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 106 毫秒
1.
2.
In this work, the geometrical and electronic properties of the mono cationic ionic liquid 1‐hexyl‐3‐methylimidazolium halides ([C6(mim)]+_X?, X=Cl, Br and I) and dicationic ionic liquid 1,3‐bis[3‐methylimidazolium‐1‐yl]hexane halides ([C6(mim)2X2], X=Cl, Br and I) were studied using the density functional theory (DFT). The most stable conformer of these two types ionic liquids (IL) are determined and compared with each other. Results show that in the most stable conformers, in both monocationic ILs and dicationic ILs, the Cl? and Br? anions prefer to locate almost in the plane of the imidazolium ring whereas the I? anion prefers nearly vertical location respect to the imidazolium ring plan. Comparison of hydrogen bonding and ionic interactions in these two types of ionic liquids reveals that these ionic liquids can be formed hydrogen bond by Cl? and Br? anion. The calculated thermodynamic functions show that the interaction of cation — anion pair in the dicationic ionic liquids are more than monocationic ionic liquids and these interactions decrease with increasing the halide anion atomic weight.  相似文献   

3.
The cation of the title compound, C13H18N3S+·NO3, consists of two subunits, viz. a planar indole moiety and a nonplanar thiouronium moiety. An isolated intermolecular hydrogen bond connects the cation with the nitrate anion. The crystal packing is additionally characterized by short intermolecular contacts between parallel indole systems. A topological analysis of the electron density revealed C—S single bonds and partial double bonding in the N—C—N group.  相似文献   

4.
The equilibrium geometries and vibrational frequencies of the ionic liquid 1-butyl-3-methylimidazolium cation and the alanine anion [BMIM][Ala] are studied using density functional theory (DFT) at the B3PW91/6-311+G(d,p) leve1. The most stable structures of the anion, the cation, and the ion pairs are obtained and characterized, and the geometry parameters of the ion pairs confirm the presence of a hydrogen bonding interaction between the anion and the cation. Natural bond orbital (NBO) analysis is also performed to analyze the atomic charge distribution and charge transfer in the [BMIM]+ cation and [BMIM][Ala] ionic liquids. The results show that there are the electrostatic interaction and multiple hydrogen bond interactions between the cation and the anion of the ionic liquids, and the stability of the ground state of the ion pairs mostly results from the hydrogen bonding between the lone pairs of O atoms in the anion and H in the imidazole cycle of the cation. There are some changes in microstructures and the charge distribution during the formation of the ion pairs.  相似文献   

5.
Several recent studies of hydroxyl-functionalized ionic liquids (ILs) have shown that cation-cation interactions can be dominating these materials at the molecular level when the anion involved is weakly interacting. The hydrogen bonds between the like ions led to the formation of interesting chain-like, ring-like, or distinct dimeric (i. e. two ion pairs) supermolecular clusters. In the present work, vibrational spectroscopy (ATR-IR and Raman) and density functional theory (DFT) calculations of the hydroxyl-functionalized imidazolium ionic liquid C2OHmimCl indicate that anion-cation hydrogen bonding interactions are dominating, leading to the formation of distinct dimeric ion pair clusters. In this arrangement, the Cl anions function as a bridge between the cations by establishing bifurcated hydrogen bonds with the OH group of one cation and the C(2)-H of another cation. Cation–cation interactions, on the other hand, do not play a significant role in the observed clusters.  相似文献   

6.

The effect of Tf2N, as a common anion in aqueous and ionic liquid (IL) phases, on Cs extraction in the IL system was investigated using C2mimTf2N as an IL and DCH18C6 as an extractant. The “common anion (Tf2N) effect” operated via the movement of Cs+·Tf2N from the aqueous phase into the IL phase by extraction in the form of Cs+·extractant·Tf2N, without any traditional cation exchange. The extraction product was recovered as a precipitate using a small amount of IL, which facilitated precipitation via supersaturation.

  相似文献   

7.
The title compound, Rb2[Co(H2O)6](C8H5O4)4·4H2O, consists of nearly regular octahedral [Co(H2O)6]2+ cations with the CoII cations on the inversion centre (special position 2a), Rb+ cations, hydrogen phthalate (Hpht) anions and disordered water molecules. The Rb+ cation is surrounded by nine O atoms from Hpht anions and water molecules, with a strongly deformed pentagonal–bipyramidal geometry and one apex split into three positions. The crystal packing is governed by numerous hydrogen bonds involving all water molecules and Hpht anions. In this way, layers parallel to the ab plane are formed, with the aromatic rings of the Hpht anions esentially directed along the c axis. While Hpht anions form the outer part of the layers, disordered water molecules and Rb+ cations alternate with [Co(H2O)6]2+ cations in the inner parts. The only interactions between the layers are van der Waals forces between the atoms of the aromatic rings. A search of the Cambridge Structural Database for coordination modes and types of hydrogen‐bonding interaction of the Hpht anion showed that, when uncoordinated Hpht anions are present, compounds with intermolecular hydrogen bonds are more numerous than compounds with intramolecular hydrogen bonds. For coordinated Hpht anions, chelating and bridging anions are almost equally common, while monodentate anions are relatively scarce. The same coordination modes appear for Hpht anions with or without intramolecular hydrogen bonds, although intramolecular hydrogen bonds are less common.  相似文献   

8.
Both cation and anion in the title compound, C2H12BN2+·I3, lie on a crystallographic mirror plane and are bound in the lattice by N—H⋯I hydrogen bonds, forming layers. Methyl‐H–borane‐H dihydride [–C—H(δ+)⋯(δ)H—B–] inter­actions between mol­ecules crosslink adjacent layers, giving `sandwich' stacking along the a axis.  相似文献   

9.
Crystals of 1‐(diaminomethylene)thiouron‐1‐ium chloride, C2H7N4S+·Cl, 1‐(diaminomethylene)thiouron‐1‐ium bromide, C2H7N4S+·Br, and 1‐(diaminomethylene)thiouron‐1‐ium iodide, C2H7N4S+·I, are built up from the nonplanar 1‐(diaminomethylene)thiouron‐1‐ium cation and the respective halogenide anion. The conformation of the 1‐(diaminomethylene)thiouron‐1‐ium cation in each case is twisted. Both arms of the cation are planar and rotated in opposite directions around the C—N bonds involving the central N atom. The dihedral angles describing the twisted conformation are 22.9 (1), 15.2 (1) and 4.2 (1)° in the chloride, bromide and iodide salts, respectively. Ionic and extensive hydrogen‐bonding interactions join oppositely charged units into a supramolecular network. The aim of the investigation is to study the influence of the size of the ionic radii of the Cl, Br and I ions on the dimensionality of the hydrogen‐bonding network of the 1‐(diaminomethylene)thiouron‐1‐ium cation. The 1‐(diaminomethylene)thiouron‐1‐ium system should be of use in crystal engineering to form multidimensional networks.  相似文献   

10.
L‐Cysteine hydrogen fluoride, or bis(L‐cysteinium) difluoride–L‐cysteine–hydrogen fluoride (1/1/1), 2C3H8NO2S+·2F·C3H7NO2S·HF or L‐Cys+(L‐Cys...L‐Cys+)F(F...H—F), provides the first example of a structure with cations of the `triglycine sulfate' type, i.e.A+(A...A+) (where A and A+ are the zwitterionic and cationic states of an amino acid, respectively), without a doubly charged counter‐ion. The salt crystallizes in the monoclinic system with the space group P21. The dimeric (L‐Cys...L‐Cys+) cation and the dimeric (F...H—F) anion are formed via strong O—H...O or F—H...F hydrogen bonds, respectively, with very short O...O [2.4438 (19) Å] and F...F distances [2.2676 (17) Å]. The F...F distance is significantly shorter than in solid hydrogen fluoride. Additionally, there is another very short hydrogen bond, of O—H...F type, formed by a L‐cysteinium cation and a fluoride ion. The corresponding O...F distance of 2.3412 (19) Å seems to be the shortest among O—H...F and F—H...O hydrogen bonds known to date. The single‐crystal X‐ray diffraction study was complemented by IR spectroscopy. Of special interest was the spectral region of vibrations related to the above‐mentioned hydrogen bonds.  相似文献   

11.
The title compound is a salt, 2H5O2+·C10H6O6S22−, in which the anion lies across an inversion centre in the space group C2/c, while the cation contains a short but noncentred O—H...O hydrogen bond. The ionic components are linked by charge‐assisted O—H...O hydrogen bonds into a three‐dimensional framework structure.  相似文献   

12.
The title compound, C16H16N5+·Cl (nbbH+·Cl), displays N—H⋯N, N—H⋯Cl and π–π inter­actions in the crystal packing. The Cl anion is chelated by the nbbH+ cation via two N—H⋯Cl hydrogen bonds. Inter‐ion N—H⋯N and N—H⋯Cl hydrogen bonds link ions related by 21 screw axes into chains along the c axis. These chains are further linked by glide‐plane operations to generate a three‐dimensional network, which is additionally stabilized by inter­chain π–π inter­actions.  相似文献   

13.
The asymmetric unit of the optically resolved title salt, C8H12N+·C4H5O4S, contains a 1‐phenylethanaminium monocation and a thiomalate (3‐carboxy‐2‐sulfanylpropanoate) monoanion. The absolute configurations of the cation and the anion are determined to be S and R, respectively. In the crystal, cation–anion N—H...O hydrogen bonds, together with anion–anion O—H...O and S—H...O hydrogen bonds, construct a two‐dimensional supramolecular sheet parallel to the ab plane. The two‐dimensional sheet is linked with the upper and lower sheets through C—H...π interactions to stack along the c axis.  相似文献   

14.
Mixtures of 4‐carboxypyridinium perchlorate or 4‐carboxypyridinium tetrafluoroborate and 18‐crown‐6 (1,4,7,10,13,16‐hexaoxacyclooctadecane) in ethanol and water solution yielded the title supramolecular salts, C6H6NO2+·ClO4·C12H24O6·2H2O and C6H6NO2+·BF4·C12H24O6·2H2O. Based on their similar crystal symmetries, unit cells and supramolecular assemblies, the salts are essentially isostructural. The asymmetric unit in each structure includes one protonated isonicotinic acid cation and one crown ether molecule, which together give a [(C6H6NO2)(18‐crown‐6)]+ supramolecular cation. N—H...O hydrogen bonds between the protonated N atoms and a single O atom of each crown ether result in the 4‐carboxypyridinium cations `perching' on the 18‐crown‐6 molecules. Further hydrogen‐bonding interactions involving the supramolecular cation and both water molecules form a one‐dimensional zigzag chain that propagates along the crystallographic c direction. O—H...O or O—H...F hydrogen bonds between one of the water molecules and the anions fix the anion positions as pendant upon this chain, without further increasing the dimensionality of the supramolecular network.  相似文献   

15.
The title compound, Na+·C9H7N4O5S·2H2O, presents a Z configuration around the imine C=N bond and an E configuration around the C(O)NH2 group, stabilized by two intra­molecular hydrogen bonds. The packing is governed by ionic inter­actions between the Na+ cation and the surrounding O atoms. The ionic unit, Na+ and 2‐oxo‐3‐semicarbazono‐2,3‐dihydro‐1H‐indole‐5‐sulfonate, forms layers extending in the bc plane. The layers are connected by hydrogen bonds involving the water mol­ecules.  相似文献   

16.
In the three isomeric salts, all C6H7N2O+·C6HCl2O4, of chloranilic acid (2,5‐dichloro‐3,6‐dihydroxy‐1,4‐benzoquinone) with 2‐, 3‐ and 4‐carbamoylpyridine, namely, 2‐carbamoylpyridinium hydrogen chloranilate (systematic name: 2‐carbamoylpyridinium 2,5‐dichloro‐4‐hydroxy‐3,6‐dioxocyclohexa‐1,4‐dienolate), (I), 3‐carbamoylpyridinium hydrogen chloranilate, (II), and 4‐carbamoylpyridinium hydrogen chloranilate, (III), acid–base interactions involving H‐atom transfer are observed. The shortest interactions between the cation and the anion in (I) and (II) are pyridinium N—H...(O,O) bifurcated hydrogen bonds, which act as the primary intermolecular interaction in each crystal structure. In (III), an amide N—H...(O,O) bifurcated hydrogen bond, which is much weaker than the bifurcated hydrogen bonds in (I) and (II), connects the cation and the anion.  相似文献   

17.
A novel layered cobalt phosphate with a racemic 1,2-diaminopropane templating agent, [CH2(NH3)CH(NH3)-CH3][CoPO4]2, has been synthesized and structurally characterized by single crystal X-ray diffraction. The compound crystallizes in the monoclinic system, space group P21/c with a = 10.457(2), b = 7.849(1), c = 6.692(1)?Å, β = 103.30(1)°, V = 534.5(2)?Å3, Z = 2. The combination of cobalt phosphate anion layers and doubly protonated cationic 1,2-diaminopropane templates is mainly through ionic bonding. The crystal structure is stabilized by a network of hydrogen bonds between layers and templating agents.  相似文献   

18.
In the title compound, (5‐oxo‐3a,6a‐diphenyl­perhydro­imidazo[4,5‐d]imidazol‐2‐ylidene)oxonium hydrogen sulfate, C16H15N4O2+·HSO4, the asymmetric unit contains a hydrogen sulfate anion and a 3a,6a‐diphenyl­glycoluril oxonium cation. The hydrogen sulfate anion is joined to the oxonium cation via a strong O—H⋯O hydrogen bond (H⋯O = 1.69 Å). The crystal packing is mainly dominated by inter­actions involving the hydrogen sulfate anion. The diphenyl­glycoluril oxonium cations also self‐assemble through N—H⋯O hydrogen bonds, forming mol­ecular chains along the [001] vector. Four intra­molecular C—H⋯N hydrogen bonds are observed, having an S(5) motif.  相似文献   

19.
The paradigm of supramolecular chemistry relies on the delicate balance of noncovalent forces. Here we present a systematic approach for controlling the structural versatility of halide salts by the nature of hydrogen bonding interactions. We synthesized halide salts with hydroxy-functionalized pyridinium cations [HOCnPy]+ (n=2, 3, 4) and chloride, bromide and iodide anions, which are typically used as precursor material for synthesizing ionic liquids by anion metathesis reaction. The X-ray structures of these omnium halides show two types of hydrogen bonding: ‘intra-ionic’ H-bonds, wherein the anion interacts with the hydroxy group and the positively charged ring at the same cation, and ‘inter-ionic’ H-bonds, wherein the anion also interacts with the hydroxy group and the ring system but of different cations. We show that hydrogen bonding is controllable by the length of the hydroxyalkyl chain and the interaction strength of the anion. Some molten halide salts exhibit a third type of hydrogen bonding. IR spectra reveal elusive H-bonds between the OH groups of cations, showing interaction between ions of like charge. They are formed despite the repulsive interaction between the like-charged ions and compete with the favored cation-anion H-bonds. All types of H-bonding are analyzed by quantum chemical methods and the natural bond orbital approach, emphasizing the importance of charge transfer in these interactions. For simple omnium salts, we evidenced three distinct types of hydrogen bonds: Three in one!  相似文献   

20.
The title salt, C3H8NO2+·C2HO4, formed between l ‐cysteine and oxalic acid, was studied as part of a comparison of the structures and properties of pure amino acids and their cocrystals. The structure of the title salt is very different from that formed by oxalic acid and equivalent amounts of d ‐ and l ‐cysteine molecules. The asymmetric unit contains an l ‐cysteinium cation and a semioxalate anion. The oxalate anion is only singly deprotonated, in contrast with the double deprotonation in the crystal structure of bis(dl ‐cysteinium) oxalate. The oxalate anion is not planar. The conformation of the l ‐cysteinium cation differs from that of the neutral cysteine zwitterion in the monoclinic and orthorhombic polymorphs of l ‐cysteine, but is similar to that of the cysteinium cation in bis(dl ‐cysteinium) oxalate. The structure of the title salt can be described as a three‐dimensional framework formed by ions linked by strong O—H...O and N—H...O and weak S—H...O hydrogen bonds, with channels running along the crystallographic a axis containing the bulky –CH2SH side chains of the cysteinium cations. The cations are only linked through hydrogen bonds via semioxalate anions. There are no direct cation–cation interactions via N—H...O hydrogen bonds between the ammonium and carboxylate groups, or via weaker S—H...S or S—H...O hydrogen bonds.  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