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.
The σ-charges on various atoms of methyl glycosides have been computed by using the MO-LCAO method of Del Re. The potential and free energies of methyl aldohexopyranosides and methyl aldopentopyranosides in their C1(d) and 1C(d) conformations have been calculated. Minimization of the energies of these conformations has been studied by suitably tilting the axial C-C and C-O bonds. Considerable release of strain is achieved when tilts of 4.5 and 3° are given to the axial hydroxymethyl and hydroxyl groups, respectively, that are involved in Hassel-Ottar effect. A tilt of 3° is also found necessary for the axial OMe group involved in the Hassel-Ottar effect. The calculated free-energy values are in accord with experimental ones, after adding a value of 0.8 kcal.mole?1 for the anomeric effect of -OMe group. These studies predict that all of the methyl aldohexopyranosides, except methyl α-d- and methyl β-d-idopyranosides, favour the C1 conformation. On the other hand, the energy calculations also predict that, of the eight methyl aldopentopyranosides studied, only methyl α-d- and methyl β-d-xylopyranosides and methyl α-d -ribopyranoside favour the C1(d) conformation; for the other pentopyranosides, considerable amounts of both C1(d) and 1C(d) conformations are present in the equilibrium mixture. The calculated values of the percentage of α-anomer present in the equilibrium mixture agree fairly well with those obtained experimentally.  相似文献   

2.
Ab initio RHF/4–31G molecular-orbital calculations have been conducted on methoxymethyl formate and methoxymethyl acetate as models for examining the anomeric effect and stereochemistry of 1-O-acetylglycopyranoses. The results indicate that, as with the methyl glycopyranosides, the α-4C1(D) configurations are more stable than the β-4C1(D), except that the energy difference is more dependent on the disposition about the glycosidic bond. The lowest-energy conformations occur with glycosidic torsion-angles of ?  180°, where the anomeric energy is about 4 kcal/mol. There is a secondary energy-minimum at ?  90°, for which the anomeric energy is less, about 2 kcal/mol. This orientation corresponds to the conformation most commonly observed in the crystal structures of peracetylated glycopyranoses. Small differences in the CO single-bond lengths, which are observed experimentally in both the α and β anomers, are reproduced by the theoretical calculations.  相似文献   

3.
DL-1,2,4,5,6-Penta-O-acetyl-myo-inositol was resolved into antipodes through the salts of its acid oxalate with quinidine and (?)-α-phenethylamine. The optically active pentaacetates were regenerated from the oxalates by lead tetraacetate oxidation. The optical purity of these pentaacetates was confirmed by their transformation into optically pure 1D- and 1L-myo-inositol 1-phosphates. The 1D-isomer of the pentaacetate was used as starting compound for the synthesis of a mixed acid phosphatidylinositol having the natural configuration at all the asymmetric centres and the normally predominating 1-saturated 2-unsaturated distribution of the fatty acid residues. From the 1L-pentaacetate an isomer of phosphatidylinositol with the unnatural 1L-configuration of the myo-inositol residue has been obtained.  相似文献   

4.
Crystal structure of a lysozyme-tetrasaccharide lactone complex   总被引:6,自引:0,他引:6  
The binding of a proposed transition-state analogue, the δ-lactone derived from tetra-N-acetylchitotetraose, to lysozyme in the crystal at pH 2.6 has been studied by X-ray diffraction techniques to a resolution of 2.5 Å. The tetrasaccharide lactone is bound in sites A, B, C, D with sugar residues located in sites A, B and C in similar positions to those observed previously in the complex with tri-N-acetylchitotriose. Analysis of the electron density map for site D, by direct model-building and with a computer model-building programme, indicates that the δ-lactone ring is in a conformation close to a sofa or a boat which brings the hydroxymethyl group C(6)O(6) axial. These studies provide support for the role of strain in the proposed mechanism of lysozyme catalysis. The orientation of the lactone group in site D is slightly different from that originally derived by hypothetical model-building.  相似文献   

5.
The conformational characteristics of the peptide sequence X-l-Pro, where X  Gly or l-Ala and the peptide bond joining X and l-Pro is cis, are evaluated. Semi-empirical potential functions are used to estimate the contributions to the conformational energy made by the non-bonded van der Waals' and electrostatic interactions and the intrinsic torsional potentials about the NCa and CaC′ bonds. Rotations φ1 and ψ1 about the NCa and CaC′ bonds in residue X and rotation ψ2 about the CaC′ bond in l-Pro are permitted, while the angle of rotation φ2 about the NCa bond in l-Pro is fixed at 120 ° by the pyrrolidine ring. The presence of the cis peptide bond connecting X and l-Pro renders the backbone rotations φ1, ψ1 in X dependent upon the rotation ψ2 about the CaC′ bond in l-Pro. (Interdependence of rotations in neighboring residues joined by a cis peptide bond was previously observed in l-alanine oligomers.) The number of energetically allowed conformations for the Gly and l-Ala residues preceding a cis peptide bond l-Pro residue are found to be substantially reduced from those permitted when the peptide bond is trans or when l-Pro is replaced by an amino acid residue. On the other hand, ψ2 = 100 to 160 ° (cis′) and 300 to 0 ° (trans′) are found to be the lowest energy conformations of the l-Pro residue irrespective of the cis or trans conformation of the X-l-Pro peptide bond.  相似文献   

6.
The monomeric composition of the suberins from 16 species of higher plants was determined by chromatographic methods following depolymerization of the isolated extractive-free cork layers with sodium methoxide-methanol. 1-Alkanols (mainly C18C28), alkanoic (mainly C16C30), α,ω-alkanedioic (mainly C16C24), ω-hydroxyalkanoic (mainly C16C21), dihydroxyhexadecanoic (mainly 10,16-dihydroxy- and 16-dihydroxyhexadecanoic), monohydroxyepoxyalkanoic (9,10-epoxy-18-hydroxyoctadecanoic), trihydroxyalkanoic (9,10, 18-trihydroxyoctadecanoic), epoxyalkanedioic (9,10-epoxyoctadecane-1,18-dioic) and dihydroxyalkanedioic (9,10-dihydroxyoctadecane-1 18-dioic) acids were detected in all species. The suberins differed from one another mainly in the relative proportions of these monomer classes and in the homologue content of their 1-alkanol, alkanoic, α,ω-alkanedioic and ω-hydroxyalkanoic acid fractions. C18 epoxy and vic-diol monomers were major components (32–59%) of half of the suberins examined (Quercus robur, Q. ilex, Q. suber, Fagus sylvatica, Castanea sativa, Betula pendula, Acer griseum, Fraxinus excelsior) where as ω-hydroxyalkanoic and α,ω-alkanedioic acids predominated in those that contained smaller quantities of such polar C18 monomers (Acer pseudoplatanus, Ribes nigrum, Euonymus alatus, Populus tremula, Solanum tuberosum, Sambucus nigra, Laburnum anagyroides, Cupressus leylandii). All species, however, contained substantial amounts (14–55 %) of ω-hydroxyalkanoic acids, the most common homologues being 18:1 (9) and 22: 0. The dominant α,ω-alkanedioic acid homologues were 16: 0 and 18: 1 (9) whereas 22: 0, 24: 0 and 26: 0, and 20: 0, 22: 0 and 24: 0 were usually the principal homologues in the 1-alkanol and alkanoic acid fractions, respectively. The most diagnostic feature of the suberins examined was the presence of monomers greater than C18 in chain length; most of the C16 and C18 monomers identified in the suberins also occur in plant cutins emphasizing the close chemical similarity between the two anatomical groups of lipid biopolymer.  相似文献   

7.
《Inorganica chimica acta》1988,141(2):253-261
Even though the α-diimino complexes [MCl2(RNCHCHNR)] and [MCl2(py-2-CHNR)] (M=Pd, Pt;R=C6H4OMe-p) are poorly soluble in chlorinated solvents, such as chloroform and 1,2-dichloroethane, or in acetonitrile, the electronic and 1H NMR spectra indicate that these compounds are generally present as undissociate monomers with σ, σ′-N,N′ chelate N-ligands in dilute solutions. Only for [PdCl2(RNCHCHNR)], some dissociation of the α-diimine occurs in acetonitrile. In dimethylsulfoxide, where the solubility is much higher, no dissociation is observed for the pyridine-2-carbaldimine complexes [MCl2(py-2-CHNR)], whereas the 1,2-bis(imino) ethane derivatives [MCl2(RNCHCHNR)] are extensively dissociated through a step-wise process involving intermediates with a σ-N monodentate α-diimino group. As is shown by the course of substitution reactions with 2,2′-bipyridine, the higher stability of [MCl2(py-2-CHNR)] in dimethylsulfoxide is mainly due to thermodynamic factors (ground state stabilization for the presence of stronger MN bonds) rather than by kinetic factors (higher activation energy for steric strain in the activation states or transients).  相似文献   

8.
To assign the observed vibrationsl modes in the resonance Raman spectrum of the retinylidene chromophore of rhodopsin, we have studied chemically modified retinals. The series of analogs investigated are the n-butyl retinals substituted at C9 and C13. The results obtained for the 11-cis isomer have clearly assigned the CCH3 vibrational frequencies observed in the spectrum of the retinylidene chromophore. The data show that the C(9)CH3 stretching vibration can be assigned to the vibrational mode observed in the 1017 cm?1 region, and the vibration detected at 997 cm?1 can be assigned to the C(13CH3 vibration. The C(5)CH3 stretching mode does not contribute to the vibrations observed in this region. The splitting in the C(n)CH3 (n = 9, 13) vibration is characteristic of the 11-cis conformation. The results on the modified retinals do not support the hypothesis that the splitting arises from equilibrium mixtures of 11-cis, 12-s-cis and 11-cis, 12-s-trans in solution. Thus, this splitting cannot be used to determine whether the chromophore in rhodopsin is in a 12-s-cis or 12-s-trans conformation. However, our results demonstrate that there are other vibrational modes in the spectra which are sensitive to this conformational equilibrium and we use the presence of a strong ~ 1271 cm?1 mode in bovine and squid rhodopsin spectra as an indication that the chromophore in these pigments is 11-cis, 12-s-trans.  相似文献   

9.
Intramolecularly hydrogen bonded conformations of (Aib-Pro)n sequences have been analysed theoretically. Both 4→1 (C10 and 3→1 (C7 hydrogen bonded regular structures are shown to be stereochemically feasible. Conformational energies for the helical structures have been estimated using classical potential energy methods. Both C10 and C7 conformations have very similar energies. Pyrrolidine ring puckering has a pronounced effect on the energies, and only -endo puckered Pro residues can be accommodated. The theoretical calculations using spectroscopic data suggest that the recently proposed novel 310 helical conformation for benzyloxycarbonyl(Aib-Pro)4-methyl ester is in solution, is indeed energetically and stereochemically favourable.  相似文献   

10.
The synthesis of 1,3,4,6-tetra-O-acetyl-2-deoxy-2-(N-nitroso)acetamido-α- and β-D-glucopyranose is described. Decomposition of the α-nitrosoamide in chloroform containing 2% of ethanol at room temperature afforded β-D-glucopyranose pentaacetate and ethyl β-D-glucopyranoside tetraacetate as major products, the former predominating. Reaction in 1:5 (v/v) acetic acid—acetic anhydride containing sodium acetate also gave β-D-glucose pentaacetate as major product, together with 1,1,3,4,6-penta-O-acetyl-2,5-anhydro-D-mannose aldehydrol. Decompositions of both α and β-nitrosoamides in 1:1 (v/v) acetone—water gave mainly 3,4,6-tri-O-acetyl-2,5-anhydro-D-mannose and its aldehydrol form. The synthesis, from 2,5-anhydro-D-mannose, of four new derivatives is also reported.  相似文献   

11.
12.
《Inorganica chimica acta》1986,125(3):159-166
This paper describes the synthesis of compounds of Ni(II) with uracil, uridine and the nucleotides 5′UMP, 5′CMP, 5′GMP and 5′IMP, and their characterization, carried out by elemental analysis, by studying the infrared spectra, diffuse reflectance and conductivity measurement.In the complexes of NiURA (and NiURD) with acetate, direct coordination of the metal ion to the C4O group of the pyrimidine ring is inferred from the changes observed on the infrared spectrum of the corresponding bands at vCO. The variations in frequency of the vCOO symmetric and asymmetric bands of the acetate group together with the conductivity and reflectance results seem to indicate the dimer structure of the compounds.In the compounds of NiURA (and NiURD) with ethylenediamine indirect bonding of Ni(II)to the pyrimidine ring is inferred, probably established through hydrogen bonds involving the C4O groups in the base or nucleoside and the −NH2 groups in the ethylenediamine.In the complexes of Ni-nucleotide, bonding seems to occur through the heterocyclic ring (C4O for 5′UMP, N(3) for 5′CMP, N(7) for 5′GMP and 5′IMP) together with additional interactions through the phosphate group.  相似文献   

13.
2,3-O-Isopropylidene-d-ribose diethyl dithioacetal, prepared from d-ribose, was converted in three steps into the corresponding dimethyl acetal, which was monotosylated at O-5, and the ester oxidized at C-4 with pyridinium chlorochromate; addition of methyl phenylphosphinate to the resulting pentos-4-ulose derivative then provided (4R,S)-4,5-anhydro-2,3-O-isopropylidene-4-C-[(R,S)-(methoxy)phenylphosphinyl]-d-erythro-pentose dimethyl acetal. Hydrogenation of this compound in the presence of Raney Ni, followed by reduction with SDMA, hydrolysis, and acetylation, yielded the title compounds (seven kinds), the structures of which were established on the basis of their 400-MHz, 1H-n.m.r. and mass spectra. A general dependence of the 2JPH and 3JPH values on the OPCH and PCCH dihedral angles provided an effective method for the assignment of the configurations and conformations of these 4-deoxy-4-phosphinyl-pentofuranoses.  相似文献   

14.
Feeding of benzoic acid-[7-14C], benzaldehyde-[7-14C] and cinnamic acid-[3-14C] to Ephedra distachya resulted in efficient incorporations of 14C into the α-carbon atom of the side chain of l-ephedrine. Thus ephedrine was shown to be biosynthesized by the condensation of a C6C1 portion which is derived from phenylalanine via cinnamate and an unidentified C2-N fragment.  相似文献   

15.
The crystal structures of the organocobalt complexes, pyCo(GH)2Me(1), pyCo(GH)2Et(2) and pyCo(GH)2Pri(3) (py = pyridine, GH = monoanion of glyoxime) are reported. Compound (1) crystallizes in the space group P212121 with cell parameters a = 8.508(1), b = 13.586(2) and c = 11.614(6) Å; (2) crystallizes in the space group P212121 with cell parameters a = 8.448(4), b = 12.164(2) and c = 13.651(2) Å; (3) crystallizes in the space group P21/c with cell parameters a = 8.443(7), b = 12.913(2), c = 14.341(2) Å and β = 92.86(4).The three structures have been solved by Patterson and Fourier methods and refined by least squares methods to final R values of 0.045(1), 0.068(2) and 0.057(3) using 1819(1), 1653(2) and 1582(3) independent reflections. The pyCoalkyl fragment shows significant variation of CoN and CoC bond lengths. The latter increase from 2.003(4) to 2.084(9) Å following the increase of the alkyl bulk. The CoN(py) distances increase from 2.064(3) to 2.101(6) Å with the increasing σ-donor power of the alkyl group trans to pyridine. In comparison with cobaloximes having the same axial ligands, pyCo(DH)2alkyl (DH = monoanion of dimethylglyoxime) does not show significant differences on the pyCo alkyl fragment. CoN axial bond lengths and exchange rates of the axial neutral ligand are consistent for the two series, although changes in bond lengths are detected only when rate constants are from two to three orders of magnitude different.  相似文献   

16.
Starting from single molecule calculations the intermolecular interactions of the glycerophosphatidylcholine (GPC) headgroup with its nearest neighbours in a layer crystal were taken into account using 1-6-12 interatomic potential functions. By use of a steepest descent energy minimisation procedure over all variable torsion angles (θ1, α1α6) of the GPC headgroup the minima of the seven-dimensional energy hypersurface were calculated. The torsion angles and the energies of the most stable conformations are given in polar coordinates. The components of the headgroup dipolar moment of these conformations were calculated to be μ6 = 3.0 … 9.5 D, μ = 17.5 … 24 D, μ = 18.5 … 25 D using the net atomic charge distribution in space. The results demonstrate a high flexibility of the GPC headgroup. Around the C-1O-11 bond (α1), only antiperiplanar conformations are allowed. The α4α56) correlation diagram shows that the choline group exists in mirror-image enantiomeric conformations. Our results yield a foundation of models of the dynamical behaviour of the phosphatidylcholine headgroup at the level of conformational behaviour and are in agreement with experimental data.  相似文献   

17.
The present study is the first report demonstrating that ionic liquids consisting of cholinium cations and linear carboxylate anions ([Ch][CA] ILs) can be used for pretreatment of lignocellulosic materials to enhance subsequent enzymatic saccharification. Six variants of [Ch][CA] ILs were systematically prepared by combining cholinium cations with linear monocarboxylate anions ([CnH2n+1–COO], n = 0–2) or dicarboxylate anions ([HOOC–CnH2n+1–COO], n = 0–2). These [Ch][CA] ILs were analyzed for their toxicity to yeast cell growth and their ability to pretreat kenaf powder for subsequent enzymatic saccharification. When assayed against yeast growth, the EC50 for choline acetate ([Ch][OAc]) was 510 mM, almost one order of magnitude higher than that for 1-ethyl-3-methylimidazolium acetate ([Emim][OAc]). The cellulose saccharification ratio after pretreatment at 110 °C for 16 h with [Ch][OAc] (100.6%) was almost comparable with that after pretreatment with [Emim][OAc]. Therefore, [Ch][OAc] is a biocompatible alternative to [Emim][OAc] for lignocellulosic material pretreatment.  相似文献   

18.
Conformational analysis of muscimol, a GABA agonist   总被引:3,自引:0,他引:3  
The potential energy barriers for rotation around the C5C6 bond in muscimol and two related isoxazoles have been calculated using the MINDO/3 molecular orbital method. The preferred conformations have N7C6C5C4 torsion angles near ± 100 °, in agreement with crystallographic data. The activities of muscimol and related isoxazoles as bicuculline-sensitive inhibitors of neuronal firing, however, are best accounted for in terms of “active conformations” with N7C6C5C4 torsion angles in the range +(32–46) °. These findings predict a limited range of possible “active conformations” for the flexible neurotransmitter GABA at postsynaptic receptors common to GABA, (+)-bicuculline salts and muscimol.  相似文献   

19.
《Inorganica chimica acta》1988,146(1):135-142
The proton NMR spectra of ethylenediamine-NN′- diacetate-NN′-di-3-propionate and its complexes with alkaline earth and diamagnetic lanthanide ions are described. Quartet splittings of the methylenic protons of the acetate groups upon complexation with metal ions of high charge density is indicative of long-lived metalnitrogen bonds and short-lived metaloxygen bonds. The observation of two AB quartets for the acetate protons complexed to trivalent ions was attributed to an unsymmetrical distribution of the acetate arms around the central ion. For Mg2+ complex, the two acetate arms are equivalently disposed and the two quartets collapse into one. For the other alkaline earths, a singlet is observed, indicating that the metalnitrogen bond lifetime is also short. The spectra of the propionate protons are consistent of either an ABCD (LuENDPDA and YENDPDA) or an AA′BB′ (MgENDPDA) spectral pattern. The thermodynamic data support these conclusions and show that substitution of C2H4CO2 groups for CH2CO2 groups decreases the stability of the complex. This results from weakening of the metalnitrogen interaction due to the expansion of the
  相似文献   

20.
《Inorganica chimica acta》1988,149(2):193-208
The reactions of Fe(CO)3(R-DAB; R1, H(4e)) (1a: R = i-Pr, R1 = H; 1b: R = t-Bu, R1 = H; 1c: R = c-Hex, R1 = H; 1e: R = p-Tol, R1 = H; 1f: R = i-Pr, R1 = Me) with Ru3(CO)12 and of Ru(CO)3(R-DAB; R1, H(4e)) (2a: R = i-Pr, R1 = H; 2d: R = CH(i-Pr)2, R1 = H) with Fe2(CO)9 in refluxing heptane both afforded FeRu(CO)6(R-DAB; R1, H(6e)) (3) in yields between 50 and 65%.The coordination mode of the ligand has been studied by a single crystal X-ray structure determination of FeRu(CO)6(i-Pr-DAB(6e)) (3a). Crystals of 3a are monoclinic, space group P21/a, with four molecules in a unit cell of dimensions: a = 22.436(3), b = 8.136(3), c = 10.266(1) Å and β = 99.57(1)°. The structure was refined to R = 0.049 and Rw = 0.052 using 3045 reflections above the 2.5σ(I) level. The molecule contains an FeRu bond of 2.6602(9) Å, three terminally bonded carbonyls to Fe, three terminally bonded carbonyls to Ru and bridging 6e donating i-Pr-DAB ligand. The i-Pr-DAB ligand is coordinated to Ru via N(1) and N(2) occupying an apical and equatorial site respectively (RuN(1) = 2.138(4) RuN(2) = 2.102(3) Å). The C(2)N(2) moiety of the ligand is η2-coordinated to Fe with C(2) in an apical and N(2) in an equatorial site (FeC(2) = 2.070(5) and FeN(2) = 1.942(3) Å).The 1H and 13C NMR data indicate that in all FeRu(CO)6(R-DAB(6e)) complexes (3a to 3f) exclusively η2-CN coordination to the Fe atom and not to the Ru atom is present irrespective of whether 3 was prepared by reaction of Fe(CO)3(R-DAB(4e)) (1) with Ru3(CO)12 or by reaction of Ru(CO)3(R-DAB(4e)) (2) with Fe2(CO)9. In the case of FeRu(CO)6(i-Pr-DAB; Me, H(6e)) (3f) the NMR data show that only the complex with the C(Me)N moiety of the ligand σ-N coordinated to the Ru atom and the C(H)N moiety η2-coordinated to the Fe atom was formed. Variable temperature NMR experiments up to 140 °C showed that the α-diimine ligand in 3a is stereochemically rigid bonded.FeRu(CO)6(R-DAB(6e)) (3a and 3e) reacted with allene to give FeRu(CO)5(R-DAB(4e))(C3H4) (4a and 4e). A single crystal X-ray structure determination of FeRu(CO)5(i-Pr-DAB(4e))(C3H4) (4a) was performed. Crystals of 4a are triclinic, space group P1, with two molecules in a unit cell of dimensions: a = 9.7882(7), b = 12.2609(9), c = 8.3343(7) Å, α = 99.77(1)°, β = 91.47(1)° and γ = 86.00(1)°. The structure was refined to R = 0.028 and Rw = 0.043 using 4598 reflections above the 2σ(I) level. The molecule contains an FeRu bond of 2.7405(7) Å and three terminally bonded carbonyls to iron. Two carbonyls are terminally bonded to the Ru atom together with a chelating 4e donating i-Pr-DAB ligand [RuN = 2.110(1) (mean)]. The allene ligand is coordinated in an η3-allylic fashion to the Fe atom while the central carbon of the allene moiety is σ-bonded to the Ru atom (FeC(14) = 2.166(3), FeC(15) = 1.970(2), FeC(16) = 2.127(3) and RuC(15) = 2.075(2) Å). The 1H and 13C NMR data show that in solution the coordination modes of the R-DAB and the allene ligands are the same as in the solid state.Thermolysis reactions of 3a with R-DAB or carbodiimides gave decomposition and did not afford C(imine)C(reactant) coupling products. Thermolysis reactions of 3a with M3(CO)12 (M = Ru, Os) and Me3NO gave decomposition. When the reaction of 3a with Me3NO was performed in the presence of dimethylacetylenedicarboxylate (DMADC) the known complex FeRu(CO)4(i-Pr-DAB(8e))(DMADC) (5a) was formed in low yield. In 5a the R-DAB ligand is in the 8e coordination mode with both the imine bonds η2-coordinated to iron. The acetylene ligand is coordinated in a bridging fashion, parallel with the FeRu bond.  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