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 140 毫秒
1.
The sidechain of the germacranolide stizolicin was oxidized by CrO3 in Py to form a lactone. The structure of the product was established as 4,5α-epoxy-7α,5,6,8β(H)-germacr-1(10),11(13)-dien-8α (3′-oxo-2′,5′-dihydrofuran-3′-carboxylyl)-12,6-olide using spectral data.__________Translated from Khimiya Prirodnykh Soedinenii, No. 2, pp. 117–118, March–April, 2005.  相似文献   

2.
Selective formation of (η3-siloxyallyl)tungsten complexes by reaction of hydrido(hydrosilylene)tungsten complexes with α,β-unsaturated carbonyl compounds was reported experimentally. The mechanisms have been investigated by employing the model reaction of [Cp(CO)2(H)WSi(H)–{C(SiH3)3}] (R), derived from the original experimental complex Cp′(CO)2(H)WSi(H)–[C(SiMe3)3] (1a, Cp′ = Cp*; 1b, Cp′ = η5-C5Me4Et), with methyl vinyl ketone, under the aid of the density functional calculations at the b3lyp level of theory. It is theoretically predicted that the route involving migration of the hydride to silicon to afford a 16e intermediate [Cp(CO)2W–SiH2–{C(SiH3)3}] is inaccessible (route 2), supporting the proposition by experiments. Another route, via [2 + 4] cycloaddition followed by directly Si–H reductive elimination, is theoretically predicted to be accessible (route 1). In route 1, two possible paths with different attacking directions of the oxygen of methyl vinyl ketone at Si (WSi) are put forward. The attack at the Si atom from the hydride (H1) side of the plane W–Si–H1 in R is found to be preferred kinetically. The regioselectivity for formation of (η3-siloxyallyl)tungsten complexes, where only the exo-anti isomer was obtained, is discussed based on the consideration of thermodynamics and kinetics.  相似文献   

3.
The melting points of anhydrous 1-O-α- -glucopyranosyl- -mannitol, 1-O-α- -glucopyranosyl- -mannitol dihydrate and a new compound, 1-O-α- -glucopyranosyl- -mannitol-ethanol (2/1) were determined using differential scanning calorimetry. The melting onset values were 169.2 (3), 104.3 (18) and 158.7 (9), respectively, and the melting peak values were 171.4 (5), 107.9 (15) and 160.1 (6), respectively. 1-O-α- -glucopyranosyl- -mannitol dihydrate and 1-O-α- -glucopyranosyl- -mannitol-ethanol (2/1) decompose to anhydrous form when heated at slow heating rates.According to TG-FTIR measurements, 1-O-α- -glucopyranosyl- -mannitol-ethanol (2/1) lost its ethanol in the 110–190°C range, and 1-O-α- -glucopyranosyl- -mannitol dihydrate lost its crystal water in the 60–210°C range. After removal of ethanol and crystal water, both decomposed in air totally as carbohydrates usually do, forming lower hydrocarbons with OH-groups, CO2 and H2O.  相似文献   

4.
Hydrated powders of non-gelatinised starch and hydrogels of gelatinised starch alone or with addition of modified nanosilica (with grafted aminopropylmethylsilyl groups substituting one-third of surface silanols) were studied using broadband dielectric relaxation spectroscopy (DRS), thermally stimulated depolarisation current (TSDC) method and 1H NMR spectroscopy with layer-by-layer freezing-out of bulk and interfacial waters. The 1H NMR and TSDC techniques with the use of Gibbs–Thomson relation for the freezing point depression allow us to calculate: (i) the thermodynamic parameters of interfacial water weakly and strongly bound to polymer molecules and nanoparticles; (ii) size distributions of pores filled by structured water; (iii) surface area and volume of micro-, meso- and macropores. The DRS and TSDC results for hydrogels and hydrated powders with starch/modified fumed silica show that the β- and γ-relaxations of starch are strongly affected by water and functionalised silica nanoparticles which slow down both low- and high-frequency and low- and high-temperature relaxations.  相似文献   

5.
The ligands [Ph2P(O)NP(E)Ph2] (E=S I; E=Se II) can readily be complexed to a range of palladium(II) starting materials affording new six-membered Pd–O–P–N–P–E palladacycles. Hence ligand substitution reaction of the chloride complexes [PdCl2(bipy)] (bipy=2,2′-bipyridine), [{Pd(μ-Cl)(L–L)}2] (HL–L=C9H13N or C12H13N), [{Pd(μ-Cl)Cl(PMe2Ph)}2] or [PdCl2(PR3)2] [PR3=PPh3; 2PR3=Ph2PCH2CH2PPh2or cis-Ph2PCH=CHPPh2] with either I (or II) in thf or CH3OH gave [Pd{Ph2P(O)NP(E)Ph2-O,E}(bipy)]PF6, [Pd{Ph2P(O)NP(E)Ph2-O,E}(L–L)], [Pd{Ph2P(O)NP(E)Ph2-O,E}Cl(PMe2Ph)] or [Pd{Ph2P(O)NP(E)Ph2-O,E} (PR3)2]PF6 in good yields. All compounds described have been characterised by a combination of multinuclear NMR [31 P{1 H} and 1 H] and IR spectroscopy and microanalysis. The molecular structures of five complexes containing the selenium ligand II have been determined by single-crystal X-ray crystallography. Three different ring conformations were observed, a pseudo-butterfly, hinge and in the case of all three PR3 complexes, pseudo-boat conformations. Within the Pd–O–P–N–P–Se rings there is evidence for π-electron delocalisation.  相似文献   

6.
β-Cyclodextrin (β-CD) was used to complex the photoinitiator, 5-thia pentacene-14-one (TX-A), yielding a water-soluble host/guest complex. IR, UV–Vis and fluorescence spectroscopy were employed to characterize complexed β-CD/TX-A. Photoinitiated polymerization of acrylamide in water was achieved with β-CD/TX-A in the presence of N-methyldiethanolamine (MDEA). Excellent polymerization yields were observed in air saturated solutions when MDEA was added.  相似文献   

7.
The acidity of the free base (H2–P(X)) forms of the tetra- (N-methyl-4–(3 or 2)pyridyl)porphyrins were studied in basic solutions at 25 C, I=0.50. Equilibrium constants for both the H2– P(X)=P(X)2–+2H+ and the Li++ P(X)2–=Li- P(X)– reactions are reported. Log (KS) values for the Mn++P(4)2–=M- P(4)(n–2) reactions are 2.6 for Li+, 17.6 for Cd2+, 17.8 for Pb2+, 19.6 for (OH)2Hg, 25.9 for Zn2+ and 19.6 for the formation of Hg2–P(4)2+. Zn(OH)3– shows similar kinetic reactivity with both H2–P(4) and P(4)2– to form Zn- P(4) and HO-Zn- P(4)–,whileZn(OH)42– is unreactive with either species. For Zn2+, Zn(OH)+ and Zn(OH)3– with H2–P(4) the relative kinetic order for this tetrapositive macrocycle was ca. 1:300:20,000, while the trend Zn(OH)+>Zn2+>Zn(OH)3– is the usual pattern for peripherally negatively charged porphyrins.  相似文献   

8.
The mononuclear fragments [Cu(HDopn)(OH)2]+ and [Cu(HPopn)(OH)2]+, [H2Dopn=3,3-(trimethylene- dinitrilo)-dibutan-2–one dioximate and H2Popn, = 3, 3-(phenylenedinitrilo)-dibutan-2–one dioximate] were used to prepare four binuclear complexes [(OH2)Cu (Dopn)Cu(ditn)]2+, [(OH2)Cu(Dopn)Ni(ditn)(H2O)]2+ (ditn=diethylenetriamine) and [(OH2)Cu(Popn)Cu(L) (H2O)]2+ (L=2,2-bipyridine or 1,10–phenanthroline). Two trinuclear complexes, [{Cu(Popn)(OH2)}2M (H2O)n]2+ (when M=CuII, n=1; M=ZnII, n= 2), have been synthesised and characterised by elemental analyses, f.a.b. mass, i.r., electronic, e.s.r. spectroscopy and variable temperature (5–300K) magnetic susceptibility measurements. A strong antiferromagnetic interaction (J=–545cm–1 to –700cm–1) has been found for the binuclear copper(II) complexes. The X-band e.s.r. spectra of these complexes at 300K and for trinuclear complexes at 120K indicate square-pyramidal geometry for the copper centres with a (dx2–y2)1 ground state. The binuclear complex of copper(II)–nickel(II) centres with antiferromagnetic interaction (J=–107 cm–1) is described, and moderately strong zero-field splitting within the quartet state leads to Kramers doublet, as indicated by X-band e.s.r. spectra of this complex. The trinuclear copper(II) complex with an antiferromagnetic interaction (J= –350cm–1) is also described. The heterometallic trinuclear copper(II)–zinc(II)–copper(II) system shows a very weak interaction (J–1cm–1).  相似文献   

9.
The metal contents of Nigerian coal minerals were analyzed using an atomic absorption spectrophotometer. Calcium, Na, and Fe occurred as the major elements with concentrations ranging from 9 782 μg/g for Ca to 432 μg/g for Na whereas K, Mg, Mn, Ni, Cr, Zn, Pb, and Cu, which occurred at trace levels ranged from 673.73 μg/g for Mg to 2.97 μg/g for Mn. The results of the quantitative analysis of porphyrins extracted from the coal minerals showed that Onyeama coal has the highest amount of porphyrins (ca~0.96 μg/g) while Okpara has the lowest (ca~0.30 μg/g). The porphyrins were qualitatively characterized by a combination of thin layer chromatography (TLC), infrared, and ultraviolet-visible spectrophotometers. The results of the mid infrared analysis (MIR) showed the presence of absorption bands at 3 440 cm~1~3 450 cm-1 and 1 640 cm-1~1 680 cm-1 , which are owing to the stretching vibrations of N - H and C = C of aromatics, with C- H out of plane (oop) bending vibrations at wavenumbers less than 900 cm-1, all of which are characteristic absorptions of porphyrin free base. The ultraviolet-visible data showed prominent peaks at ~400 nm(Soret band) and at wavelength ranges of 535 nm~550 nm(β-band) and 565 nm~ 600 nm (α-band) for the coal porphyrins analyzed. The geochemical significance of the metals and porphyrins in coal minerals are discussed.  相似文献   

10.
Complexation of ketoconazole (KET), a broad-spectrum antifungal drug, with β- and γ-cyclodextrins (CDs), heptakis (2,6-di-O-methyl)-β-CD (2,6-DM-β-CD), heptakis (2,3,6-tri-O-methyl)-β-CD (TM-β-CD), 2-hydroxypropyl-β-CD (2HP-β-CD) and carboxymethyl-β-CD (CM-β-CD) was studied. The stability constants were determined by the solubility method at pH = 6 and for 2,6-DM-β-CD and CM-β-CD at pH = 5. At pH = 6, the stability constants increased in the order: TM-β-D < γ-CD < 2HP-β-CD < β-CD < CM-β-CD < 2,6-DM-β-CD. At pH = 5, due to the increased ionization of KET, the stability constant with CM-β-CD increased and with 2,6-DM-β-CD decreased. For complexes of KET with 2HP-β-CD and 2,6-DM-β-CD, the thermodynamic parameters of complexation were determined from the temperature dependence of the corresponding stability constants. For β–γ and TM-β-CD complexes, calculations using HyperChem 6 software by the Amber force field were carried out to gain some insight into the host–guest geometry.  相似文献   

11.
Fumed oxide alumina/silica/titania was studied in comparison with fumed alumina, silica, titania, alumina/silica, and titania/silica by means of XRD, 1H NMR, IR, optical, dielectric relaxation, and photon correlation spectroscopies, electrophoresis, and quantum chemical methods. The explored Al2O3/SiO2/TiO2 consists of amorphous alumina (22 wt%), amorphous silica (28 wt%), and crystalline titania (50 wt%, with a blend of anatase (88%) and rutile (12%)) and has a wide assortment of Brønsted and Lewis acid sites, which provide a greater acidity than that of individual fumed alumina, silica, or titania and an acidity close to that of fumed alumina/silica or titania/silica. The changes in the Gibbs free energy (ΔG) of interfacial water in an aqueous suspension of Al2O3/SiO2/TiO2 are close to the ΔG values of the dispersions of pure rutile but markedly lower than those of alumina, anatase, or rutile covered by alumina and silica. The zeta potential of Al2O3/SiO2/TiO2 (pH of the isoelectric point (IEP) equals ≈3.3) is akin to that of fumed titania (pH(IEPTiO2) ≈ 6) at pH > 6, but it significantly differs from the ζ of fumed alumina (pH(IEPAl2O3) ≈ 9.8) at any pH value as well as those of fumed silica, titania/silica, and alumina/silica at pH < 6. The particle size distribution in the diluted aqueous suspensions of Al2O3/SiO2/TiO2 studied by means of photon correlation spectroscopy depends relatively slightly on pH in contrast to the titania/silica or alumina/silica dispersions. Theoretical calculations of oxide cluster interaction with water show a high probability of hydrolysis of Al–O–Ti and Si–O–Ti bonds strained at the interface of alumina/titania or silica/titania due to structural differences in the lattices of the corresponding individual oxides. Ab initio calculated chemical shift δH values of H atoms in different hydroxyl groups at the oxide clusters and in bound water molecules are in agreement with the 1H NMR data and show a significant impact of charged particles (H3O+ or OH) on the average δH values of water droplets with (H2O)n at n between 2 and 48.  相似文献   

12.
New inclusion complexes(C2H5)4N+HCO2-·(NH2)2CS·H2O (1),[(C2H5)4N]2+[(HCO2)2H]-(HCO2-)·2(NH2)2CS(2), (n-C3H7)4N+HCO2-·3(NH2)2CS·H2O (3)and (n-C4H9)4N+[(HCO2)2H]-·2(NH2)2CS(4) have been prepared and characterized by X-ray crystallography. Crystal data, MoK radiation: 1, space group P21/c, a = 7.199(2), b = 16.851(2),c = 13.044(2) Å, = 100.13(2)°, Z = 4, and RF = 0.065 for 1011 observed data; 2, space group Pca21, a = 25.803(5), b = 7.190(2), c = 17.394(2) Å, Z = 4, and RF= 0.073 for 1515 observed data; 3, space group P21/n, a = 8.533(2), b = 9.423(5), c = 33.517(7) Å, = 90.44(2)°, Z = 4, and RF = 0.052 for 2521 observed data; 4, space group Pbca, a = 17.389(3), b = 16.622(2),c = 20.199(3) Å, Z = 8, and RF = 0.056for 1910 observed data. In both 1 and 2 the tetraethylammonium ions are sandwiched between puckered layers, which are constructed by the cross-linkage of a parallel arrangement of infinite chains. In 1 each chain is composed of twisted(thiourea–formate)2 tetramers bridged by water molecules, whereas in 2 it comprises an alternate arrangement of thiourea dimers and protonated formate trimers each formed by the linkage of a hydrogen diformate ion, [(HCO2)2H]-, to a formate ion via} a C–-H·sO hydrogen bond. In compound 3 two independent thiourea molecules are used to construct a hydrogen-bonded puckered layer normal to thec axis, whereas the remaining thiourea molecule, together with the formate ion and water molecule, generate another puckered layer that is parallel to the first one. Hydrogen bonding between these two types of layers gives rise to a network containing channels running parallel to the [100] direction, and the cations are stacked regularly within each column. In the crystal structure of 4, the thiourea molecules form hydrogen-bonded zigzag ribbons running parallel to the b axis, which are cross-linked by the dimeric formate moieties [(HCO2)2H]- to form a puckered layer, and the(n-C4H9)N+ cations occupy the space between adjacent layers.  相似文献   

13.
Excess molar volumes VmE of the binary mixtures of (trifluoroethanol + 1-propanol), (trifluoroethanol + 2-propanol), (acetone + water), (methanol + water), (ethanol + water), (1-propanol + water), (2-propanol + water), and the ternary mixtures of (trifluoroethanol + methanol + water), (trifluoroethanol + ethanol + water), (trifluoroethanol~+ 1-propanol + water), (trifluoroethanol + 2-propanol + water) and (trifluoroethanol + acetone + water) were measured with a vibrating tube densimeter at the temperature of 298.15 K and the pressure 101 kPa. The extrema in VmE of trifluoroethanol mixtures occur at –0.690 cm3-mol–1 for (trifluoroethanol + 1-propanol), at –0.990~cm3-mol–1 for (trifluoroethanol + 2-propanol); at 0.562 and –0.973 cm3-mol–1 for (trifluoroethanol + methanol + water), at 0.629 and –0.973 cm3-mol–1 for (trifluoroethanol + ethanol + water), at 1.082 and –0.659 cm3-mol–1 for (trifluoroethanol~+ 1-propanol + water), at 0.998 and –0.991 cm3-mol–1 for (trifluoroethanol~+ 2-propanol + water), and at 0.515 and –1.472 cm3-mol–1 for (trifluoroethanol + acetone + water). The experimental ternary VmE values were predicted by empirical expressions using binary solution data.  相似文献   

14.
The lanthanide complexes of bis(benzimidazole-2′-yl-methyl)amine (BImz) having molecular formula [M(BImz)X3H2O]·nH2O (M = La, Pr, Nd, or Gd; X = Cl or ClO4 and n = 1, 2 or 3) were prepared and characterized spectroscopically through IR, 1H and 13C NMR, FAB-mass, UV–visible and luminescence spectroscopy. TGA data suggested presence of coordinated and the lattice water. The oscillator strengths of the f–f transitions and the covalency parameters (β, b1/2 and δ) have been evaluated from the electronic spectral data. The proposed hepta-coordinate geometry for the complexes has been ascertained from the molecular model computations. CV studies indicate formation of a stable quasi-reversible redox couple GdIII/IV in the solution. The in vivo antimicrobial activities of the complexes have been evaluated against gram +ve and gram −ve bacteria and fungi.  相似文献   

15.
The synthesis of cobalt(II), nickel(II) and zinc(II) complexes of 2-pyridine-2-yl-3(pyridine-2-carboxylideneamino)-1,2-dihydroquinazolin-4(3H)-one is described. The ligand and metal complexes were characterized by elemental analysis, conductivity measurements, spectral (u.v.–vis., i.r., 1D n.m.r., 2D hetcor and mass) and thermal studies. The cobalt(II) complex crystallizes as pink crystals in the monoclinic crystal system, space group P21/n with a = 10.066(6) Å, b = 15.929(9) Å, c = 12.624(7) Å, α = 90.00(9)°, β = 110.850 (8)°, γ = 90.00, V = 1891.5 (18) Å3 and Z = 4. The geometry around the cobalt atom is distorted trigonal bipyramidal with τ = 0.83 [structural parameter, τ = (βα)/60; where α and β are the two basal angles in a five coordinate complex].  相似文献   

16.
Condensation reactions between 4'-formyl-5'-hydroxybenzo-15-crown-5and 2-aminopyridine, 2-amino-6-methylpyridine, 2-amino-4-methylpyridine or2-(aminomethyl)furan yielded the new laterally functionalized crown ethers1–4. The crown compounds 1–3 form crystalline 1:1 (Na+:ligand) complexes 1a–3a with sodium perchlorate. Ligands and complexes have been characterized by elemental analyses, IR, UV-Vis, 1H-, 13C-NMR and mass spectra. The tautomeric equilibria (phenol-imine, O...H–N and keto-amine, O...H–N forms) have been systematically studied by using UV-Vis absorption spectra. The spectra of the ligands 1–4 and complexes 1a–3a were recorded in polar, non-polar, acidic, and basic media. In solutions of polar solvents, tautomeric interconversion of the Schiff base into the keto-amine form has been observed. A crystal structure [monoclinic, space group P21/c,a = 14.292(2), b = 9.449(6), c = 16.059(2) Å, = 114.20(1)°,V = 1978.4(13) Å3, Z = 4 and Dx = 1.314 g cm-3] shows that compound 4 is in the form of phenol-imine in solid state. There is a strong intramolecular [O–H...N 1.78(6), O...N 2.581(7), O–H 0.89(6) Å and N...H–O 148.4(5)°] hydrogen bond between the phenolic oxygen and imine nitrogen atoms. The C=N imine bond reveals a trans planar (1E) configuration. The molecules stack in columns parallel to the a/c plane of the unit cell.  相似文献   

17.
In the present work, a novel heterocyclic hybrid of a spirooxindole system was synthesized via the attachment of ferrocene and triazole motifs into an azomethine ylide by [3 + 2] cycloaddition reaction protocol. The X-ray structure of the heterocyclic hybrid (1″R,2″S,3R)-2″-(1-(3-chloro-4-fluorophenyl)-5-methyl-1H-1,2,3-triazole-4-carbonyl)-5-methyl-1″-(ferrocin-2-yl)-1″,2″,5″,6″,7″,7a″-hexahydrospiro[indoline-3,3″-pyrrolizin]-2-one revealed very well the expected structure, by using different analytical tools (FTIR and NMR spectroscopy). It crystallized in the triclinic-crystal system and the P-1-space group. The unit cell parameters are a = 9.1442(2) Å, b = 12.0872(3) Å, c = 14.1223(4) Å, α = 102.1700(10)°, β = 97.4190(10)°, γ = 99.1600(10)°, and V = 1484.81(7) Å3. There are two molecules per unit cell and one formula unit per asymmetric unit. Hirshfeld analysis was used to study the molecular packing of the heterocyclic hybrid. H···H (50.8%), H···C (14.2%), Cl···H (8.9%), O···H (7.3%), and N···H (5.1%) are the most dominant intermolecular contacts in the crystal structure. O···H, N···H, H···C, F···H, F···C, and O···O are the only contacts that have the characteristic features of short and significant interactions. AIM study indicated predominant covalent characters for the Fe–C interactions. Also, the electron density (ρ(r)) at the bond critical point correlated inversely with the Fe–C distances.  相似文献   

18.
[K(18-cr-6)]2[Pt2Cl6] reacts with olefins in CH2Cl2 to give crown ether adducts of olefin complexes of Zeise's salt type [K(18-cr- 6)][PtCl3(olefin)] (2) [olefin: EtCH=CH2 (2a), cis-MeCH=CHMe (2b), trans-MeCH=CHMe (2c), nPrCH=CH2 (2d), cis-EtCH=CHMe (2e), trans-EtCH=CHMe (2f), nBuCH=CH2 (2g), cis-nPrCH=CH-Me (2h), cis-EtCH=CHEt (2i), trans-EtCH=CHEt (2j), cyclohexene (2k), PhCH=CH2 (2l)]. The complexes were characterized by microanalysis, n.m.r. (1H, 13C) and i.r. spectroscopy. The crystal structures of the cis-but-2-ene (2b) and cis-pent-2-ene complexes (2e) were determined by single-crystal X-ray analyses. The Pt atom is square-planar, coordinated by three chloro ligands and the olefin ligand [interplanar angle PtCl3/Pt(C=C) 89.6° (2b), 86.6° (2e)]. The C=C bond length is 1.38(1)Å (2b) and 1.32(2)Å (2e), respectively. The substituents (Me, Et) on the double bond are bent back away from Pt [torsion angles C–C=C–Pt 109(2)–110.7(8)°]. There are close contacts of the [K(18- cr-6)]+ cations both to the trans and one of the cis chloro ligands [K···Cl 3.208(2)/3.230(2)Å (2b); 3.220(5)/3.370(5)Å (2e)].  相似文献   

19.
Steady-state and time-resolved fluorescence techniques were used to study (R)- and (S)-[1,1′-binaphthalene]-2,2′-diol (1,1′-binaphthol or BINOL) dilute solutions of different polarity solvents, as well as their inclusion complexes with α- and βcyclodextrins (CDs) in water. BINOLs in dilute water solutions exhibited a surprisingly high fluorescence anisotropy that was explained as being due to the formation of fairly large order π–π stacking aggregates in aqueous polar media. Stoichiometries, formation constants and the changes of enthalpy and entropy upon inclusion were also obtained by measuring the variation of the fluorescence intensity with [CD] and temperature. Results agree with the formation of 1:1 stoichiometry complexes, but the association constants are rather low and very similar for both enantiomers. Molecular mechanic calculations in the presence of water were employed to study the formation of BINOL complexes with both α- and βCDs. For the most stable structures of any of the complexes only a small portion of the guests, in agreement with thermodynamics parameters and quenching experiments, penetrates inside the CD cavities. Driving forces for 1:1 inclusion processes may be dominated by non-bonded van der Waals host:guest interactions. The low guest:host binding constants and poor enantioselectivity of α- and βCDs for BINOLS may be a consequence of the BINOL aggregation in water.  相似文献   

20.
The C–HX (X=N, O, S) intramolecular hydrogen bond between the α-hydrogen of the vinyl group and the corresponding heteroatom in the series of 1-vinyl-2-(2′-heteroaryl)pyrroles was examined by ab initio calculations at the B3LYP/6-311(d,p) level. It was shown that the C–HN hydrogen bond is stronger than the C–HO hydrogen bond and the latter is, in turn, stronger than the C–HS hydrogen bond. This conclusion is supported by calculations of 1H NMR chemical shieldings.  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