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.
采用等温溶解平衡法研究了五元体系Na+, Mg2+//Cl-, SO42-, NO3-, H2O在298.16 K下氯化钠饱和平衡体系的溶解度, 获得了相应的投影干盐图、氯图和水图. 研究结果表明, 在298.16 K下氯化钠饱和时, 该五元体系投影干盐图由8个二盐共饱和的双变面、13条三盐共饱的单变线和6个四盐共饱的零变点构成, 存在两种复盐, 8个二盐共饱双变面分别对应于NaCl+NaNO3, NaCl+Na2SO4, NaCl+MgCl2·6H2O, NaCl+MgSO4·Na2SO4·4H2O, NaCl+Mg(NO3)2·6H2O, NaCl+NaNO3·Na2SO4·2H2O, NaCl+MgSO4·7H2O 和NaCl+MgSO4·(1—6)H2O. 讨论了该相图在新疆硝酸盐矿开发利用过程中的应用.  相似文献   

2.
The molecular interactions between sulfur ylide (SY) and HSX molecules (X = F, Cl and Br) were investigated using the MP2 method with the 6-311++G (2d, 2p) basis set. The SY (CH2=SH2) have two reactive sites: CH2 (denoted as C-interaction) and SH2 (S-interaction) that both could interact with three atoms of HSX molecules. The results show that S···C, X···C, and H···C interactions (C-interactions) is preference over the X···S, H···S, and H···X interactions. Quantum theories of atoms in molecules and natural bond orbitals methods have been applied to analyze the intermolecular interactions. Good correlations have been found between the interaction energies, the second-order perturbation energies E(2), and the charge transfer qCT in the studied systems.  相似文献   

3.
The reactions of tetraphenylstibium nitrate with nitric acid and of tetraphenylstibium acetate with acetic acid yield adducts Ph4SbONO2 · HNO3 (I) and Ph4SbOC(O)CH3 · CHH3COOH (II). According to X-ray diffraction data, the antimony atom in [Ph4Sb]+[O2N-O···H···O-NO2]? has a tetrahedral coordination. The CSbC bond angles and Sb-C bond lengths vary within 108.04(6)°–109.75(4)° and 2.096(1)–2.098(1) Å, respectively. The anion includes the intermolecular hydrogen bond O(1)–H(1)···O(1)″: the O(1)-H(1), H(1)···O(1)″, and O(1)···O(1)″ distances are 0.91(4), 1.56(4), and 2.460(2) Å, respectively; and the OHO angle is 169(5)°. The nitrate groups are usually planar. Complex II also contains the intermolecular hydrogen bond with the following parameters: O(3)-H(3), 0.92 Å; H(3)···O(2), 1.68 Å; and O(3)···O(2) 2.594 Å; the O(2)H(3)O(3) angle is 172.1°. This H-bond noticeable changes the coordination polyhedron of the antimony atom compared to that in tetraphenylstibium acetate.  相似文献   

4.
A.T. Bottini  L.L. Hilton  J. Plott 《Tetrahedron》1975,31(17):1997-2001
Generation of 1,2-cyclohexadiene in the presence of conjugated dienes leads to (2 + 2) and/or (2 + 4) cycloaddition products. Methods used to generate 1,2-cyclohexadiene were: (1) treatment of 6,6-dibromobicyclo[3.1.0]hexane with methyllithium in tetrahydrofuran-ether at 0° and 60°; (2) treatment of 1,6-dichlorocyclohexene with magnesium in tetrahydrofuran at 60°; and (3) treatment of 1-bromocyclohexene with t-BuOK in THF at 60° or dimethyl sulfoxide at 40°. Comparison of relative reactivities with various dienes and styrene in ether solvents at 60° confirmed that the same intermediate, uncomplexed 1,2-cyclohexadiene, was involved in these reactions. Relative reactivities at 0° and 60° were found to be: 2-methylfuran (0·12, 0·14); furan (0·17, 0·16); 2,4-hexadiene (0·17,—); cis-pentadiene (0·53, 0·53); 2,3-dimethylbutadiene (2·35, 1·9); 1,3-cyclohexadiene (1·85,—); styrene (2·35, 1·9); and 1,3-cyclopentadiene (47, 14).  相似文献   

5.
The reaction of the carbodiphosphorane Ph3P=C=PPh3 ( 1 ) with MeI in the presence of iodine gives the oxidation product (IC(PPh3)2)2I[I3]·(I2)2 ( 2 ). In the solid state dimeric units linked by indefinite ···I?···I2···I3?···I2···I?··· chains are found. An additional I···I contact between the cation and the I2 molecule is formed, amounting to 359.23(5) pm. 2 crystallizes in the monoclinic space group P2/c, with the unit cell dimensions a = 2053.9(1), b = 1011.4(1), c = 1889.8(1) pm; β = 105.21(1)° and Z = 4.  相似文献   

6.
The region of positive electrostatic potentials (σ-hole) has been found along the extension of the C–I bond in the iodine-ylide CH2IH, which suggests that the iodine-ylide could interact with nucleophiles to form weak, directional noncovalent interactions. MP2 calculations confirmed that the I···N σ-hole interaction exists in the CH2IH···NCX (X = H, F, Cl, Br, I) bimolecular complexes. The NCCl···CH2IH···NCX (X = H, F, Cl, Br, I) termolecular complexes were constructed to investigate the weakly bonded σ-hole interactions to be strengthened by Cl···C halogen bond. And then, the NCY···CH2IH···NCCl (Y = H, F, Cl, Br, I) termolecular complexes were designed to investigate the enhancing effects of the I···N σ-hole interaction on the Y···C halogen/hydrogen-bonded interactions. Accompany with the mutual enhancing processes of the σ-hole interactions and halogen/hydrogen-bonded interactions in the iodine-ylide containing termolecular complexes, both the I···N σ-hole interactions and Y···C halogen/hydrogen-bonded interactions become more polarizable.  相似文献   

7.
Two new zinc(II) and cadmium(II) complexes, [Zn(PDT)2(NCS)2] (1) and [Cd((PDT)2I1.6(H2O)0.4(OH)0.4] · 0.4H2O (2) (“PDT” is the abbreviation of 3-(2-pyridyl)-5, 6-diphenyl-1,2,4-triazine), have been synthesized and characterized by elemental analysis, IR, 1H NMR spectroscopy, and studied by X-ray crystallography. Zinc(II) in 1 is six coordinate ZnN6. 2 is a co-crystal with cadmium(II) being 60% six-coordinated with a CdN4I2 environment and 40% seven-coordinated with a CdN4O2I environment. The supramolecular features in these complexes are guided/controlled by weak directional intermolecular S ··· π, C–H ··· π, C–H ··· I, and π ··· π interactions.  相似文献   

8.
The complexes of XH2NH2···HNO(X = B, Al, Ga) are characterized as head to tail with hydrogen bonding interactions. The structural characteristics can be confirmed by atoms in molecules (AIM) analysis, which also provide comparisons of hydrogen bonds strengths. The calculated interaction energies at G2MP2 level show that stability of complexes decrease as BH2NH2···HNO > AlH2NH2···HNO > GaH2NH2···HNO. On the basis of the vibrational frequencies calculations, there are red‐shifts for ν(X1? H) and blue‐shifts for ν(N? H) in the complexes on dihydrogen bonding formations (X1? H···H? N). On hydrogen bonding formations (N? H···O), there are red‐shifts for ν(N? H) compared to the monomers. Natural bond orbital (NBO) analysis is used to discuss the reasons for the ν(X1? H) and ν(N? H) stretching vibrational shifts by hyperconjugation, electron density redistribution, and rehybridization. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2008  相似文献   

9.
Two new palladium complexes derived from the di(2-pyridinyl)methanone N-(2-pyridinyl)hydrazone (DPMNPH) ligand are reported. The compounds were characterized by elemental analysis, spectroscopic studies, and, for the DPMNPH ligand, single-crystal X-ray diffraction analysis. The DPMNPH ligand crystallized as orthorhombic with the space group P212121. The H1 atom is intramolecularly bonded to the pyridinic N4 with N1–H1 = 0.92(3) Å, H1···N4 = 1.87(2) Å, N1···N4 = 2.615(2) Å, and N1–H1···N4 = 137(2)°. Both complexes were excellent catalysts in the Heck reaction in the presence of base.  相似文献   

10.
The analogy of chloride–chloride contacts in compounds containing Fe–Cl1···Cl2–Fe synthons with well-studied organic C–Cl1···Cl2–C interactions has been investigated. The crystal structures of the two tetrahaloferrate(III) salts, [(2-iodopyridinium)2FeX4]X (X = Cl, Br) have been determined. Analysis of these two isomorphous structures and related published structures shows that the arrangement of Fe–Cl1···Cl2–Fe synthons is similar to that of C–Cl1···Cl2–C with the Fe–Cl1···Cl2 and Cl1···Cl2–Fe angles being ~150°. While inter-chlorine distances are less than the sum of van der Waals radii in C–Cl1···Cl2–C units, they are equal to, or longer, than the sum of van der Waals radii in the corresponding Fe–Cl1···Cl2–Fe contacts. This might indicate that the arrangement of Fe–Cl1···Cl2–Fe synthons occurs predominately to reduce repulsive forces rather than as a result of attractive forces. However, it is observed that the halide–halide distance in [(2-iodopyridinium)2FeBr4]Br is shorter than in the isostructural chloride species, which can be explained by the fact that bromine is softer than chlorine. Several intermolecular forces unite the cations and anions within the crystalline lattice of [(2-iodopyridinium)2FeX4]X including N–H···X?, C–I···X–Fe, N(π)···X–Fe, N(π)···I–C, and Fe–X1···X2–Fe contacts. The calculated electron density and electrostatic potential of the [FeX4]? anions and the organic iodopyridinium cations was used to describe the arrangement of these synthons and the hierarchy of the strengths of the respective contacts.  相似文献   

11.
Herein, we first address the question posed in the title by establishing the tautomerization trajectory via the double proton transfer of the adenine·guanine (A·G) DNA base mispair formed by the canonical tautomers of the A and G bases into the A*·G* DNA base mispair, involving mutagenic tautomers, with the use of the quantum‐mechanical calculations and quantum theory of atoms in molecules (QTAIM). It was detected that the A·G ? A*·G* tautomerization proceeds through the asynchronous concerted mechanism. It was revealed that the A·G base mispair is stabilized by the N6H···O6 (5.68) and N1H···N1 (6.51) hydrogen bonds (H‐bonds) and the N2H···HC2 dihydrogen bond (DH‐bond) (0.68 kcal·mol?1), whereas the A*·G* base mispair—by the O6H···N6 (10.88), N1H···N1 (7.01) and C2H···N2 H‐bonds (0.42 kcal·mol?1). The N2H···HC2 DH‐bond smoothly and without bifurcation transforms into the C2H···N2 H‐bond at the IRC = ?10.07 Bohr in the course of the A·G ? A*·G* tautomerization. Using the sweeps of the energies of the intermolecular H‐bonds, it was observed that the N6H···O6 H‐bond is anticooperative to the two others—N1H···N1 and N2H···HC2 in the A·G base mispair, while the latters are significantly cooperative, mutually strengthening each other. In opposite, all three O6H···N6, N1H···N1, and C2H···N2 H‐bonds are cooperative in the A*·G* base mispair. All in all, we established the dynamical instability of the А*·G* base mispair with a short lifetime (4.83·10?14 s), enabling it not to be deemed feasible source of the A* and G* mutagenic tautomers of the DNA bases. The small lifetime of the А*·G* base mispair is predetermined by the negative value of the Gibbs free energy for the A*·G* → A·G transition. Moreover, all of the six low‐frequency intermolecular vibrations cannot develop during this lifetime that additionally confirms the aforementioned results. Thus, the A*·G* base mispair cannot be considered as a source of the mutagenic tautomers of the DNA bases, as the A·G base mispair dissociates during DNA replication exceptionally into the A and G monomers in the canonical tautomeric form. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
Abstract

Spectroscopic and single crystal X-ray diffraction studies of coordination compounds of CoII, NiII, ZnII, and PdII with phenylsulfonyl imidazole and benzimidazole derivatives (2-mfsiz, 2-mfsbz) were performed. The relevance of non-covalent interactions on the stabilization of intra and intermolecular arrangements in the ligands and their coordination compounds was investigated. The imidazole 2-mfsiz ligand presents two enantiomeric conformers, where the ethylphenylsulfone moiety stabilizes intermolecular lone pair···π (S–O···π(phe)) and H···π contacts, while its tetrahedral coordination compounds, [M(2-mfsiz)2X2] (M2+?=?Co, Ni, Zn; X?=?Cl, Br) showed intramolecular lone pair···π interactions (S–O···π(iz)). On the other hand, compounds [Cu2(2-mfsiz)22-AcO)4] and trans-[Pd(2-mfsiz)2Cl2] do not present lone pair···π interactions due to the metal ion geometry (square base pyramidal or square planar), which leads to formation of π(iz)···π(phe) interactions. For the benzimidazole ligand 2-mfsbz, an intramolecular, H(phe)···π(bz) contact was observed, remaining in its tetrahedral and octahedral coordination compounds, [M(2-mfsbz)2X2] (M2+?=?Co, Zn; X?=?Cl, Br, NO3). This interaction limits the free rotation of the ethylphenylsulfone moiety for stabilization of an intermolecular lone pair···π interaction (S–O···π(iz)). The dimeric [Zn2(2-mfsiz)22-AcO)4] compound has a π(bz)···π(phe) contact. Theoretical calculations confirmed the non-covalent interactions in the ligands and their coordination compounds.  相似文献   

13.
Four new bridged silver(I) complexes, namely [Ag22‐teda)(μ2‐fbc)2] ( 1 ), [Ag22‐1,6‐dah)2](bpdc) · 4H2O ( 2 ), [Ag22‐2‐ap)(2‐ap)(bnb)] · 0.34H2O ( 3 ), [Ag22‐pyc)2(2‐apy)2] · 0.5H2O ( 4 ), have been synthesized and characterized by elemental analysis and crystallographic methods [fbc = 4‐fluorobenzoate, teda = triethylenediamine ( 1 ); bpdc = biphenyl‐4,4′‐dicarboxylate, 1,6‐dah = 1,6‐diaminohexane ( 2 ); bnb = 3,5‐binitrobenzoate, 2‐ap = 2‐aminopyrimidine ( 3 ); pyc = 3‐pyridinecarboxylate acid, 2‐apy = 2‐aminopyridine ( 4 )]. Complex 1 contains a 1D linear chain paralleling to the c‐axis, whereas in complex 2 silver(I) atoms were bridged by the 1,6‐dah ligand into a zigzag chain, further giving a 1D ribbon by weak Ag ··· Ag interactions. Complex 3 consists of a dinuclear silver(I) [Ag22‐2‐ap)(2‐ap)(bnb)] moiety and a lattice water molecule, forming a 3D network via a number of hydrogen‐bonding interactions such as N–H ··· O, N–H ··· N and C–H ··· O hydrogen bond and other weak interactions such Ag ··· Ag, Ag ··· N, N ··· O as well as O ··· O interaction. Similar to 3 , the asymmetric unit of 4 consists of one dinuclear silver(I) [Ag22‐pyc)2(2‐apy)2] moiety and half lattice water molecule, further generating a tetranuclear silver(I) {[Ag22‐pyc)2(2‐apy)2]2 · H2O} moiety. These moieties construct a 3D supramolecular network structure of 4 through N–H ··· O, O–H ··· O and C–H ··· O hydrogen bonds as well as other weak interactions such as Ag ··· O and N ··· O interactions.  相似文献   

14.
Semiempirical (AM1 and PM3) and density functional theory (DFT) calculations were performed on about 50 porphyrin isomers with 25 each of 1,2 (syn) and 1,3 (anti) tautomeric forms. The corresponding oxa-and thia-core-modified analogues were also computed. The variations of relative energies and stabilities of the core-modified analogues were compared with parent porphyrin1 and the corresponding oxa-and thia-analogues. The trends in relative energies are not significantly changed while going from parent system to oxa-and thia-core-modified porphyrins in case of bothsyn andanti tautomers. Isomers of types [2·2·0·0], [3·0·1·0], [3·1·0·0], and [4·0·0·0] are destabilized due to the absence of methine bridge, which results in angle strain for tetrapyrroles. Isomers having [2·1·1·0], [2·1·0·1], [2·0·2·0] and [2·2·0·0] connectivity, the Z isomers, are more stable compared to the correspondingE isomers in bothsyn andanti forms of parent and core-modified analogues.  相似文献   

15.
The enhancing effects of molecule X (X = PH2Cl, SHCl, ClCl) on S···S and Se···Se chalcogen–chalcogen bonds in the cyclic trimers SHCl···SHCl···X and SeHCl···SeHCl···X were investigated by calculations at the MP2/aug‐cc‐pVTZ level. When molecule X is added to the dimer SHCl···SHCl (SeHCl···SeHCl), cyclic trimers are formed. Compared with the dimer, all the cyclic trimers have shorter S···S (Se···Se) lengths, greater electron densities, negative three‐body interaction energies, and larger second‐order perturbation energies. These results indicate that the addition of molecule X strengthens the original S···S (Se···Se) bond. For the SHCl···SHCl···X cyclic trimers, the S···S bond is strongest in SHCl···SHCl···PH2Cl, weaker in SHCl···SHCl···SHCl, and weakest in SHCl···SHCl···ClCl. This same trend is observed for the Se···Se bond in SeHCl···SeHCl···X. This means that PH2Cl provides the greatest enhancement to the S···S (Se···Se) interaction.  相似文献   

16.
The three-component systems RbClMnCl2H2O, 2RbCl · CoCl2 · 2H2O2RbCl · CuCl2 · 2H2OH2O, 2RbCl · CoCl2 · 2H2O2RbCl · MnCl2 · 2H2OH2O have been studied at 25°C. In the 2RbCl · CoCl2 · 2H2O2RbCl · CuCl2 · 2H2OH2O system, a discontinuous series of mixed crystals is formed and in the 2RbCl · CoCl2 · 2H2O2RbCl · MnCl2 · 2H2OH2O system, a continuous series is present.The unit cell parameters of the 2RbCl · CoCl2 · 2H2O double salt were determined: a = 5.586(2) Å, b = 6.469(3) Å, c = 6.988(2) Å, α = 65.31(3)°, β = 87.69(3)°, γ = 84.65(4)°, volume 228.4 Å3, Z = 1.The results obtained and discussed in conjunction with the crystal structure data suggest that for 2MICl · MIICl2 · 2H2O type salts the triclinic structure is stable only when the large rubidium and cesium ions participate in combinations with non-Jahn-Teller metal(II) ions. In the cases of Jahn-Teller metal(II) ions or with potassium or ammonium ions a tetragonal structure is always stable.  相似文献   

17.
郭文生  徐赫男  郭放  佟健 《中国化学》2005,23(3):272-274
设计、合成了两种蝶形主体分子:2,5-二(三苯甲基)对苯二酚1,2,5-二(二苯甲基)对苯二酚2.1和2可与许多有机小分子形成配位包合物。用IR表征了主体分子1和2 的包结物, 用1H NMR测定了主客体分子的摩尔比:1•DMF (1:2),1•DMSO (1:2),1 •吡啶 (1:2),1•环戊酮 (2:3)和2•DMF 1:2),2•DMSO (1:2),2 •THF (1:1),2•苯甲醛(1:2),2•苯乙酮 (1:2),2•2,5-己二酮 (1:1),2 •N-甲基-2-吡咯烷酮 (1:3)。单晶X-射线衍射分析了包结物2·苯甲醛的晶体结构,在分子间氢键的相互作用下晶体得以稳定。  相似文献   

18.
The title complex {[Co(dimb)2(H2O)2]·(NO3)2·(H2O)2}n ( 1 ) (dimb = 1,3‐di(imidazol‐1‐ylmethyl)‐5‐methylbenzene) has been hydrothermally synthesized by the reaction of dimb with Co(NO3)2·6H2O in aqueous solution. The cobalt(II) atoms are linked by bridging dimb ligands to form 2D corrugated and wavy networks containing Co4(dimb)4 macrocyclic motifs. Two neighboring independent layers interlinked each other in a parallel fashion to construct three‐dimensional structure by O–H···O, N–H···O and C–H···O hydrogen bonds. Magnetic measurement shows the weak antiferromagnetic interaction with a one‐dimensional chain model in the range of 5–300 K, with J of –0.68 cm−1.  相似文献   

19.
The [2]pseudorotaxane of cucurbit[6]uril (Q6) with 1,6-bis(imidazol-1-yl)hexane dihydrobromide was synthesized and its crystal structure was described. The structure of [2]pseudorotaxane was mainly stablized by host–guest C–H···O interactions. Self-assembly of the [2]pseudorotaxane produces infinite one-dimensional chains with intermolecular N–H···O, C–H···O, and π···π interactions; thus, a linear non-covalent pseudopolyrotaxane is formed.  相似文献   

20.
t-Butyl methyl ether has been UV photolysed (λ = 185 nm) to a maximal conversion of less than 0·1%. A study of the products (quantum yields) has been made: methanol (0·405), t-butanol (0·20), isobutene (0·178), t-butyl neopentyl ether (0·142), t-butyl ethyl ether (0·134), 1,2-di-t-butoxyethane (0·097), methane (0·056), isobutane (0·046), isopropenyl methyl ether (0·030), hydrogen (0·020), neopentane (0·016), ethane (0·015), formaldehyde (0·012), 2-methoxy-2-methyl-4-t-butoxybutane (0·005), hexamethylethane (0·0048), 2-methoxy-2-methylbutane (0·0027), 2-methoxy-2-methyl-3-t-butoxypropane (0·002), isopropyl methyl ether (0·0015), formaldehyde t-butyl methyl acetal (0·001), formaldehyde di-t-butyl acetal (0·001), 2-methoxy-2-methyl-4,4-dimethylpentane (0-001), 2-methoxy-2-methyl-3,3-dimethylbutane (0·0003), 2,5-dimethoxy-2,5-dimethylhexane (0·0002), di-t-butyl ether (5 · 10?5), 2,2-dimethyloxirane (?, <- 0·001). There is no decomposition of the t-BuO radical into acetone (< 5 · 10?4) and CH3. Cyclisation reactions leading to α,α-dimethyloxetane (< 10?4) and 1-methoxy-1-methylcyclopropane (< 10?4) do not occur. The material balance yields C5H11·97O1·018.The main modes of fragmentation (ca 82%) are represented by the homolytic CO bond split, either into t-butyl and methoxy (ca 52%) or into t-butoxy and methyl (ca 30%), Fragmentation into methanol and isobutene (8·5%) as well as into formaldehyde and isobutane (2%) are further modes of decomposition. The break of a CC linkage (4·5%) mainly occurs by elimination of molecular methane. A CH bond split has a probability of ca 3% with the methoxy CH bond the more likely one to break.  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