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 62 毫秒
1.
The structure of ammonium hydrogensquarate squaric acid monohydrate has been determined by single crystal X-ray diffraction. The compound crystallizes in the monoclinic space group C2/c and exhibits a 3D network with molecules linked by intermolecular interactions with participation of the H2Sq, HSq?, NH4 +, and H2O species. The HSq? anion and the neutral H2Sq form a strong head-to-tail dimer through O–H···O hydrogen bonding with lengths of 2.587 and 2.494 Å (protected space between numeral and unit). The layers are connected by ammonium cations and water molecules in a plane through the O···N (2.950, 2.978, 3.036 Å) and O···O (2.953, 2.781 Å) bonds. Another such layer is connected to the NH4 + cation in the adjacent plane through bifurcated N–H···O hydrogen-bonding to form a double layer (NH···O bond lengths are 3.036, 2.978, 2.857, 2.909, 2.958, and 2.742 Å, respectively). The IR-band assignment of the compound was achieved using the polarized IR-spectroscopy of oriented colloids in a nematic host. Theoretical ab initio calculations were performed and achieved with a view to explain the IR-bands of the H2Sq.HSq? motif.  相似文献   

2.
The changes in stabilization energy upon the formation of intermolecular hydrogen, dihydrogen and lithium bond complexes between C2B3H7, LiH and HF have been investigated using MP2 method with aug-cc-pVDZ basis set. The interaction of HF with nido-C2B3H7 could occur through the formation of B–H···H–F, C–H···F–H and B–C···H–F classical and non-classical hydrogen bonds. The B–C bonds in backbone of the C2B3H7 as electron donor interact with σ* orbital of HF as electron acceptor. Also interaction of LiH with nido-C2B3H7 resulted in B–C···Li–H and B–H···LiH lithium bonds as well as C–H···H–Li dihydrogen bond complexes. In some of these complexes, LiH interacts with B–C bonds. Results are indicating that more stable complexes belong to interaction of HF and LiH with backbone of the nido-C2B3H7. The AIM and NBO methods were used to analyze the intermolecular interactions; also the electron density at the bond critical point and the charge transfer of obtained complexes were studied.  相似文献   

3.
We have performed theoretical studies on sixteen molecular cubes for both (NH3·HCl)(H2O)6 and (NH3·HF)(H2O)6. We use an empirical gauge, based upon the N?CH and H?CX bond lengths, to categorize the degree to which the cubes are neutral adduct or ion pair in character. On this basis, we describe all sixteen cubes of the former as highly ionized, but only five of the latter as greater than 85% ionic in character. Addition of one or two bridging water molecules to form (NH3·HF)(H2O)7 or (NH3·HF)(H2O)8 raises the percent ionic character to greater than 85% for these systems. The relative energy of the cubes can be categorized based on simple chemical principles. The computed vibrational frequency corresponding to the proton stretch in the N?CH?CF framework shows the highest degree of redshifting for systems near 50% ion-pair character. Molecular cubes close to neutral adduct or to ion-pair character show less redshifting of this vibrational motion.  相似文献   

4.
Tryptammonium hydrogentartarate (1) crystallizes in the orthorhombic space group P212121 and exhibits a crystal structure consisting of the tryptammonium cation, hydrogentartarate anion and a solvent methanol. The cations and anions are joined into a 3D network by intermolecular NH···OH(CH3) and NH3···O(tart) bonds with lengths of 2.998, 2.772, 2.902, and 2.847 Å, respectively. Hydrogentartarate anions are themselves connected by strong intermolecular O···H-O hydrogen bonds with lengths of 2.481 Å into infinite chains. The anions also participate in moderate hydrogen bonding with solvent methanol molecules of the (tart)O···O(CH3) type with bond lengths of 2.736 and 2.762 Å). The conformational preference of the tryptammonium cation is discussed by comparing the results with the data for other crystallographically determined structures of salts. Quantum chemical calculations at (density functional theory) DFT (B3LYP) level of theory and 6-311++G** basis set are performed for the interacting system tryptammonium cation/H2O with a view to predicting the geometry of the most stable conformer. The optical properties of tryptammonium hydrogentartarate have been elucidated in the solid-state by means of linear-polarized solid state IR-spectroscopy (IR-LD) of oriented solids as a colloid suspension in nematic hosts. Some limitations of the method are discussed as well.  相似文献   

5.
The five trimers of H2O···HNC···H2O, H2O···H2O···HNC, HNC···H2O···H2O, H2O···HNC···HNC, and HNC···HNC···H2O have been studied with quantum chemical calculations. Their structures, harmonic vibrational frequencies and interaction energies have been calculated at the B3LYP and MP2 levels with the aug‐cc‐pVDZ and aug‐cc‐pVTZ basis sets. The cooperative effect on these properties has also been studied quantitatively. For HNC:(H2O)2 systems, the cyclic H2O···H2O···HNC trimer is most stable with an interaction energy of ?16.01 kcal/mol and a large cooperative energy of ?3.25 kcal/mol at the MP2/aug‐cc‐pVTZ level. For H2O:(HNC)2 systems, the interaction energy and cooperative energy in the H2O···HNC···HNC trimer are larger than those in the HNC···HNC···H2O trimer. The NH stretch frequency has a blue shift for the terminal HNC molecule in the HNC···H2O···H2O and HNC···HNC···H2O trimers and a red shift in other cases. A many‐body analysis has also been performed to understand the interaction energies in these hydrogen‐bonded clusters. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

6.
Calcium peroxosilicates CaO·nSiO2·xH2yH2O2 (n = 1, 2) were synthesized by three methods: (1) the reaction of CaSiO3·nH2O (~7% CaSiO3 suspension in water) with 50.7, 73.4, and 92% H2O2 at 0–5°C produced compositions CaSiO3·6H2O·2H2O2, CaSiO3·2H2O·4H2O2, and CaSiO3·3H2O·8H2O2, respectively; (2) the reaction of CaSiO3·17H2O with 50.7 and 73.4% H2O2 at 0–5°C produced the solvate CaO·2SiO2·4H2O·0.15H2O2; and (3) the reaction of CaSiO3·3H2O with H2O2 vapor at 5°C in the absence of anhydrone produced the solvate CaSiO3·3.5H2O·0.5H2O2. The products were characterized by X-ray powder diffraction, thermogravimetry, and IR spectroscopy.  相似文献   

7.
Two new precious metal coordination complexes [Pd(HL)2] (1) and [PtL2] (2) (where H2L is 1-(carboxymethyl)-1H-pyrazole-3-carboxylic acid) were synthesized by one-pot in situ hydrolysis. The complexes are assembled into a 3D structure by hydrogen bonding interactions and intermolecular contacts. The structures have been established by single-crystal X-ray diffraction, and characterized by FT-IR and liquid state fluorescent spectroscopy. Hirshfeld surface analysis reveals that the O···H contacts outnumber the other contacts in both structures (49.4 % of the total interactions for 1 and 41.6 % for 2).  相似文献   

8.
Substances crystallizing under various conditions from the MVO3(MF, HF)H2O2H2O (M = NH4, K) systems have been characterized by elemental analysis, infrared and Raman spectra and X-ray powder patterns. Besides the known M2[VO(O2)2F] complexes, complexes of two new types have been obtained: M3[HV2O2(O2)3F4·2H2O and (NH4)3[V2O2(O2)4F]·nH2O (n≈2). Vibrational spectra of new complexes are consistent with the presence of dimeric anions containing V(μ2O2)V and VFV bridges, respectively.  相似文献   

9.
The intermolecular interactions existing at three different sites between phenylacetylene and LiX (X = OH, NH2, F, Cl, Br, CN, NC) have been investigated by means of second‐order Møller?Plesset perturbation theory (MP2) calculations and quantum theory of “atoms in molecules” (QTAIM) studies. At each site, the lithium‐bonding interactions with electron‐withdrawing groups (? F, ? Cl, ? Br, ? CN, ? NC) were found to be stronger than those with electron‐donating groups (? OH and ? NH2). Molecular graphs of C6H5C?CH···LiF and πC6H5C?CH···LiF show the same connectional positions, and the electron densities at the lithium bond critical points (BCPs) of the πC6H5C?CH···LiF complexes are distinctly higher than those of the σC6H5C?CH···LiF complexes, indicating that the intermolecular interactions in the C6H5C?CH···LiX complexes can be mainly attributed to the π‐type interaction. QTAIM studies have shown that these lithium‐bond interactions display the characteristics of “closed‐shell” noncovalent interactions, and the molecular formation density difference indicates that electron transfer plays an important role in the formation of the lithium bond. For each site, linear relationships have been found between the topological properties at the BCP (the electron density ρb, its Laplacian ?2ρb, and the eigenvalue λ3 of the Hessian matrix) and the lithium bond length d(Li‐bond). The shorter the lithium bond length d(Li‐bond), the larger ρb, and the stronger the π···Li bond. The shorter d(Li‐bond), the larger ?2ρb, and the greater the electrostatic character of the π···Li bond. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
Crystallochemical treatment of the Nieuwland reaction is carried out on the basis of structural data obtained for crystalline acetylene complexes of the formulas NH4Cu8Cl9·4C2H2·1/2HCu2Cl3·H2O (I), NH4Cu3Cl4·C2H2 (II), KCu8Cl9·4C2H2·1/2HCu2Cl3·H2O (III), KCu3Cl4·C2H2 (IV), (NH4)2Cu3Cl5·4/9H2O·(xC2H2) with x=0 (Va), 1/9 (Vb), and 4/9 (Vc) and divinylacetylene (DVA) copper chloride compounds 2CuCl·DVA (VI) and 3CuCl·DVA (VII). Because of the π-coordination of a copper atom, the C≡C bond of the acetylene molecule is activated, as indicated by its significant (up to 1.32 Å) stretch (complexes I and II). The zeolite-like structure of complexes Va-Vc, which form in a catalytic solution, is realized as an infinite {[Cu108Cl168(H2O)16]60?}n anion with discrete [Cl(NH4)6]5+ cations inside. In this structure, only 16 Cu(1) atoms have a trigonal-pyramidal environment with the oxygen atom of the crystallization water located in the vertex (dCu?O=2.79 Å). Under the liquid-phase conditions of the Nieuwland reaction, these copper atoms are active centers stimulating the reaction to the subsequent acetylene oligomerization due to the π-interaction with the C2H2 molecule. The mutual arrangement of the catalytically active Cu(1) atoms in structure Va serves as a matrix for the synthesis of DVA, as shown by the structure of the 2CuCl·DVA adduct.  相似文献   

11.
On the refluxing ofM(II) oxalate (M=Mn, Co, Ni, Cu, Zn or Cd) and 2-ethanolamine in chloroform, the following complexes were obtained: MnC2O4·HOCH2CH2NH2·H2O, CoC2O4·2HOCH2CH2NH2, Ni2(C2O4)2·5HOCH2CH2NH2·3H2O, Cu2(C2O4)2·5HOCH2CH2NH2, Zn2(C2O4)2·5HOCH2CH2NH2·2H2O and Cd2(C2O4)2·HOCH2CH2NH2·2H2O. Following the reaction ofM(II) oxalate with 2-ethanolamine in the presence of ethanolammonium oxalate, a compound with the empirical formula ZnC2O4·HOCH2CH2NH2·2H2O1 was isolated. The complexes were identified by using elemental analysis, X-ray powder diffraction patterns, IR spectra, and thermogravimetric and differential thermal analysis. The IR spectra and X-ray powder diffraction patterns showed that the complexes obtained were not isostructural. Their thermal decompositions, in the temperature interval between 20 and about 900°C, also take place in different ways, mainly through the formation of different amine complexes. The DTA curves exhibit a number of thermal effects.  相似文献   

12.
The H‐bonded complexes formed from interaction between NH2NO (NA) and H2O2 (HP) have been investigated by using B3LYP and MP2 methods with a wide range of basis sets. We found six H‐bonded complexes in which three of them have cyclic structure. Calculations carried out at various levels show that the seven‐membered cyclic structure with O···HO and O···HN hydrogen bonding interactions is the most stable complex. The large binding energy obtained for A1 complex probably results from a more linear arrangement of the O···H N and O H···OH‐bonds in the seven‐membered structure A1. The natural bond orbital (NBO) analysis and the Bader's quantum theory of atoms in molecules have been used to elucidate the interaction characteristics of the NA‐HP complexes. The NBO results reveal that the charge transfer energy corresponds to the H‐bond interactions for A1 complex is grater than other complexes. The electrostatic nature of H‐bond interactions is predicted from QTAIM analysis. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

13.
Acetamide and thioacetamide react with the superacid solutions HF/MF5 (M = As, Sb) under formation of the corresponding salts [H3CC(OH)NH2]+MF6 and [H3CC(SH)NH2]+MF6 (M = As, Sb), respectively. The reaction of DF/AsF5 with acetamide and thioacetamide lead to the corresponding deuterated salts [H3CC(OD)ND2]+AsF6 and [H3CC(SD)ND2]+AsF6, respectively. The salts are characterized by vibrational and NMR spectroscopy, and in the case of [H3CC(OH)NH2]+AsF6 and [H3CC(SH)NH2]+AsF6 also by single‐crystal X‐ray analyses. The [H3CC(OH)NH2]+AsF6( 1 ) salt crystallizes in the triclinic space group P$\bar{1}$ with two formula units per unit cell, and the [H3CC(SH)NH2]+AsF6( 2 ) salt crystallizes in the monoclinic space group P21/c with four formula units per unit cell. In both crystal structures three‐dimensional networks are observed which are formed by intra‐ and intermolecular N–H ··· F and O–H ··· F or S–H ··· F hydrogen bonds, respectively. For the vibrational analyses, quantum chemically calculated spectra of the cations [H3CC(OH)NH2 · 3HF]+ and [H3CC(SH)NH2 · 2HF]+ are considered.  相似文献   

14.
The shifts in ionization energies which occur when a molecule is incorporated as an asymmetric dimer or in an intermolecular complex are analyzed theoretically. MO ? SCF calculations with 4–31G basis sets were performed on closed- and open-shell states of (HF)2, H2O·HF, and their valence–hole ions, as well as on the heterodimers incorporating the higher homologues CH3F, CH3OH, and (CH3)2O. The analysis concerns the influence of electrostatic, polarization, and charge transfer effects associated with complexation on the initial molecular state of each monomer system, as well as monomer–dimer differences in the electronic relaxation mechanism considered as a final state effect in the ionization process. The calculated ionization energy shifts which agree well with the experimental data available for (CH3)2O·HF, show that the shifts are dominated by electrostatic effects, but some effects arising from differences in molecular size and electric polarizability of the monomers can be discerned.  相似文献   

15.
It has been demonstrated in several instances that the 0.001 a.u. (electrons per bohr3) isodensity mapped electrostatic surface potentials on the fluorines along the outermost extensions of the C? F covalent bonds in tetrafluoromethane (CF4) are entirely negative, they are thereby unable to engage in σhole bonding interactions with the negative sites on another molecules. In this study, we have attempted at resolving this controversy by performing various high‐level electronic structure calculations with Quadratic Configuration Integrals of Singles and Doubles QCISD(full), second‐order Møller–Plesset MP2(full), and 12 other Density Functional Theory (DFT) based functionals with and without dispersion corrections, all in conjunction with the 6–311++G(2d,2p) basis set. The results achieved with all the levels of theory utilized suggest that the fluorine's σholes in CF4 are positive regardless of the 0.001‐, 0.0015‐, and 0.002‐a.u. isodensity mapped electrostatic surfaces examined. Because of this specific quality, the fluorines in CF4 have displayed their capacities to form not only 1:1 clusters with the Lewis bases such as water (H2O), ammonia (NH3), formaldehyde (H2C?O), hydrogen fluoride (HF), and hydrogen cyanide (HCN), but also 1:2, 1:3, and 1:4 clusters with the latter three randomly chosen Lewis bases. Various topological and nontopological features obtained from applications of atoms in molecules, noncovalent interaction reduced‐density‐gradient and natural bond orbital analytical tools reveal that the N···F, O···F, and F···F long‐ranged interactions developed between the interacting monomers in H3N···FCF3, H2O···FCF3, and (Y? D)n=1–4···F4C (Y? D = H2C?O, HCN, and HF) are reminiscent of halogen bonding. The nonadditive cooperative and anticooperative energetic effects emerged on cluster formations are discussed in detail. © 2015 Wiley Periodicals, Inc.  相似文献   

16.
The thermal decomposition of copper(II) complexes with salicylaldehyde S-methylthiosemicarbazone of general formula Cu(HL)X·nH2O (X=Py+NO3, NCS, 0.5SO4) and [Cu(L)NH3]·H2O was investigated in air atmosphere in the interval from room temperature to 1000°C. Decomposition of the complexes occurred in several successive endothermic and exothermic processes, and the residue was in all cases CuO.  相似文献   

17.
For the treatment of hydrogen bonding in SINDO1, 2p orbitals are introduced on hydrogen. The optimization of the orbital exponent together with the generation of approximate formulas for the core attraction integrals is sufficient to obtain good geometries and binding energies in hydrogen bonded systems. The method is applied to the dimers (H2O)2, (NH3)2, (HF)2, (HCOOH)2, (HCN)2, (H2S)2, and (HCI)2, mixed dimers NH3 · H2O and H2O · HCN, and cyclic polymers (HF)n(n = 3, 4, 6). © 1993 John Wiley & Sons, Inc.  相似文献   

18.
In this work, the time-dependent density functional theory (TD-DFT) method was used to study the electronic excited-state dynamics of the hydrogen-bonded p-Cresol–NH3–H2O complex. The intermolecular hydrogen bonds O1–H1···N and C–O1···H2 were demonstrated by the optimized geometric structure of the hydrogen-bonded p-Cresol–NH3–H2O complex. The infrared spectra (IR spectra) of the hydrogen-bonded p-Cresol–NH3–H2O complex in the ground and excited states were also calculated by using the density functional theory (DFT) and TD-DFT methods. It is demonstrated that hydrogen bond O1–H1···N can be strengthened while hydrogen bond C–O1···H2 is weakened upon photoexcitation to the S1 state. The significant changes of the hydrogen bond from the calculated bond lengths in different electronic states can be observed. In addition, the spectral shifts of the stretching vibrational mode of the hydrogen-bonded O–H group in different electronic states are accounted for the hydrogen bond changes in the S1 state too.  相似文献   

19.
Hydrates of 3-phenylpropenal thiosemicarbazone (HL·H2O) and semicarbazone (HL′·H2O) react in methanol with cobalt, nickel, copper, and zinc chlorides, nitrates, and acetates to form coordination compounds MX2·2HL·nSolv [M = Co, Ni, Cu, Zn; X = Cl, NO3; HL = C6H5CH=CH-CH=N-NHC(O)NH2; n = 0–3; Solv = H2O, CH3OH], CuX2·HL·nH2O [M = Ni, Cu; n = 0, 1], ML2·nH2O and ML′·nH2O [M = Co, Ni, Zn; HL′ = C6H5CH=CH-CH=N-NHC(O)NH2; n = 0–3]. In the presence of amines (A = C5H5N, 2-CH3C5H4N, 3-CH3C5H4N, and 4-CH3C5H4N) these reactions yield the complexes Cu(A)LCl·CH3OH and M(A)LX·nH2O [M = Cu, Ni; X = Cl, NO3; n = 0–2]. The copper complexes with the amine ligands are of polynuclear structure, and other complexes are monomeric. Carbazones (HL and HL′) are included in the complexes as bidentate N,S-and N,O-ligands. The thermolysis of the complexes involves the stages of removing solvent crystallization molecules (70–90°C), deaquation (150–170°C), and full thermal decomposition (500–580°C).  相似文献   

20.
Two new supramolecular complexes,[Cu(H_2dhbd)(3-pyOH)(H_2O)]_2·3-pyOH·2H_2O(1)and[Cu_2(dhbd)(dpa)_2-(H_2O)]·6H_2O(2)(H_4dhbd=2,3-dihydroxybutanedioic acid,3-pyOH=3-hydroxypyddine,dpa=2,2'-dipyridylamine),have been synthesized in aqueous solution and characterized by single-crystal X-ray diffraction,elemental analyses,UV-Vis and IR spectra,and TGA analysis.X-ray structural analysis revealed that,through four pairs of strong O…H—O hydrogen bonds,the cyclic dinuclear units in 1 together with four adjacent neighbors are connected into a 2Dhoneycomb network encapsulating free 3-pyOH ligands.Unexpectedly,the water-dimers are fixed in interlayers of2D honeycomb network and act as hydrogen-bond bridging to further extend these 2D networks into 3D hydro-gen-bonded framework.Complex 2 includes interesting 2D grids constructed from chiral dinuclear units throughstrong O…H—O and O…H—N hydrogen bonding,which are extended through other crystallization water mole-cules into three dimension with channels.Variable-temperature magnetic susceptibility measurements for bothcomplexes indicate the presence of weak antiferromagnetic exchange interactions between adjacent copper(Ⅱ)ions.  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