Journal articles on the topic 'Anion-π'

To see the other types of publications on this topic, follow the link: Anion-π.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Anion-π.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Li, Lei, Yu-Jian Hong, Dong-Yang Chen, Wang-Chuan Xiao, and Mei-Jin Lin. "Anion–π interactions in lithium–organic redox flow batteries." Chemical Communications 55, no. 16 (2019): 2364–67. http://dx.doi.org/10.1039/c8cc09834d.

Full text
Abstract:
The interactions between the electrolyte anions and electron-deficient redox-active organic molecules (anion–π interactions) have strong influences on the battery properties due to the anion–π-induced formation of radical anions or sandwich-like aggregates.
APA, Harvard, Vancouver, ISO, and other styles
2

Mirocki, Artur, and Artur Sikorski. "Structural Characterization of Multicomponent Crystals Formed from Diclofenac and Acridines." Materials 15, no. 4 (February 17, 2022): 1518. http://dx.doi.org/10.3390/ma15041518.

Full text
Abstract:
Multicomponent crystals containing diclofenac and acridine (1) and diclofenac and 6,9-diamino-2-ethoxyacridine (2) were synthesized and structurally characterized. The single-crystal XRD measurements showed that compound 1 crystallizes in the triclinic P-1 space group as a salt cocrystal with one acridinium cation, one diclofenac anion, and one diclofenac molecule in the asymmetric unit, whereas compound 2 crystallizes in the triclinic P-1 space group as an ethanol solvate monohydrate salt with one 6,9-diamino-2-ethoxyacridinium cation, one diclofenac anion, one ethanol molecule, and one water molecule in the asymmetric unit. In the crystals of the title compounds, diclofenac and acridines ions and solvent molecules interact via N–H⋯O, O–H⋯O, and C–H⋯O hydrogen bonds, as well as C–H⋯π and π–π interactions, and form heterotetramer bis[⋯cation⋯anion⋯] (1) or heterohexamer bis[⋯cation⋯ethanol⋯anion⋯] (2). Moreover, in the crystal of compound 1, acridine cations and diclofenac anions interact via N–H⋯O hydrogen bond, C–H⋯π and π–π interactions to produce blocks, while diclofenac molecules interact via C–Cl⋯π interactions to form columns. In the crystal of compound 2, the ethacridine cations interact via C–H⋯π and π–π interactions building blocks, while diclofenac anions interact via π–π interactions to form columns.
APA, Harvard, Vancouver, ISO, and other styles
3

Martínez-Camarena, Álvaro, Matteo Savastano, Carla Bazzicalupi, Antonio Bianchi, and Enrique García-España. "Stabilisation of Exotic Tribromide (Br3−) Anions via Supramolecular Interaction with a Tosylated Macrocyclic Pyridinophane. A Serendipitous Case." Molecules 25, no. 14 (July 10, 2020): 3155. http://dx.doi.org/10.3390/molecules25143155.

Full text
Abstract:
Tetraaza-macrocyclic pyridinophane L-Ts, decorated with a p-toluenesulfonyl (tosyl; Ts) group, appear to be a useful tool to provide evidence on how the interplay of various supramolecular forces can help stabilise exotic anionic species such as tribromide (Br3−) anions. Indeed, crystals of (H2L-Ts)(Br3)1.5(NO3)0.5 unexpectedly grew from an acidic (HNO3) aqueous solution of L-Ts in the presence of Br− anions. The crystal structure of this compound was determined by single crystal XRD analysis. Hydrogen bonds, salt-bridges, anion-π, π-π stacking, and van der Waals interactions contribute to stabilising the crystal lattice. The observation of two independent Br3− anions stuck over the π-electron densities of pyridine and tosyl ligand groups, one of them being sandwiched between two pyridine rings, corroborates the significance of anion-π interactions for N-containing heterocycles. We show herein the possibility of detecting anion-π contacts from fingerprint plots generated by Hirshfeld surface analysis, demonstrating the effective usage of this structural investigation technique to further dissect individual contributions of stabilising supramolecular forces.
APA, Harvard, Vancouver, ISO, and other styles
4

Kim, Hee-Joon. "Assembly of Sn(IV)-Porphyrin Cation Exhibiting Supramolecular Interactions of Anion···Anion and Anion···π Systems." Molbank 2022, no. 4 (September 25, 2022): M1454. http://dx.doi.org/10.3390/m1454.

Full text
Abstract:
Trans-diaqua[meso-tetrakis(4-pyridyl)porphyrinato]Sn(IV) dinitrate complexes were assembled in a two-dimensional manner via hydrogen bonding between aqua ligands and pyridyl substituents. Interestingly, this supramolecular assembly was accompanied by unconventional noncovalent interactions, such as anion···anion and anion···π interactions, which were confirmed by X-ray crystallographic analysis. Two nitrate anions close to 2.070 Å were constrained in a confined space surrounded by four hydrogen-bonded Sn(IV)-porphyrin cations. The nitrate anion was also 3.433 Å away from the adjacent pyrrole ring, and the dihedral angle between the two mean planes was estimated to be 7.39°. The preference of the anion···π interaction was related to the electron-deficient π-system owing to the high-valent Sn(IV) center and cationic nature of the porphyrin complex. These two unconventional noncovalent interactions played an important role in the formation of a one-dimensional array with pairs of Sn(IV)-porphyrin cation and nitrate anion.
APA, Harvard, Vancouver, ISO, and other styles
5

Panteleieva, Olha S., Vira V. Ponomarova, Alexander V. Shtemenko, and Kostiantyn V. Domasevitch. "Supramolecular networks supported by the anion...π linkage of Keggin-type heteropolyoxotungstates." Acta Crystallographica Section C Structural Chemistry 76, no. 8 (July 21, 2020): 753–62. http://dx.doi.org/10.1107/s205322962000950x.

Full text
Abstract:
Anion...π interactions are newly recognized weak supramolecular forces which are relevant to many types of electron-deficient aromatic substrates. Being less competitive with respect to conventional hydrogen bonding, anion...π interactions are only rarely considered as a crystal-structure-defining factor. Their significance dramatically increases for polyoxometalate (POM) species, which offer extended oxide surfaces for maintaining dense aromatic/inorganic stacks. The structures of tetrakis(caffeinium) μ12-silicato-tetracosa-μ2-oxido-dodecaoxidododecatungsten trihydrate, (C8H11N4O2)4[SiW12O40]·3H2O, (1), and tris(theobrominium) μ12-phosphato-tetracosa-μ2-oxido-dodecaoxidododecatungsten ethanol sesquisolvate, (C7H9N4O2)3[PW12O40]·1.5C2H5OH, (2), support the utility of anion...π interactions as a special kind of supramolecular synthon controlling the structures of ionic lattices. Both caffeinium [(HCaf)+ in (1)] and theobrominium cations [(HTbr)+ in (2)] reveal double stacking patterns at both axial sides of the aromatic frameworks, leading to the generation of anion...π...anion bridges. The latter provide the rare face-to-face linkage of the anions. In (1), every square face of the metal–oxide cuboctahedra accepts the interaction and the above bridges yield flat square nets, i.e. {(HCaf+)2[SiW12O40]4−} n . Two additional cations afford single stacks only and they terminate the connectivity. Salt (2) retains a two-dimensional (2D) motif of square nets, with anion...π...anion bridges involving two of the three (HTbr)+ cations. The remaining cations complete a fivefold anion...π environment of [PW12O40]3−, acting as terminal groups. This single anion...π interaction is influenced by the specific pairing of (HTbr)+ cations by double amide-to-amide hydrogen bonding. Nevertheless, invariable 2D patterns in (1) and (2) suggest the dominant role of anion...π interactions as the structure-governing factor, which is applicable to the construction of noncovalent linkages involving Keggin-type oxometalates.
APA, Harvard, Vancouver, ISO, and other styles
6

Aramaki, Yoshitaka. "Anion-π Catalyst." Journal of Synthetic Organic Chemistry, Japan 75, no. 9 (2017): 965–66. http://dx.doi.org/10.5059/yukigoseikyokaishi.75.965.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Zhao, Yingjie, César Beuchat, Yuya Domoto, Jadwiga Gajewy, Adam Wilson, Jiri Mareda, Naomi Sakai, and Stefan Matile. "Anion−π Catalysis." Journal of the American Chemical Society 136, no. 5 (January 23, 2014): 2101–11. http://dx.doi.org/10.1021/ja412290r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Cotelle, Yoann, Vincent Lebrun, Naomi Sakai, Thomas R. Ward, and Stefan Matile. "Anion-π Enzymes." ACS Central Science 2, no. 6 (May 23, 2016): 388–93. http://dx.doi.org/10.1021/acscentsci.6b00097.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Schottel, Brandi L., Helen T. Chifotides, and Kim R. Dunbar. "Anion-π interactions." Chem. Soc. Rev. 37, no. 1 (2008): 68–83. http://dx.doi.org/10.1039/b614208g.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Giese, Michael, Markus Albrecht, Gergana Ivanova, Arto Valkonen, and Kari Rissanen. "Geometrically diverse anions in anion–π interactions." Supramolecular Chemistry 24, no. 1 (November 3, 2011): 48–55. http://dx.doi.org/10.1080/10610278.2011.622384.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Sapronov, Alexander A., Alexey S. Kubasov, Victor N. Khrustalev, Alexey A. Artemjev, Gleb M. Burkin, Evgeny A. Dukhnovsky, Alexander O. Chizhov, et al. "Se⋯π Chalcogen Bonding in 1,2,4-Selenodiazolium Tetraphenylborate Complexes." Symmetry 15, no. 1 (January 11, 2023): 212. http://dx.doi.org/10.3390/sym15010212.

Full text
Abstract:
The series of substituted 1,2,4-selenodiazolium tetraphenylborate complexes were synthesized via cyclization between 2-pyridylselenylchloride, followed by the anion metathesis, and fully characterized. The utilization of tetraphenylborate anion, a strong π-electron donor via its phenyl rings, promoted the formation of assemblies exhibiting selenium–π interactions. The chalcogen bonding (ChB) interactions involving the π-systems of the tetraphenylborate anion were studied using density functional theory (DFT) calculations, where “mutated” anions were used to estimate the strength of the Se···π chalcogen bonds. Moreover, molecular electrostatic potential (MEP) surfaces were used to investigate the electron-rich and poor regions of the ion pairs. The quantum theory of atoms-in-molecules (QTAIM) and the noncovalent interaction (NCI) plot methods based on the topology of the electron density were used and combined to characterize the ChBs. The investigation reported herein disclosed that the formation of symmetrical dimers can be broken by the introduction of a stronger π-acceptor and, consequently, forming stronger Se···π contacts with selenodiazolium cations.
APA, Harvard, Vancouver, ISO, and other styles
12

Frontera, Antonio, David Quiñonero, and Pere M. Deyà. "Cation-π and anion-π interactions." Wiley Interdisciplinary Reviews: Computational Molecular Science 1, no. 3 (April 11, 2011): 440–59. http://dx.doi.org/10.1002/wcms.14.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Zhang, Jian, Bin Zhou, Zhen-Rong Sun, and Xue-Bin Wang. "Photoelectron spectroscopy and theoretical studies of anion–π interactions: binding strength and anion specificity." Physical Chemistry Chemical Physics 17, no. 5 (2015): 3131–41. http://dx.doi.org/10.1039/c4cp04687k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Yamuna, Thammarse S., Jerry P. Jasinski, Manpreet Kaur, Brian J. Anderson, and H. S. Yathirajan. "5-(4-Fluorophenyl)-2H-pyrazol-1-ium 2,2,2-trifluoroacetate." Acta Crystallographica Section E Structure Reports Online 70, no. 4 (March 15, 2014): o429—o430. http://dx.doi.org/10.1107/s1600536814005200.

Full text
Abstract:
The title salt, C9H8FN2+·C2F3O2−, crystallizes with two independent cations (AandB) and two independent anions (CandD) in the asymmetric unit. In the cations, the dihedral angles between the benzene and pyrazolium rings are 23.7 (3)° in cationAand 1.8 (8)° in cationB. In the crystal, each anion links to the two cationsviaN—H...O hydrogen bonds, forming a U-shaped unit with anR44(14) ring motif. These U-shaped units stack along theaaxis and are linkedviaC—H...O and C—H...F hydrogen bonds, forming slabs lying parallel to (100). Within the slabs there are π–π interactions between the pyrazolium rings [inter-centroid distance = 3.6326 (15) Å] and between the benzene rings [inter-centroid distance = 3.7244 (16) Å]. In the anions, the F atoms of the trifluoromethyl groups are disordered over two sets of sites, with refined occupancy ratios of 0.58 (3):0.42, 0.540 (14):0.46 (14), and 0.55 (2):0.45 (2) for anionC, and 0.73 (5):0.27 (5), 0.63 (5):0.37 (5), and 0.57 (8):0.43 (8) for anionD.
APA, Harvard, Vancouver, ISO, and other styles
15

Quiñonero, David, Antonio Frontera, Carolina Garau, Pablo Ballester, Antoni Costa, and Pere M. Deyà. "Interplay Between Cation-π, Anion-π and π-π Interactions." ChemPhysChem 7, no. 12 (October 30, 2006): 2487–91. http://dx.doi.org/10.1002/cphc.200600343.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Zhang, Zhu-Yan, Zhao-Peng Deng, Li-Hua Huo, Shu-E. Zhang, Hui Zhao, and Shan Gao. "Role of Anion–π Interactions in the Supramolecular Assembly of Salts Containing Asymmetrical Bis(pyridyl) Cations." Australian Journal of Chemistry 67, no. 10 (2014): 1504. http://dx.doi.org/10.1071/ch13673.

Full text
Abstract:
Self-assembly of three flexible bis(pyridyl) molecules with different inorganic acids (HPF6, HClO4, and HNO3) leads to the formation of eight salts, which exhibit diverse architectures involving hydrogen bonding and anion–π interactions. The three types of inorganic anions in these salts formed anion–π interactions with HM+ and H2M2+ except for 2, in which the six F atoms were involved in hydrogen bonds. Anion–π interactions produced diverse motifs of one (anion)-to-one (cation) in 1, 3, 4, and 6, two (anion)-to-one (cation) in 5 and 7, and (4,4) layer in 8. Hydrogen bonds resulted in interesting supramolecular architectures, such as right- and left-handed helical chains in 3, 2-fold interpenetrating networks in 5, and 3-fold interpenetrating networks in 8. Structural analyses indicated that the conformations of the three flexible asymmetrical bis(pyridyl) molecules and the non-covalent bonding interactions, such as hydrogen bonds and anion···π interactions, play crucial roles in the final architectures of these salts. Thermogravimetric analyses indicated that the thermal stability of the eight salts decreased in the order of perchlorates, hexafluorophosphates, and nitrates. The emission intensity of the perchlorates is much stronger than that of the hexafluorophosphates, nitrates, and their corresponding organic molecules in the solid state at room temperature.
APA, Harvard, Vancouver, ISO, and other styles
17

Lu, Tongxiang, and Steven E. Wheeler. "Quantifying the Role of Anion−π Interactions in Anion−π Catalysis." Organic Letters 16, no. 12 (June 10, 2014): 3268–71. http://dx.doi.org/10.1021/ol501283u.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Garau, Carolina, David Quiñonero, Antonio Frontera, Pablo Ballester, Antoni Costa, and Pere M. Deyà. "Approximate Additivity of Anion−π Interactions: An Ab Initio Study on Anion−π, Anion−π2and Anion−π3Complexes." Journal of Physical Chemistry A 109, no. 41 (October 2005): 9341–45. http://dx.doi.org/10.1021/jp053380p.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Geronimo, Inacrist, N. Jiten Singh, and Kwang S. Kim. "Nature of anion-templated π+–π+ interactions." Physical Chemistry Chemical Physics 13, no. 25 (2011): 11841. http://dx.doi.org/10.1039/c1cp20348g.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Wang, Yun Jie, Hao Hong Li, Zhi Rong Chen, Chang Cang Huang, and Ji Bo Liu. "1-Propylquinolinium triiodidocuprate(I)." Acta Crystallographica Section E Structure Reports Online 63, no. 11 (October 17, 2007): m2736. http://dx.doi.org/10.1107/s1600536807049136.

Full text
Abstract:
In the title compound, (C12H14N)2[CuI3], the asymmetric unit contains two N-propylquinolinium cations which lie on opposite sides of the CuI3 2− anion. In the anion, Cu—I bond distances lie in the range 2.5161 (13)–2.5529 (12) Å. All of the atoms in the anion are essentially coplanar, with an r.m.s. deviation from the [CuI3]2− mean plane of 0.0001 Å. In the crystal structure, an extensive network of C—H...I hydrogen bonds links the cations and anions into an extended three-dimensional network, with the cations further aggregated through π–π stacking interactions [centroid–centroid distances 3.48 (7) and 3.43 (5) Å].
APA, Harvard, Vancouver, ISO, and other styles
21

Yeston, Jake. "Measuring anion-π interactions." Science 364, no. 6438 (April 25, 2019): 348.3–348. http://dx.doi.org/10.1126/science.364.6438.348-c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Liu, Le, and Stefan Matile. "Anion-π transaminase mimics." Supramolecular Chemistry 29, no. 10 (November 18, 2016): 702–6. http://dx.doi.org/10.1080/10610278.2016.1258118.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Pook, Niels-Patrick. "Supramolecular Architecture in a Ni(II) Complex with a Weakly Bonded N,N′-(1,4-phenylenedi- carbonyl)Diglycinate Counter-Anion: Crystal Structure Investigation and Hirshfeld Surface Analysis." Crystals 9, no. 12 (November 23, 2019): 615. http://dx.doi.org/10.3390/cryst9120615.

Full text
Abstract:
In this work, we describe the structural investigation of a Ni(II) complex, [Ni(C12H8N2)2(H2O)2]2·(C12H10N2O6)·(NO3)2·10H2O, with phenanthroline ligands, a deprotonated aromatic dicarboxylic acid, N,N′-(1,4-phenylenedicarbonyl)diglycine, and a nitrate as counter-anions, as well as water molecules. Noncovalent interactions, such as π–π stacking, lone-pair···π, and C–H···π between the phenanthrolines of the cationic complex, [Ni(C12H8N2)2(H2O)2]2+, and counter-anions are observed. Moreover, the solvated and noncoordinating counter-anion, N,N′-(1,4-phenylenedicarbonyl)diglycinate, is embedded in classical and nonclassical hydrogen-bonding interactions with water molecules and phenanthrolines. The two water molecules coordinated by the NiII atom and hydrogen bonded to the carboxylate of the N,N′-(1,4-phenylenedicarbonyl)diglycinate show attractive secondary electrostatic interactions, and a DD/AA hydrogen bonding pattern is formed. The noncovalent interactions of the cationic complex and the solvated N,N′-(1,4-phenylenedicarbonyl)diglycinate counter anion were explored with a Hirshfeld surface analysis, and related contributions to crystal cohesion were determined. The results of the N,N′-(1,4-phenylenedicarbonyl)diglycinate counter anion were compared to those of a solvated N,N′-(1,4-phenylenedicarbonyl)diglycine molecule of a previously described copper(II) complex.
APA, Harvard, Vancouver, ISO, and other styles
24

Kuzniak-Glanowska, Emilia, Dorota Glosz, Grzegorz Niedzielski, Jedrzej Kobylarczyk, Monika Srebro-Hooper, James G. M. Hooper, and Robert Podgajny. "Binding of anionic Pt(ii) complexes in a dedicated organic matrix: towards new binary crystalline composites." Dalton Transactions 50, no. 1 (2021): 170–85. http://dx.doi.org/10.1039/d0dt03535a.

Full text
Abstract:
Square-planar [PtL4]2− (L = CN, Cl, Br) anions are bound by π-acidic HAT(CN)6 in solution and in the solid state to provide the basis for the first epitaxially grown anion–π crystalline composites.
APA, Harvard, Vancouver, ISO, and other styles
25

Bornhof, Anna-Bea, Antonio Bauzá, Alexander Aster, Marion Pupier, Antonio Frontera, Eric Vauthey, Naomi Sakai, and Stefan Matile. "Synergistic Anion–(π)n–π Catalysis on π-Stacked Foldamers." Journal of the American Chemical Society 140, no. 14 (April 2, 2018): 4884–92. http://dx.doi.org/10.1021/jacs.8b00809.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Çevik, Sabri, Musa Sarı, Murat Sarı, and Tuncay Tunç. "Crystal structure of 4,4′-bipyridine-1,1'-diium naphthalene-2,6-disulfonate dihydrate." Acta Crystallographica Section E Structure Reports Online 70, no. 9 (August 9, 2014): o989—o990. http://dx.doi.org/10.1107/s160053681401784x.

Full text
Abstract:
The title hydrated molecular organic salt, C10H10N22+·C10H6O6S22−·2H2O, crystallized with half a bipyridinium cation, half a naphthalene-2,6-disulfonate anion and a water molecule in the asymmetric unit. The whole cation and anion are generated by inversion symmetry, the inversion centers being at the center of the bridging C—C bond of the cation, and at the center of the fused C—C bond of the naphthalene group of the anion. In the crystal, the anions and cations stack alternately along theaaxis with π–π interactions [inter-centroid distance = 3.491 (1) Å]. The anions are linkedviaO—H...O(sulfonate) hydrogen bonds involving two inversion-related water molecules, forming chains along [10-1]. These chains are bridged by bifurcated N—H...(O,O) hydrogen bonds, forming a three-dimensional framework structure. There are also C—H...O hydrogen bonds present, reinforcing the framework structure.
APA, Harvard, Vancouver, ISO, and other styles
27

Kim, Ajeong, Rashid Ali, Seok Ho Park, Yong-Hoon Kim, and Jung Su Park. "Probing and evaluating anion–π interaction in meso-dinitrophenyl functionalized calix[4]pyrrole isomers." Chemical Communications 52, no. 74 (2016): 11139–42. http://dx.doi.org/10.1039/c6cc04562f.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Savastano, Matteo, Celeste García-Gallarín, Claudia Giorgi, Paola Gratteri, Maria Dolores López de la Torre, Carla Bazzicalupi, Antonio Bianchi, and Manuel Melguizo. "Solid State and Solution Study on the Formation of Inorganic Anion Complexes with a Series of Tetrazine-Based Ligands." Molecules 24, no. 12 (June 16, 2019): 2247. http://dx.doi.org/10.3390/molecules24122247.

Full text
Abstract:
Four molecules (L1–L4) constituted by an s-tetrazine ring appended with two identical aliphatic chains of increasing length bearing terminal morpholine groups were studied as anion receptors in water. The basicity properties of these molecules were also investigated. Speciation of the anion complexes formed in solution and determination of their stability constants were performed by means of potentiometric (pH-metric) titrations, while further information was obtained by NMR and isothermal titration calorimetry (ITC) measurements. The crystal structures of two neutral ligands (L3, L4) and of their H2L3(ClO4)2∙2H2O, H2L4(ClO4)2∙2H2O, H2L3(PF6)2, and H2L3(PF6)2∙2H2O anion complexes were determined by single crystal X-ray diffraction. The formation of anion–π interactions is the leitmotiv of these complexes, both in solution and in the solid state, although hydrogen bonding and/or formation of salt-bridges can contribute to their stability. Evidence of the ability of these ligands to form anion–π interactions is given by the observation that even the neutral (not-protonated) molecules bind anions in water to form complexes of significant stability, including elusive OH− anions.
APA, Harvard, Vancouver, ISO, and other styles
29

Zhou, Chun-Shan, Ya-Mei Pei, and Xiang-Gao Meng. "2,2′,2′′,2′′′-(3,6-Dioxaoctane-1,8-diyldinitrilo)tetrabenzimidazolium tetrakis(perchlorate) dihydrate." Acta Crystallographica Section E Structure Reports Online 63, no. 11 (October 17, 2007): o4334—o4335. http://dx.doi.org/10.1107/s160053680704980x.

Full text
Abstract:
In the title crystal structure, C38H44N10O2 4+·4ClO4 −·2H2O, components are linked into a two-dimensional framework by a combination of N—H...O, C—H...O, O—H...O and N—H...N hydrogen bonds. In addition, weak π–π stacking interactions and anion–π noncovalent interactions between perchlorate anions and heteroaromatic imidazole rings [O...Cg = 3.328 (10) and 3.386 (11) Å; Cg is the centroid of an imidazole ring] consolidate the crystal structure.
APA, Harvard, Vancouver, ISO, and other styles
30

Ng, Chee Koon, Teck Lip Dexter Tam, Fengxia Wei, Xuefeng Lu, and Jishan Wu. "Anion–π and anion–π-radical interactions in bis(triphenylphosphonium)-naphthalene diimide salts." Organic Chemistry Frontiers 6, no. 1 (2019): 110–15. http://dx.doi.org/10.1039/c8qo01122b.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Li, Lei, Yu-Jian Hong, Yun Lin, Wang-Chuan Xiao, and Mei-Jin Lin. "An electron-deficient nanosized polycyclic aromatic hydrocarbon with enhanced anion–π interactions." Chemical Communications 54, no. 84 (2018): 11941–44. http://dx.doi.org/10.1039/c8cc06522e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Capó, Magdalena, Jordi Benet-Buchholz, and Pablo Ballester. "Anion−π−π Interactions in a Dinuclear M2L2Metallocycle." Inorganic Chemistry 47, no. 22 (November 17, 2008): 10190–92. http://dx.doi.org/10.1021/ic801282r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Frontera, Antonio, David Quinonero, and Pere M. Deya. "ChemInform Abstract: Cation-π and Anion-π Interactions." ChemInform 42, no. 47 (October 27, 2011): no. http://dx.doi.org/10.1002/chin.201147277.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Quiñonero, David, and Antonio Frontera. "Benzene, an Unexpected Binding Unit in Anion–π Recognition: The Critical Role of CH/π Interactions." Sci 4, no. 3 (August 22, 2022): 32. http://dx.doi.org/10.3390/sci4030032.

Full text
Abstract:
We report high-level ab initio calculations (CCSD(T)(full)/CBS//SCS-RI-MP2(full)/aug-cc-pwCVTZ) that demonstrate the importance of cooperativity effects when Anion–π and CH/π interactions are simultaneously established with benzene as the π-system. In fact, most of the complexes exhibit high cooperativity energies that range from 17% to 25.3% of the total interaction energy, which is indicative of the strong influence of the CH/π on the Anion–π interaction and vice versa. Moreover, the symmetry-adapted perturbation theory (SAPT) partition scheme was used to study the different energy contributions to the interaction energies and to investigate the physical nature of the interplay between both interactions. Furthermore, the Atoms in Molecules (AIM) theory and the Non-Covalent Interaction (NCI) approach were used to analyze the two interactions further. Finally, a few examples from the Protein Data Bank (PDB) are shown. All results stress that the concurrent formation of both interactions may play an important role in biological systems due to the ubiquity of CH bonds, phenyl rings, and anions in biomolecules.
APA, Harvard, Vancouver, ISO, and other styles
35

Rusanova, Julia A., Olesia V. Kozachuk, Valentyna V. Semenaka, and Viktoriya V. Dyakonenko. "Bis(2,9-dimethyl-1,10-phenanthroline)copper(I) pentacyanidonitrosoferrate(II)." Acta Crystallographica Section E Structure Reports Online 69, no. 12 (November 27, 2013): m684—m685. http://dx.doi.org/10.1107/s1600536813031760.

Full text
Abstract:
The asymmetric unit of the title complex [Cu(C14H12N2)2]2[Fe(CN)5(NO)], consists of a [Cu(dmp)2]+cation (dmp is 2,9-dimethyl-1,10-phenanthroline) and half an [Fe(CN)5(NO)]2−anion. The anion is disordered across an inversion center with the FeIIion slightly offset (ca0.205Å) from the inversion center in the direction of the disorderedtrans-coordinating CN/NO ligands. The anion has a distorted octahedral coordination geometry. The CuIion is coordinated by two phenanthroline ligands in a distorted tetrahedral geometry. The dihedral angle between the phenanthroline ligands is 77.16 (4) Å. In the crystal, the cations are connected to the anions by weak C—H...N hydrogen bonds. In addition, weak π–π stacking interactions are observed, with centroid–centroid distances in the range 3.512 (3)–3.859 (3) Å.
APA, Harvard, Vancouver, ISO, and other styles
36

Caballero, Antonio, Fabiola Zapata, Lidia González, Pedro Molina, Ibon Alkorta, and José Elguero. "Discovery of anion–π interactions in the recognition mechanism of inorganic anions by 1,2,3-triazolium rings." Chem. Commun. 50, no. 36 (2014): 4680–82. http://dx.doi.org/10.1039/c4cc00169a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Esrafili, Mehdi D., Fariba Mohammadian-Sabet, and Mohammad Mehdi Baneshi. "Cooperative and substitution effects in enhancing the strength of fluorine bonds by anion−π interactions." Canadian Journal of Chemistry 93, no. 11 (November 2015): 1169–75. http://dx.doi.org/10.1139/cjc-2015-0154.

Full text
Abstract:
In this work, the cooperative effects between anion−π and fluorine bond interactions are studied by ab initio calculations at the MP2/6-311++G** level. Cooperative effects are observed in complexes in which anion−π and fluorine bond interactions coexist. For each complex, the shortening of the binding distance in the fluorine bond is more prominent than that in the anion−π bond. Favorable cooperativity energies are found with values that range between –0.51 and –0.76 kcal/mol. The atoms in molecules and molecular electrostatic potential analyses are carried out for these complexes to understand the nature of anion−π and fluorine bond interactions and the origin of the cooperativity.
APA, Harvard, Vancouver, ISO, and other styles
38

Garau, Carolina, Antonio Frontera, David Quiñonero, Pablo Ballester, Antoni Costa, and Pere M. Deyà. "Cation-π vs anion-π interactions: a complete π-orbital analysis." Chemical Physics Letters 399, no. 1-3 (November 2004): 220–25. http://dx.doi.org/10.1016/j.cplett.2004.10.014.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Pook, Niels-Patrick, Arnold Adam, and Mimoza Gjikaj. "Crystal structure and Hirshfeld surface analysis of (μ-2-{4-[(carboxylatomethyl)carbamoyl]benzamido}acetato-κ2 O:O′)bis[bis(1,10-phenanthroline-κ2 N,N′)copper(II)] dinitrate N,N′-(1,4-phenylenedicarbonyl)diglycine monosolvate octahydrate." Acta Crystallographica Section E Crystallographic Communications 75, no. 5 (April 25, 2019): 667–74. http://dx.doi.org/10.1107/s2056989019005164.

Full text
Abstract:
The centrosymmetric binuclear complex cation of the title compound, [Cu2(C12H10N2O6)(C12H8N2)4](NO3)2·C12H12N2O6·8H2O, is composed of a CuII atom with a distorted trigonal–bipyramidal coordination environment defined by four N atoms from two bidentate 1,10-phenanthroline ligands and one oxygen atom from one-half of the monodentate N,N′-(1,4-phenylenedicarbonyl)diglycinate anion. The asymmetric unit is completed by one-half of the N,N′-(1,4-phenylenedicarbonyl)diglycine solvent molecule, which is located on a centre of inversion, by one nitrate counter-anion and four water molecules. In the crystal, the cationic complexes are linked via intermolecular π–π stacking and through lone-pair...π interactions involving the N,N′-(1,4-phenylenedicarbonyl)diglycinate anion and the phenanthroline ligands. The N,N′-(1,4-phenylenedicarbonyl)diglycine solvent molecule is involved in classical and non-classical hydrogen-bonding interactions, as well as π–π stacking interactions. The centroid-to-centroid distances between aromatic entities are in the range 3.5402 (5)–4.3673 (4) Å. The crystal structure is stabilized by further C—H...O contacts as well as by O—H...O and N—H...O hydrogen bonds between water molecules, the nitrate anions, the N,N′-(1,4-phenylenedicarbonyl)diglycinate ligands, N,N′-(1,4-phenylenedicarbonyl)diglycine solvent molecules and phenanthroline ligands, giving rise to a supramolecular framework. A Hirshfeld surface analysis was carried out to quantify these interactions.
APA, Harvard, Vancouver, ISO, and other styles
40

Mirocki, Artur, and Artur Sikorski. "The Influence of Solvent on the Crystal Packing of Ethacridinium Phthalate Solvates." Materials 13, no. 22 (November 10, 2020): 5073. http://dx.doi.org/10.3390/ma13225073.

Full text
Abstract:
The synthesis, structural characterization and influence of solvents on the crystal packing of solvated complexes of ethacridine with phthalic acid: 6,9-diamino-2-ethoxyacridinium phthalate methanol solvate (1), 6,9-diamino-2-ethoxyacridinium phthalate ethanol solvate (2), 6,9-diamino-2-ethoxyacridinium phthalate isobutanol solvate (3), and 6,9-diamino-2-ethoxyacridinium phthalate tert-butanol solvate monohydrate (4) are described in this article. Single-crystal XRD measurements revealed that the compounds 1–4 crystallized in the triclinic P-1 space group, and the 6,9-diamino-2-ethoxyacridinium cations, phthalic acid anions and solvent molecules interact via strong N–H···O, O–H···O, C–H···O hydrogen bonds, and C–H···π and π–π interactions to form different types of basic structural motifs, such as: heterotetramer bis[···cation···anion···] in compound 1 and 2, heterohexamer bis[···cation···alcohol···anion···] in compound 3, and heterohexamer bis[···cation···water···anion···] in compound 4. Presence of solvents molecule(s) in the crystals causes different supramolecular synthons to be obtained and thus has an influence on the crystal packing of the compounds analyzed.
APA, Harvard, Vancouver, ISO, and other styles
41

Moon, Dohyun, and Jong-Ha Choi. "Crystal structure oftrans-difluoridotetrakis(pyridine-κN)chromium(III) trichlorido(pyridine-κN)zincate monohydrate from synchrotron data." Acta Crystallographica Section E Structure Reports Online 70, no. 11 (October 4, 2014): 290–93. http://dx.doi.org/10.1107/s160053681402145x.

Full text
Abstract:
In the asymmetric unit of the title compound, [CrF2(C5H5N)4][ZnCl3(C5H5N)]·H2O, there are two independent complex cations, one trichlorido(pyridine-κN)zincate anion and one solvent water molecule. The cations lie on inversion centers. The CrIIIions are coordinated by four pyridine (py) N atoms in the equatorial plane and two F atoms in atransaxial arrangement, displaying a slightly distorted octahedral geometry. The Cr—N(py) bond lengths are in the range 2.0873 (14) to 2.0926 (17) Å while the Cr—F bond lengths are 1.8609 (10) and 1.8645 (10) Å. The [ZnCl3(C5H5N)]−anion has a distorted tetrahedral geometry. The Cl atoms of the anion were refined as disordered over two sets of sites in a 0.631 (9):0.369 (9) ratio. In the crystal, two anions and two water molecules are linkedviaO—H...Cl hydrogen bonds, forming centrosymmetric aggregates. In addition, weak C—H...Cl, C—H...π and π–π stacking interactions [centroid–centroid distances = 3.712 (2) and 3.780 (2)Å] link the components of the structure into a three-dimensional network.
APA, Harvard, Vancouver, ISO, and other styles
42

Setifi, Zouaoui, Konstantin V. Domasevitch, Fatima Setifi, Pavel Mach, Seik Weng Ng, Vaclav Petříček, and Michal Dušek. "Multiple anion...π interactions in tris(1,10-phenanthroline-κ2N,N′)iron(II) bis[1,1,3,3-tetracyano-2-(2-hydroxyethyl)propenide] monohydrate." Acta Crystallographica Section C Crystal Structure Communications 69, no. 11 (October 12, 2013): 1351–56. http://dx.doi.org/10.1107/s0108270113027108.

Full text
Abstract:
In the ionic structure of the title compound, [Fe(C12H8N2)3](C9H5N4O2)2·H2O, the octahedral tris-chelate [Fe(phen)3]2+dications [Fe—N = 1.9647 (14)–1.9769 (14) Å; phen is 1,10-phenathroline] afford one-dimensional chains by a series of slipped π–π stacking interactions [centroid-to-centroid distances = 3.792 (3) and 3.939 (3) Å]. The 1,1,3,3-tetracyano-2-(2-hydroxyethyl)propenide anions, denoted tcnoetOH−, reveal an appreciable delocalization of π-electron density, involving the central propenide [C—C = 1.383 (3)–1.401 (2) Å] fragment and four nitrile groups, and this is also supported by density functional theory (DFT) calculations at the B97D/6-311+G(2d,2p) level. Primary noncovalent inter-moiety interactions comprise conventional O—H...O(N) and weak C—H...O(N) hydrogen bonding [O...O(N) = 2.833 (2)–3.289 (5) Å and C...O(N) = 3.132 (2)–3.439 (2) Å]. The double anion...π interaction involving a nitrile group of tcnoetOH−and twocis-positioned pyridine rings (`π-pocket') of [Fe(phen)3]2+[N...centroid = 3.212 (2) and 3.418 (2) Å] suggest the relevance of anion...π stackings for charge-diffuse polycyanoanions and commonM-chelate species.
APA, Harvard, Vancouver, ISO, and other styles
43

Mouas, Nardjes, Hocine Merazig, Jean-Claude Daran, and Eric Manoury. "rac-{[2-(Diphenylthiophosphoryl)ferrocenyl]methyl}dimethylammonium diphenyldithiophosphinate." Acta Crystallographica Section E Structure Reports Online 68, no. 4 (March 7, 2012): m381—m382. http://dx.doi.org/10.1107/s1600536812009129.

Full text
Abstract:
2-(Diphenylthiophosphino)dimethylaminomethylferrocene is a key intermediate in the synthesis of various ferrocenyl ligands. During one such synthesis, the title compound, [Fe(C5H5)(C20H22NPS)](C12H10PS2), was isolated as a by-product. It is built up by association of (2-(diphenylphosphino)ferrocenyl)methyl)dimethylammonium cations and diphenylphosphino dithioate anions. N—H...S, C—H...S and C—H...π interactions link the anions and cations. Each anion–cation pair is linked two by two through C—H...π interactions, forming pseudo dimers.
APA, Harvard, Vancouver, ISO, and other styles
44

López-Andarias, Javier, Antonio Frontera, and Stefan Matile. "Anion−π Catalysis on Fullerenes." Journal of the American Chemical Society 139, no. 38 (September 13, 2017): 13296–99. http://dx.doi.org/10.1021/jacs.7b08113.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Giese, Michael, Markus Albrecht, and Kari Rissanen. "Anion−π Interactions with Fluoroarenes." Chemical Reviews 115, no. 16 (August 17, 2015): 8867–95. http://dx.doi.org/10.1021/acs.chemrev.5b00156.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Zhao, Yingjie, Yuya Domoto, Edvinas Orentas, César Beuchat, Daniel Emery, Jiri Mareda, Naomi Sakai, and Stefan Matile. "Catalysis with Anion-π Interactions." Angewandte Chemie 125, no. 38 (August 14, 2013): 10124–27. http://dx.doi.org/10.1002/ange.201305356.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Estarellas, Carolina, Antonio Frontera, David Quiñonero, and Pere M. Deyà. "Anionπ Interactions in Flavoproteins." Chemistry - An Asian Journal 6, no. 9 (June 29, 2011): 2316–18. http://dx.doi.org/10.1002/asia.201100285.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Zhao, Yingjie, Yuya Domoto, Edvinas Orentas, César Beuchat, Daniel Emery, Jiri Mareda, Naomi Sakai, and Stefan Matile. "Catalysis with Anion-π Interactions." Angewandte Chemie International Edition 52, no. 38 (August 15, 2013): 9940–43. http://dx.doi.org/10.1002/anie.201305356.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Khalil, Salim M. "MINDO-Forces Calculations of 1-Substituted Cyclopropyl Cations and Anions." Zeitschrift für Naturforschung A 43, no. 8-9 (September 1, 1988): 801–5. http://dx.doi.org/10.1515/zna-1988-8-914.

Full text
Abstract:
AbstractMINDO-Forces calculations are reported, after complete optimization of geometry, for 1-X-sub- stituted cyclopropyl cations and anions, where X is H, O-, OH, NH2, CH3, NO2, CN, F, CHO. All the substituents are stabilizing. It was found that the π-donating groups interact strongly with the cyclopropyl cations, whereas the π-withdrawing groups interact strongly with cyclopropyl anion depending on the electron demand.
APA, Harvard, Vancouver, ISO, and other styles
50

Anbarasan, Radhakrishnan, Palaniyasan Eniya, Jeyaperumal Kalyana Sundar, and Menberu Mengesha Woldemariam. "Crystal structure and Hirshfeld surface analysis of 4-bromoanilinium nitrate." Acta Crystallographica Section E Crystallographic Communications 76, no. 6 (May 29, 2020): 973–76. http://dx.doi.org/10.1107/s2056989020006945.

Full text
Abstract:
The title compound C4H7BrN+·NO3 − crystallizes in the monoclinic crystal system with space group P21/c. In the crystal, π-π stacking interactions and strong N—H...O and C—H...O hydrogen bonds link the cations and anions into layers parallel to the bc plane. The O...H/H...O interactions between the cation and anion are the major factor determining the crystal packing.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography