  Yang GD  Wang XZ  Liu ZJ  Yang ZL  Xu YZ  Wu JG 《Steroids》2002,67(5):385-392
The crystal structure of cesium cholate, C(24)H(36)(OH)(3) COOCs has been determined with three-dimensional X-ray diffractometer data. It crystallized in the monoclinic space group P2(1) with unit-cell dimensions a = 11.543(5) A, b = 8.614(3) A, and c = 12.662(5) A, beta(deg) = 107.95(2), V = 1197.7 A(3) and Z = 2. The atomic parameters were refined to a final r = 0.0269 and R(omega) = 0.0280 for 2342 observed reflections. Each Cs(+) is coordinated to 7 oxygen atoms from 5 different cholate anions with Cs-O distances ranging from 2.957(4) A to 3.678(5) A. In this crystal, 5 cholates are coordinated with 1 Cs(+), and 5 Cs(+) are coordinated with 1 cholate anion. Carboxyl and all the 3 ring hydroxyl groups of cholate anion participate in binding to Cs(+) simultaneously, and there is no water molecule coordinated with the Cs(+). The pattern of successive rows arranged with polar (p) and non-polar (n) faces in apposition leads to the formation of a sandwich sheet structure with polar and non-polar channels. The Cs ions lie within the polar interior of the sandwich. The H-bond network is reorganized in forming cesium cholate from cholic acid. All the oxygen atoms in cholate anion are involved in H-bonding reciprocally or with water molecules to form an extensive 3-dimensional network of H-bonds. Compared with cholic acid and other similar type of steroids, the coordination structure and H-bonding of Cs cholate crystal are distinct.  相似文献   

20.
O-alpha-D-Galactopyranosyl-(1---4)-D-galactopyranose, C12H22O11, Mr = 342.30, crystallises in the orthorhombic space group P2(1)2(1)2(1), and has alpha = 5.826(1), b = 13.904(3), c = 17.772(4) A, Z = 4, and Dx = 1.579 g.cm-3. Intensity data were collected with a CAD4 diffractometer. The structure was solved by direct methods and refined to R = 0.063 and Rw = 0.084 for 2758 independent reflections. The glycosidic linkage is of the type 1-axial-4-axial with torsion angles phi O-5' (O-5'-C-1'-O-1'-C-4) = 98.1(2) degrees, psi C-3 (C-3-C-4-O-1'-C-1') = -81.9(3) degrees, phi H (H-1'-C-1'-O-1'-C-4) = -18 degrees, and psi H (H-4-C-4-O-1'-C-1') = 35 degrees. The conformation is stabilised by an O-3 . . . O-5' intramolecular hydrogen-bond with length 2.787(3) A and O-3-H . . . O-5' = 162 degrees. The glycosidic linkage causes a folding of the molecule with an angle of 117 degrees between the least-square planes through the pyranosidic rings. The crystal investigated contained 56(1)% of alpha- and 44(1)% of beta-galabiose as well as approximately 70% of the gauche-trans and approximately 30% of the trans-gauche conformers about the exocyclic C-5'-C-6' and C-5-C-6 bonds. The crystal packing is governed by hydrogen bonding that engages all oxygen atoms except the intramolecular acceptor O-5' and the glycosidic O-1' oxygen atoms.  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