Journal articles on the topic 'H-Bonding Arrays'

To see the other types of publications on this topic, follow the link: H-Bonding Arrays.

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 'H-Bonding Arrays.'

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

Tadokoro, Makoto, Kyosuke Isoda, Yasuko Tanaka, Yuko Kaneko, Syoko Yamamoto, Tomoaki Sugaya, and Kazuhiro Nakasuji. "Self-Organization of -Crown Ether Derivatives into Double-Columnar Arrays Controlled by Supramolecular Isomers of Hydrogen-Bonded Anionic Biimidazolate Ni Complexes." Journal of Nanotechnology 2012 (2012): 1–10. http://dx.doi.org/10.1155/2012/216050.

Full text
Abstract:
Anionic tris (biimidazolate) nickelate (II) ([Ni(Hbim)3]−), which is a hydrogen-bonding (H-bonding) molecular building block, undergoes self-organization into honeycomb-sheet superstructures connected by complementary intermolecular H-bonds. The crystal obtained from the stacking of these sheets is assembled into channel frameworks, approximately 2 nm wide, that clathrate two cationic K+-crown ether derivatives organised into one-dimensional (1D) double-columnar arrays. In this study, we have shown that all five cationic guest-included crystals form nanochannel structures that clathrate the 1-D double-columnar arrays of one of the four types of K+-crown ether derivatives, one of which induces a polymorph. This is accomplished by adaptably fitting two types of anionic [Ni(Hbim)3]−host arrays. One is a network with H-bonded linkages alternating between the two different optical isomers of the and types with flexible H-bonded [Ni(Hbim)3]−. The other is a network of a racemate with 1-D H-bonded arrays of the same optical isomer for each type. Thus, [Ni(Hbim)3]−can assemble large cations such as K+crown-ether derivatives into double-columnar arrays by highly recognizing flexible H-bonding arrangements with two host networks of and .
APA, Harvard, Vancouver, ISO, and other styles
2

Ligthart, G. B. W. L., Dawei Guo, A. L. Spek, Huub Kooijman, Han Zuilhof, and Rint P. Sijbesma. "Ureidobenzotriazine Multiple H-Bonding Arrays: The Importance of Geometrical Details on the Stability of H-Bonds." Journal of Organic Chemistry 73, no. 1 (January 2008): 111–17. http://dx.doi.org/10.1021/jo7019338.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Piot, Luc, Carlos-Andres Palma, Anna Llanes-Pallas, Maurizio Prato, Zsolt Szekrényes, Katalin Kamarás, Davide Bonifazi, and Paolo Samorì. "Selective Formation of Bi-Component Arrays Through H-Bonding of Multivalent Molecular Modules." Advanced Functional Materials 19, no. 8 (April 23, 2009): 1207–14. http://dx.doi.org/10.1002/adfm.200801419.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Li, Jiang-Sheng, Li-Gong Chen, Ying-Ying Zhang, Yan-Jie Xu, Yi Deng, Tao Zeng, and Peng-Mian Huang. "Synthesis and Characterisation of a New Pyridinium Salt/Crown Ether Complex and its Self-Assembly." Journal of Chemical Research 2007, no. 6 (June 2007): 350–52. http://dx.doi.org/10.3184/030823407x225482.

Full text
Abstract:
The monoprotonated bipyridinium cation can complex with dibenzo-24-crown-8 by means of ion-dipole and charge-transfer interactions. They both thread to form a [2]pseudorotaxane-like complex for further intercomplex association to produce pseudopolyrotaxane arrays through N+–H···N hydrogen bonding in the solid state.
APA, Harvard, Vancouver, ISO, and other styles
5

Glidewell, Christopher, William T. A. Harrison, John N. Low, Jamie G. Sime, and James L. Wardell. "Patterns of soft C—H...O hydrogen bonding in diaryl sulfones." Acta Crystallographica Section B Structural Science 57, no. 2 (April 1, 2001): 190–200. http://dx.doi.org/10.1107/s0108768100015494.

Full text
Abstract:
In bis(4-tolyl) sulfone, C14H14O2S (1), 2,5,4′-trimethyldiphenyl sulfone, C15H16O2S (2), and 4-chlorodiphenyl sulfone, C12H9ClO2S (3), the molecules are linked by soft C—H...O hydrogen bonds into three different types of one-dimensional aggregate: simple chains in (1), molecular ladders in (2) and chains of fused rings in (3). In each of 3,4-dimethyl-4′-chlorodiphenyl sulfone, C14H13ClO2S (4), and 2,5-dimethyldiphenyl sulfone, C14H14O2S (5), the C—H...O hydrogen bonds link the molecules into two different types of two-dimensional sheet, based on a (4,4) net in (4) and a (3,6) net in (5). The patterns of soft C—H...O hydrogen bonds in (1)—(5) are compared with those in other diaryl sulfones, mainly retrieved from the Cambridge Structural Database, whose substitution patterns preclude the formation of hard hydrogen bonds. Observed aggregation modes range from the formation of no C—H...O hydrogen bonds at all, via finite (zero-dimensional) arrays through one-, two- and three-dimensional systems.
APA, Harvard, Vancouver, ISO, and other styles
6

Feng, Yang, Yui Hasegawa, Takeo Suga, Hiroyuki Nishide, Liuqing Yang, George Chen, and Shengtao Li. "Tuning Conformational H-Bonding Arrays in Aromatic/Alicyclic Polythiourea toward High Energy-Storable Dielectric Material." Macromolecules 52, no. 22 (November 11, 2019): 8781–87. http://dx.doi.org/10.1021/acs.macromol.9b01785.

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

Baljozovic, Milos, X. Liu, Mina Moradi, Igor A. Pasti, Jan Dreiser, Silvio Decurtins, Patrick Shahgaldian, Shi Xia Liu, and Thomas A. Jung. "Molecular Lego for Functional 2D Materials: Self-Assembly and Ordering of Bi-Molecular 2D Spinlattices of M(II,III) Phthalocyanines and of Highly Crystalline Free Standing Networks." ECS Meeting Abstracts MA2022-01, no. 14 (July 7, 2022): 944. http://dx.doi.org/10.1149/ma2022-0114944mtgabs.

Full text
Abstract:
Single layer low-dimensional materials are presently of emerging interest, also in the context of magnetism. We further developed on-surface supra-molecular architecturing to create two-dimensional spin arrays. By chemical programming of the modules, different checkerboards were produced containing metals of different oxidation and spin states, diamagnetic zinc, and a metal-free ‘spacer’. [1] In an in-depth, spectro-microscopy and theoretical account, we correlate the structure and the magnetic properties of these tunable systems and discuss the emergence of 2D Kondo magnetism [2] from the spin-bearing components and via the physico-chemical bonding to the underlying substrate. The contributions of the individual elements, as well as the role of the electronic surface state in the bottom substrate, are discussed. Also we show the the first free-standing single layer networks of calixarenes stabilized via functional groups enabling vdW, coordination and dipole-dipole, interaction. [3] Fig. 1) Graphical animation of the 2D spin arrays built by C--H - F--C bonding of two different magnetic atom bearing phthalocyanines that have been exploited to investigate the assembly and properties of Kondo lattices. [1] M. Baljozovic et al. Magnetochemistry 2021, 7, 119. [2] J. Girovsky et al. Nature Communications 2017 8:15388 [3] M. Moradi, N. Opara, L.G. Tulli, C. Wäckerlin, S.J. Dalgarno, S.J. Teat, M. Baljozovic, O. Popova, E. van Genderen, A. Kleibert, H. Stahlberg, J.P. Abrahams, C. Padeste, T.A. Jung, P. Shahgaldian, Sci. Adv. 2019; 5: eaav4489 Figure 1
APA, Harvard, Vancouver, ISO, and other styles
8

Langley, Philip J., Jürg Hulliger, Ram Thaimattam, and Gautam R. Desiraju. "Supramolecular synthons mediated by weak hydrogen bonding: forming linear molecular arrays via CC–H···NC and CC–H···O2N recognition." New Journal of Chemistry 22, no. 12 (1998): 1307–9. http://dx.doi.org/10.1039/a807552b.

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

Krische, Michael J., Jean-Marie Lehn, Nathalie Kyritsakas, Jean Fischer, Elina Karoliina Wegelius, and Kari Rissanen. "Self-assembly of 1- and 2-Dimensional Multicompartmental Arrays via the 2-Aminopyrimidine H-Bonding Motif and Selective Guest Inclusion." Tetrahedron 56, no. 36 (September 2000): 6701–6. http://dx.doi.org/10.1016/s0040-4020(00)00489-0.

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

Piot, Luc, Carlos-Andres Palma, Anna Llanes-Pallas, Maurizio Prato, Zsolt Szekrényes, Katalin Kamarás, Davide Bonifazi, and Paolo Samorì. "Supramolecular Architectures: Selective Formation of Bi-Component Arrays Through H-Bonding of Multivalent Molecular Modules (Adv. Funct. Mater. 8/2009)." Advanced Functional Materials 19, no. 8 (April 23, 2009): NA. http://dx.doi.org/10.1002/adfm.200990027.

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

Smith, Graham, and Urs D. Wermuth. "4-Amino-N-(4,6-dimethylpyrimidin-2-yl)benzenesulfonamide–4-nitrobenzoic acid (1/1)." Acta Crystallographica Section E Structure Reports Online 68, no. 6 (May 5, 2012): o1649—o1650. http://dx.doi.org/10.1107/s1600536812019563.

Full text
Abstract:
In the asymmetric unit of the title co-crystal, C7H5NO4·C12H14N4O2S, there are two independent but conformationally similar heterodimers, which are formed through intermolecular N—H...Ocarboxy and carboxyl–pyrimidine O—H...N hydrogen-bond pairs, giving a cyclic motif [graph set R 2 2(8)]. The dihedral angles between the rings in the sulfonamide molecules are 78.77 (8) and 82.33 (9)° while the dihedral angles between the ring and the CO2H group in the acids are 2.19 (9) and 7.02 (10)°. A two-dimensional structure parallel to the ab plane is generated from the heterodimer units through hydrogen-bonding associations between NH2 and sulfone groups. Between neighbouring two-dimensional arrays there are two types of aromatic π–π stacking interactions involving either one of the pyrimidine rings and a 4-nitrobenzoic acid molecule [minimum ring centroid separation = 3.5886 (9) Å] or two acid molecules [minimum ring centroid separation = 3.7236 (10) Å].
APA, Harvard, Vancouver, ISO, and other styles
12

Granifo, Juan, Beatriz Arévalo, Rubén Gaviño, Sebastián Suárez, and Ricardo Baggio. "Structural and theoretical characterization of a new twisted 4′-substituted terpyridine compound: 4′-(isoquinolin-4-yl)-2,2′:6′,2′′-terpyridine." Acta Crystallographica Section C Structural Chemistry 72, no. 12 (November 4, 2016): 932–38. http://dx.doi.org/10.1107/s2053229616016533.

Full text
Abstract:
4′-Substituted derivatives of 2,2′:6′,2′′-terpyridine with N-containing heteroaromatic substituents, such as pyridyl groups, might be able to coordinate metal centres through the extra N-donor atom, in addition to the chelating terpyridine N atoms. The incorporation of these peripheral N-donor sites would also allow for the diversification of the types of noncovalent interactions present, such as hydrogen bonding and π–π stacking. The title compound, C24H16N4, consists of a 2,2′:6′,2′′-terpyridine nucleus (tpy), with a pendant isoquinoline group (isq) bound at the central pyridine (py) ring. The tpy nucleus deviates slightly from planarity, with interplanar angles between the lateral and central py rings in the range 2.24 (7)–7.90 (7)°, while the isq group is rotated significantly [by 46.57 (6)°] out of this planar scheme, associated with a short Htpy...Hisqcontact of 2.32 Å. There are no strong noncovalent interactions in the structure, the main ones being of the π–π and C—H...π types, giving rise to columnar arrays along [001], further linked by C—H...N hydrogen bonds into a three-dimensional supramolecular structure. An Atoms In Molecules (AIM) analysis of the noncovalent interactions provided illuminating results, and while confirming the bonding character for all those interactions unquestionable from a geometrical point of view, it also provided answers for some cases where geometric parameters are not informative, in particular, the short Htpy...Hisqcontact of 2.32 Å to which AIM ascribed an attractive character.
APA, Harvard, Vancouver, ISO, and other styles
13

Gomes, Ligia R., Marcus V. N. de Souza, Cristiane F. Da Costa, James L. Wardell, and John Nicolson Low. "Crystal structures and Hirshfeld surfaces of four methoxybenzaldehyde oxime derivatives, 2-MeO-XC6H3C=NOH (X = H and 2-, 3- and 4-MeO): different conformations and hydrogen-bonding patterns." Acta Crystallographica Section E Crystallographic Communications 74, no. 11 (October 9, 2018): 1553–60. http://dx.doi.org/10.1107/s2056989018014020.

Full text
Abstract:
The crystal structures of four (E)-methoxybenzaldehyde oxime derivatives, namely (2-methoxybenzaldehyde oxime, 1, 2,3-dimethoxybenzaldehyde oxime, 2, 4-dimethoxybenzaldehyde oxime, 3, and 2,5-dimethoxybenzaldehyde oxime, 4, are discussed. The arrangements of the 2-methoxy group and the H atom of the oxime unit are s-cis in compounds 1–3, but in both independent molecules of compound 4, the arrangements are s-trans. There is also a difference in the conformation of the two molecules in 4, involving the orientations of the 2- and 5-methoxy groups. The primary intermolecular O—H(oxime)...O(hydroxy) hydrogen bonds generate C(3) chains in 1 and 2. In contrast, in compound 3, the O—H(oxime)...O(hydroxy) hydrogen bonds generate symmetric R 2 2(6) dimers. A more complex dimer is generated in 4 from the O—H(oxime)...O(hydroxy) and C—H(2-methoxy)...O(hydroxy) hydrogen bonds. In all cases, further interactions, C—H...O and C—H...π or π–π, generate three-dimensional arrays. Hirshfeld surface and fingerprint analyses are discussed.
APA, Harvard, Vancouver, ISO, and other styles
14

Castañeda-Calzoncit, Cesar E., Denis A. Cabrera-Munguia, Jesús A. Claudio-Rizo, Dora A. Solís-Casados, and Claudia M. López-Badillo. "Biocompatible Molybdenum Complexes Based on Terephthalic Acid and Derived from PET: Synthesis and Characterization." Asian Journal of Applied Science and Technology 06, no. 03 (2022): 25–34. http://dx.doi.org/10.38177/ajast.2022.6304.

Full text
Abstract:
Metal-organic molybdenum complexes were synthesized by the hydrothermal method using ammonium heptamolybdate as the metallic source, and as the organic ligand terephthalic acid (BDC) or bis(2-hydroxyethyl) terephthalate (BHET), obtained via glycolysis of poly(ethylene)terephthalate (PET). The BDC-Mo and BHET-Mo complexes were characterized by XRD, N2 physisorption, TGA, ATR-FTIR, SEM, XPS and their in vitro biocompatibility was tested by porcine fibroblasts viability. The results show that molybdates (MoO4-2) are coordinated to the carbonyl functional groups of BDC and BHET by urea bonding (-NH-CO-NH-) which is related to their high biocompatibility and high thermal stability. These organic molybdate complexes possess rectangular prism particles made up of rods arrays characteristics of molybdenum oxides (MoO3). The organic complexes BDC-Mo and BHET-Mo do not show to be cytotoxic for porcine dermal fibroblasts growing on their surface for up to 48 h of culture.
APA, Harvard, Vancouver, ISO, and other styles
15

Asfari, Zouhair, Eric J. Chan, Jack M. Harrowfield, Brian W. Skelton, Alexandre N. Sobolev, Pierre Thuéry, and Allan H. White. "Structural Systematics of Lanthanide(III) Picrate Solvates: Neutral, Mononuclear Ln(pic)3(dimethylsulfoxide)3 Arrays." Australian Journal of Chemistry 73, no. 6 (2020): 447. http://dx.doi.org/10.1071/ch19169.

Full text
Abstract:
Adducts of dimethylsulfoxide, dmso=Me2SO, with lanthanide(iii) picrates (picrate=2,4,6-trinitrophenoxide, pic) of stoichiometry Ln(pic)3·3dmso have been prepared and characterised by single-crystal X-ray structure determinations as discrete, neutral, mononuclear molecular species. Such complexes have been obtained across the gamut of Ln, specifically for Ln=La, Pr, Nd, Sm, Gd, Dy, Yb, Lu, and Y, presumably also accessible for other intermediate members, the series being isomorphous (monoclinic, C2/c, Z=8); a second triclinic P form has also been identified for Ln=La, Pr. In both forms, the metal atom coordination environments are nine-coordinate, tricapped trigonal prismatic, [Ln(dmso-O)3(pic-O,O′)3], two of the three unidentate ligands lying in one of the trigonal planes and one in the other (an isomer we have termed meridional, mer). A hydrated form of Ln(pic)3·2dmso·H2O stoichiometry has also been defined for Ln=Sm, Gd, Lu, the metal atom environment again nine-coordinate, [Ln(dmso-O)2(H2O)(pic-O,O′)3], but now fac, with the three unidentate ligands occupying one triangular face of the tricapped trigonal prism and involved in a centrosymmetric H-bonding array with the three similar ligands of an adjacent complex; the three capping atoms are nitro-oxygen atoms, the phenoxy-O triad occupying the other face.
APA, Harvard, Vancouver, ISO, and other styles
16

Hamann, Thomas, Dagmar Henschel, Ilona Lange, Oliver Moers, Armand Blaschette, and Peter G. Jones. "Polysulfonylamine, CLVII [1]. Molekulare Cokristalle aus Di(4-halogenbenzolsulfonyl)aminen und Sauerstoffbasen: Lamellare Schichten mit engen Zwischenschichtkontakten der Art C-H···Hal, Cl ··· Cl oder Br ··· Br / Molecular Co-crystals of Di(4-halobenzenesulfonyl)amines and Oxygen Bases: Lamellar Layers with Close C-H···Hal, Cl · · · Cl or Br · · · Br." Zeitschrift für Naturforschung B 57, no. 9 (September 1, 2002): 1051–65. http://dx.doi.org/10.1515/znb-2002-0912.

Full text
Abstract:
Di(4-fluorobenzenesulfonyl)amine (DFBSA), di(4-chlorobenzenesulfonyl)amine (DCBSA) or di(4-bromobenzenesulfonyl)amine (DBBSA) were co-crystallized with equimolar amounts of pyridine-N-oxide (PyO), dimethyl formamide (DMF) or 1,3-dimethylurea (DMU), respectively, to form the supramolecular complexes DFBSA·PyO (1; triclinic, space group P1 , Z = 1), DCBSA·DMF (2; monoclinic, P21/n, Z = 1) and DBBSA·DMU (3; triclinic, P1 , Z = 1). Throughout the triad, the molecules are ordered into lamellar layers parallel to the xy plane. Owing to the folded conformations of the disulfonylamines, the layers display an inner polar region of oxygen bases and (SO2)2NH groups, outer apolar regions of aromatic rings, and interlayer regions hosting the halogen atoms. These arrays mimic the formerly reported structures of a series of ionic metal di(arenesulfonyl)amides. The intralamellar connectivity is governed by conventional hydrogen bonding and weak C-H···O bonds, the former comprising in structure 1 a very strong N-H···O-N interaction [N···O 250,9(2) pm, N-H···O 171(3)°], in 2 an N-H···O=C bond, and in 3 a set of one S2N-H···O=C and two N-H(urea)···O=S bonds. A centrosymmetric (PyO)2 dimer is present in structure 1. The juxtapositions of adjacent layers reflect halogen-specific recognition patterns (1: three short C-H···F sequences, all F···F distances beyond the van der Waals limit dW; 2: one short C-H···Cl sequence and one close Cl ··· Cl contact < dW, all other Cl ··· Cl > dW; 3: four short C-H···Br sequences and one close Br···Br contact < dW, all other Br···Br > dW). The interhalogen contacts in 2 and 3 are of the type I, as characterized by θ(C-X···X')≈ θ(C'-X' ··· X); the four angles θ lie in the range 166-175°
APA, Harvard, Vancouver, ISO, and other styles
17

Pal, Shrinwantu, Arun K. Manna, and Swapan K. Pati. "The role of H bonding and dipole-dipole interactions on the electrical polarizations and charge mobilities in linear arrays of urea, thiourea, and their derivatives." Journal of Chemical Physics 129, no. 20 (November 28, 2008): 204301. http://dx.doi.org/10.1063/1.3020335.

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

Rodríguez-Ortega, Pilar Gema, Magdalena Sánchez-Valera, Juan Jesús López-González, and Manuel Montejo. "Fourier Transform Infrared Spectroscopy and Vibrational Circular Dichroism Assisted Elucidation of the Solution-State Supramolecular Speciation in Racemic and Enantiopure Ketoprofen." Applied Spectroscopy 76, no. 2 (January 8, 2022): 216–27. http://dx.doi.org/10.1177/00037028211060073.

Full text
Abstract:
The molecular structure and solution-state molecular interactions in the popular non-steroidal anti-inflammatory drug, ketoprofen, are extensively studied with the aim of gaining a better understanding of the chemical behavior of its solution state and its connection to its nucleation pathway and crystallization outcome. Using as reference solid-state X-ray structures of enantiomeric and racemic forms of ketoprofen, a set of self-assembly models underpinned by density functional theory calculations has been considered for the analysis of spectroscopic data, infrared (IR) and vibrational circular dichroism (VCD), obtained for solutions of the samples as a function of composition and solvent. From our results it can be concluded that, contrary to the general belief for generic carboxylic acids, there are no cyclic dimeric structures of ketoprofen present in solution, but rather linear arrays made up of two (in high polar or diluted media) or more units (in low polar or low dilution media). This observation is in line with the idea that the weak contacts (other than H-bonding) would hold the key to molecular self-assembly, in agreement with recent studies on other aromatic carboxylic acids.
APA, Harvard, Vancouver, ISO, and other styles
19

Smith, Graham. "Hydrogen-bonded two- and three-dimensional polymeric structures in the ammonium salts of 3,5-dinitrobenzoic acid, 4-nitrobenzoic acid and 2,4-dichlorobenzoic acid." Acta Crystallographica Section C Structural Chemistry 70, no. 3 (February 13, 2014): 315–19. http://dx.doi.org/10.1107/s2053229614002459.

Full text
Abstract:
The structures of ammonium 3,5-dinitrobenzoate, NH4+·C7H3N2O6−, (I), ammonium 4-nitrobenzoate dihydrate, NH4+·C7H4NO4−·2H2O, (II), and ammonium 2,4-dichlorobenzoate hemihydrate, NH4+·C7H3Cl2O2−·0.5H2O, (III), have been determined and their hydrogen-bonded structures are described. All three salts form hydrogen-bonded polymeric structures,viz.three-dimensional in (I) and two-dimensional in (II) and (III). With (I), a primary cation–anion cyclic association is formed [graph setR43(10)] through N—H...O hydrogen bonds, involving a carboxylate group with both O atoms contributing to the hydrogen bonds (denoted O,O′-carboxylate) on one side and a carboxylate group with one O atom involved in two hydrogen bonds (denoted O-carboxylate) on the other. Structure extension involves N—H...O hydrogen bonds to both carboxylate and nitro O-atom acceptors. With structure (II), the primary inter-species interactions and structure extension into layers lying parallel to (001) are through conjoined cyclic hydrogen-bonding motifs,viz.R43(10) (one cation, an O,O′-carboxylate group and two water molecules) and centrosymmetricR42(8) (two cations and two water molecules). The structure of (III) also has conjoinedR43(10) and centrosymmetricR42(8) motifs in the layered structure but these differ in that the first motif involves one cation, an O,O′-carboxylate group, an O-carboxylate group and one water molecule, and the second motif involves two cations and two O-carboxylate groups. The layers lie parallel to (100). The structures of salt hydrates (II) and (III), displaying two-dimensional layered arrays through conjoined hydrogen-bonded nets, provide further illustration of a previously indicated trend among ammonium salts of carboxylic acids, but the anhydrous three-dimensional structure of (I) is inconsistent with that trend.
APA, Harvard, Vancouver, ISO, and other styles
20

Brown, Malcolm. "Progress in understanding the biosynthesis and degradation of cellulose: The roles of light and Electron Microscopy." Proceedings, annual meeting, Electron Microscopy Society of America 47 (August 6, 1989): 770–71. http://dx.doi.org/10.1017/s0424820100155827.

Full text
Abstract:
Cellulose is the most abundant macromolecule on earth. It is a homopolymer of ß-1,4 linked glucose residues. The glucan chains have a degree of polymerization (=dp) ranging from several hundred to more than 20,000 glucose residues. The glucan chains typically associate with their neighbors to form crystalline rod-like structures known as microfibrils. Specific intermolecular associations of glucan chains (usually via H-bonding) result in several crystyalline polymorphs of cellulose such as Cellulose I (with parallel glucan chain orientation), or Cellulose II (absolute chain orientation not established). Cellulose is synthesized by many different organisms, including prokaryotic cells (Acetobacter, Rhizobium, Agrobacterium, Sarcina) and eukaryotic cells and organisms (many algae, fungi, mosses, ferns, vascular plants, Ascidians, and possibly humans).Electron microscopy has been very important in elucidating the site of cellulose synthesis. Using conventional ultrathin sections and specific cytochemical localization combined with negative staining of isolated scales, the site of cellulose synthesis was confirmed in the Golgi apparatus (Cellulose synthesis is localized in the Golgi apparatus of a few specialized algae such as Pleurochrysis. More commonly, cellulose synthesis occurs in association with the plasma membrane. Freeze fracture was instrumental in first visualizing terminal complexes (=TCs) associated with elongating microfibrils. TCs are presently known to exist either as rosettes (in all vascular plants so far examined, or in certain zygnematalean algae or linear arrays, largely confined to certain algae. The subunits of the TCs may be either transmembrane as in Valonia and Boergesenia or confined to one leaflet as in Oocystis or Vaucheria.
APA, Harvard, Vancouver, ISO, and other styles
21

Ponec, Robert, and Gleb Yuzhakov. "Evidence for 5-Center 4-Electron Bonding in (C···H···C···H···C) Array." Journal of Organic Chemistry 68, no. 21 (October 2003): 8284–86. http://dx.doi.org/10.1021/jo034676z.

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

Tantillo, Dean J., and Roald Hoffmann. "Prospecting for a 5-Center 4-Electron (C- - -H- - -C- - -H- - -C)+Bonding Array." Journal of the American Chemical Society 125, no. 14 (April 2003): 4042–43. http://dx.doi.org/10.1021/ja021394s.

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

Lynch, Daniel E., and Ian McClenaghan. "2-Amino-4-chloro-6-(4-carbamoylpiperidin-1-yl)pyrimidine hemihydrate." Acta Crystallographica Section E Structure Reports Online 57, no. 1 (December 14, 2000): o50—o51. http://dx.doi.org/10.1107/s1600536800019589.

Full text
Abstract:
The structure of the title compound, C10H15ClN5O·0.5H2O, comprises non-planar molecules that associateviaN—H...N and N—H...O interactions to form a three-dimensional hydrogen-bonded array. The water molecule resides on a twofold axis and is also involved in the hydrogen-bonding network.
APA, Harvard, Vancouver, ISO, and other styles
24

Kim, Nam Woon, Hyeonjeong Choe, Muhammad Ali Shah, Duck-Gyu Lee, and Shin Hur. "High-Density Patterned Array Bonding through Void-Free Divinyl Siloxane Bis-Benzocyclobutene Bonding Process." Polymers 13, no. 21 (October 21, 2021): 3633. http://dx.doi.org/10.3390/polym13213633.

Full text
Abstract:
Divinylsiloxane-bis-benzocyclobutene (DVS-BCB) has attracted significant attention as an intermediate bonding material, owing to its excellent properties. However, its applications are limited, due to damage to peripheral devices at high curing temperatures and unoptimized compressive pressure. Therefore, it is necessary to explore the compressive pressure condition for DVS-BCB bonding. This study demonstrates an optimization process for void-free DVS-BCB bonding. The process for obtaining void-free DVS-BCB bonding is a vacuum condition of 0.03 Torr, compressive pressure of 0.6 N/mm2, and curing temperature of 250 °C for 1 h. Herein, we define two factors affecting the DVS-BCB bonding quality through the DVS-BCB bonding mechanism. For strong DVS-BCB bonding, void-free and high-density chemical bonds are required. Therefore, we observed the DVS-BCB bonding under various compressive pressure conditions at a relatively low temperature (250 °C). The presence of voids and high-density crosslinking density was examined through near-infrared confocal laser microscopy and Fourier-transform infrared microscopy. We also evaluated the adhesion of the DVS-BCB bonding, using a universal testing machine. The results suggest that the good adhesion with no voids and high crosslinking density was obtained at the compressive pressure condition of 0.6 N/mm2. We believe that the proposed process will be of great significance for applications in semiconductor and device packaging technologies.
APA, Harvard, Vancouver, ISO, and other styles
25

Chen, Jing, Yong Lu, Wen-Shi Wu, Jin-Cao Dai, and Jian-Ming Lin. "trans-Diaquabis(2-carboxylato-4-nitropyridine 1-oxide-κ2 O 1,O 2)manganese(II) dihydrate." Acta Crystallographica Section E Structure Reports Online 62, no. 7 (June 14, 2006): m1540—m1541. http://dx.doi.org/10.1107/s1600536806021283.

Full text
Abstract:
In the centrosymmetric title complex, [Mn(C6H3N2O5)2(H2O)2]·2H2O, each MnII ion has a six-coordinate octahedral environment within an O6 donor set. The presence of O—H...O hydrogen-bonding interactions links adjacent molecules into a two-dimensional array.
APA, Harvard, Vancouver, ISO, and other styles
26

Yu, Yan-Hong, and Kun Qian. "1,3-Dihydroxy-2-(hydroxymethyl)propan-2-aminium 2,2-dichloroacetate." Acta Crystallographica Section E Structure Reports Online 65, no. 6 (May 14, 2009): o1278. http://dx.doi.org/10.1107/s1600536809016626.

Full text
Abstract:
The title compound, C4H12NO3+·C2HCl2O2−, was obtained from dichloroacetic acid and 2-amino-2-(hydroxymethyl)propane-1,3-diol. In the crystal structure, the cations and anions are connected by intermolecular N—H...O and O—H...O hydrogen bonding, forming a two-dimensional array parallel to (001). The crystal used for analysis was a merohedral twin, as indicated by the Flack parameter of 0.67 (6).
APA, Harvard, Vancouver, ISO, and other styles
27

Saravanan, Raju, Harkesh B. Singh, and Ray J. Butcher. "Bis(2-nitrophenyl) selenide, bis(2-aminophenyl) selenide and bis(2-aminophenyl) telluride: structural and theoretical analysis." Acta Crystallographica Section C Structural Chemistry 77, no. 6 (May 17, 2021): 271–80. http://dx.doi.org/10.1107/s2053229621005015.

Full text
Abstract:
Three organoselenium and organotellurium compounds containing ortho substitutents, namely, bis(2-nitrophenyl) selenide, C12H8N2O4Se, 2, bis(2-aminophenyl) selenide, C12H12N2Se, 3, and bis(2-aminophenyl) telluride, C12H12N2Te, 7, have been investigated by both structural and theoretical methods. In the structures of all three compounds, there are intramolecular contacts between both Se and Te with the ortho substituents. In the case of 2, this is achieved by rotation of the nitro group from the arene plane. For 3, both amino groups exhibit pyramidal geometry and are involved in intramolecular N—H...Se interactions, with one also participating in intermolecular N—H...N hydrogen bonding. While 3 and 7 are structurally similar, there are some significant differences. In addition to both intramolecular N—H...Te interactions and intermolecular N—H...N hydrogen bonding, 7 also exhibits intramolecular N—H...N hydrogen bonding. In the packing of these molecules, for 2, there are weak intermolecular C—H...O contacts and these, along with the O...N interactions mentioned above, link the molecules into a three-dimensional array. For 3, in addition to the N—H...N and N—H...Se interactions, there are also weak intermolecular C—H...Se interactions, which also link the molecules into a three-dimensional array. On the other hand, 7 shows intermolecular N—H...N interactions linking the molecules into R 2 2(16) centrosymmetric dimers. In the theoretical studies, for compound 2, AIM (atoms in molecules) analysis revealed critical points in the Se...O interactions with values of 0.017 and 0.026 a.u. These values are suggestive of weak interactions present between Se and O atoms. For 3 and 7, the molecular structures displayed intramolecular, as well as intermolecular, hydrogen-bond interactions of the N—H...N type. The strength of this hydrogen-bond interaction was calculated by AIM analysis. Here, the intermolecular (N—H...N) hydrogen bond is stronger than the intramolecular hydrogen bond. This was confirmed by the electron densities for 3 and 7 [ρ(r) = 0.015 and 0.011, respectively].
APA, Harvard, Vancouver, ISO, and other styles
28

Kennedy, Alan R., and Maurice O. Okoth. "2-Aminoethanaminium iodide." Acta Crystallographica Section E Structure Reports Online 68, no. 6 (May 16, 2012): o1731. http://dx.doi.org/10.1107/s160053681202065x.

Full text
Abstract:
The title salt, [NH3CH2CH2NH2]+·I−, has an array structure based on strong intermolecular N—H...N hydrogen bonding formed between the ammonium and amine groups of adjacent cations. This interaction gives a helical chain of cations that runs parallel to the b axis. The four remaining NH group H atoms all form hydrogen bonds to the iodide anion, and these iodide anions lie in channels parallel to the cation–cation chains.
APA, Harvard, Vancouver, ISO, and other styles
29

Li, Yu-Guang, Hai-Liang Zhu, and Edward R. T. Tiekink. "Tetrakis(diethylenetriammonium) tris[hexacyanoferrate(II)] octahydrate." Acta Crystallographica Section E Structure Reports Online 62, no. 4 (March 15, 2006): m760—m762. http://dx.doi.org/10.1107/s1600536806007525.

Full text
Abstract:
The asymmetric unit of the title complex, (C4H16N3)4[Fe(CN)6]3·8H2O, comprises two [H3N(CH2)2NH2(CH2)2NH3]3+ cations, 1.5 octahedral [Fe(CN)6]4− anions and four solvent water molecules; one anion is located on a center of inversion. Extensive hydrogen bonding of the types O—H...O, N and N—H...N, involving all components of the structure, leads to a three-dimensional array.
APA, Harvard, Vancouver, ISO, and other styles
30

Qiao, Liang, Xiao-Gang Chen, Ji-Xing Gao, and Yong Ai. "Three new quinuclidine-based structures: second harmonic generation response for 1,2-bis(1-azoniabicyclo[2.2.2]octan-3-ylidene)hydrazine dichloride." Acta Crystallographica Section C Structural Chemistry 75, no. 6 (May 21, 2019): 728–33. http://dx.doi.org/10.1107/s2053229619005898.

Full text
Abstract:
The crystal structures of three quinuclidine-based compounds, namely (1-azabicyclo[2.2.2]octan-3-ylidene)hydrazine monohydrate, C7H13N3·H2O (1), 1,2-bis(1-azabicyclo[2.2.2]octan-3-ylidene)hydrazine, C14H22N4 (2), and 1,2-bis(1-azoniabicyclo[2.2.2]octan-3-ylidene)hydrazine dichloride, C14H24N4 2+·2Cl− (3), are reported. In the crystal structure of 1, the quinuclidine-substituted hydrazine and water molecules are linked through N—H...O and O—H...N hydrogen bonds, forming a two-dimensional array. The compound crystallizes in the centrosymmetric space group P21/c. Compound 2 was refined in the space group Pccn and exhibits no hydrogen bonding. However, its hydrochloride form 3 crystallizes in the noncentrosymmetric space group Pc. It shows a three-dimensional network structure via intermolecular hydrogen bonding (N—H...C and N/C—H...Cl). Compound 3, with its acentric structure, shows strong second harmonic activity.
APA, Harvard, Vancouver, ISO, and other styles
31

Hong, Minghwei, Chao-Kai Cheng, Yen-Hsun Lin, Lawrence Boyu Young, Ren-Fong Cai, Chia-Hung Hsu, Chien-Ting Wu, and Jueinai Kwo. "Epitaxy from a Periodic Y–O Monolayer: Growth of Single-Crystal Hexagonal YAlO3 Perovskite." Nanomaterials 10, no. 8 (August 2, 2020): 1515. http://dx.doi.org/10.3390/nano10081515.

Full text
Abstract:
The role of an atomic-layer thick periodic Y–O array in inducing the epitaxial growth of single-crystal hexagonal YAlO3 perovskite (H-YAP) films was studied using high-angle annular dark-field and annular bright-field scanning transmission electron microscopy in conjunction with a spherical aberration-corrected probe and in situ reflection high-energy electron diffraction. We observed the Y–O array at the interface of amorphous atomic layer deposition (ALD) sub-nano-laminated (snl) Al2O3/Y2O3 multilayers and GaAs(111)A, with the first film deposition being three cycles of ALD-Y2O3. This thin array was a seed layer for growing the H-YAP from the ALD snl multilayers with 900 °C rapid thermal annealing (RTA). The annealed film only contained H-YAP with an excellent crystallinity and an atomically sharp interface with the substrate. The initial Y–O array became the bottom layer of H-YAP, bonding with Ga, the top layer of GaAs. Using a similar ALD snl multilayer, but with the first film deposition of three ALD-Al2O3 cycles, there was no observation of a periodic atomic array at the interface. RTA of the sample to 900 °C resulted in a non-uniform film, mixing amorphous regions and island-like H-YAP domains. The results indicate that the epitaxial H-YAP was induced from the atomic-layer thick periodic Y–O array, rather than from GaAs(111)A.
APA, Harvard, Vancouver, ISO, and other styles
32

Swinton Darious, Robert, Packianathan Thomas Muthiah, and Franc Perdih. "Supramolecular hydrogen-bonding patterns in salts of the antifolate drugs trimethoprim and pyrimethamine." Acta Crystallographica Section C Structural Chemistry 74, no. 4 (March 23, 2018): 487–503. http://dx.doi.org/10.1107/s2053229618004072.

Full text
Abstract:
Nine salts of the antifolate drugs trimethoprim and pyrimethamine, namely, trimethoprimium [or 2,4-diamino-5-(3,4,5-trimethoxybenzyl)pyrimidin-1-ium] 2,5-dichlorothiophene-3-carboxylate monohydrate (TMPDCTPC, 1:1), C14H19N4O3 +·C5HCl2O2S−, (I), trimethoprimium 3-bromothiophene-2-carboxylate monohydrate, (TMPBTPC, 1:1:1), C14H19N4O3 +·C5H2BrO2S−·H2O, (II), trimethoprimium 3-chlorothiophene-2-carboxylate monohydrate (TMPCTPC, 1:1:1), C14H19N4O3 +·C5H2ClO2S−·H2O, (III), trimethoprimium 5-methylthiophene-2-carboxylate monohydrate (TMPMTPC, 1:1:1), C14H19N4O3 +·C6H5O2S−·H2O, (IV), trimethoprimium anthracene-9-carboxylate sesquihydrate (TMPAC, 2:2:3), C14H19N4O3 +·C15H9O2 −·1.5H2O, (V), pyrimethaminium [or 2,4-diamino-5-(4-chlorophenyl)-6-ethylpyrimidin-1-ium] 2,5-dichlorothiophene-3-carboxylate (PMNDCTPC, 1:1), C12H14ClN4 +·C5HCl2O2S−, (VI), pyrimethaminium 5-bromothiophene-2-carboxylate (PMNBTPC, 1:1), C12H14ClN4 +·C5H2BrO2S−, (VII), pyrimethaminium anthracene-9-carboxylate ethanol monosolvate monohydrate (PMNAC, 1:1:1:1), C12H14ClN4 +·C15H9O2 −·C2H5OH·H2O, (VIII), and bis(pyrimethaminium) naphthalene-1,5-disulfonate (PMNNSA, 2:1), 2C12H14ClN4 +·C10H6O6S2 2−, (IX), have been prepared and characterized by single-crystal X-ray diffraction. In all the crystal structures, the pyrimidine N1 atom is protonated. In salts (I)–(III) and (VI)–(IX), the 2-aminopyrimidinium cation interacts with the corresponding anion via a pair of N—H...O hydrogen bonds, generating the robust R 2 2(8) supramolecular heterosynthon. In salt (IV), instead of forming the R 2 2(8) heterosynthon, the carboxylate group bridges two pyrimidinium cations via N—H...O hydrogen bonds. In salt (V), one of the carboxylate O atoms bridges the N1—H group and a 2-amino H atom of the pyrimidinium cation to form a smaller R 2 1(6) ring instead of the R 2 2(8) ring. In salt (IX), the sulfonate O atoms mimic the role of carboxylate O atoms in forming an R 2 2(8) ring motif. In salts (II)–(IX), the pyrimidinium cation forms base pairs via a pair of N—H...N hydrogen bonds, generating a ring motif [R 2 2(8) homosynthon]. Compounds (II) and (III) are isomorphous. The quadruple DDAA (D = hydrogen-bond donor and A = hydrogen-bond acceptor) array is observed in (I). In salts (II)–(IV) and (VI)–(IX), quadruple DADA arrays are present. In salts (VI) and (VII), both DADA and DDAA arrays co-exist. The crystal structures are further stabilized by π–π stacking interactions [in (I), (V) and (VII)–(IX)], C—H...π interactions [in (IV)–(V) and (VII)–(IX)], C—Br...π interactions [in (II)] and C—Cl...π interactions [in (I), (III) and (VI)]. Cl...O and Cl...Cl halogen-bond interactions are present in (I) and (VI), with distances and angles of 3.0020 (18) and 3.5159 (16) Å, and 165.56 (10) and 154.81 (11)°, respectively.
APA, Harvard, Vancouver, ISO, and other styles
33

Dai, Ji-Xiang, Fang-Hui Wu, Wen-Rui Yao, and Qian-Feng Zhang. "One-dimensional Hydrogen-bonded Chloride-Hydrate Assembly {[(H2O)4Cl2]2–}∞." Zeitschrift für Naturforschung B 62, no. 4 (April 1, 2007): 491–94. http://dx.doi.org/10.1515/znb-2007-0401.

Full text
Abstract:
A hydrogen-bonded chloride-hydrate assembly {[(H2O)4Cl2]2−}∞ has been ion-countered by the complex cations [Fe([9]aneS3)2]2+ ([9]aneS3 = 1,4,7-trithiacyclononane). In {[(H2O)4Cl2]2−}∞, four water molecules and two chloride ions are self-assembled to form a one-dimensional supramolecular array of O-H···O and O-H···Cl hydrogen bonding, which consists of fused fourand six-membered rings. The discrete cation [Fe([9]aneS3)2]2+ has a nearly regular octahedral FeS6 core with an average Fe-S bond length of 2.2586(5) Å .
APA, Harvard, Vancouver, ISO, and other styles
34

Chérif, Ichraf, Jawher Abdelhak, Mohamed Faouzi Zid, and Ahmed Driss. "2-Amino-5-chloropyridinium cis-diaquadioxalatochromate(III) sesquihydrate." Acta Crystallographica Section E Structure Reports Online 68, no. 6 (May 26, 2012): m824—m825. http://dx.doi.org/10.1107/s1600536812023392.

Full text
Abstract:
In the crystal structure of the title compound, (C5H6ClN2)[Cr(C2O4)2(H2O)2]·1.5H2O, the CrIII atom adopts a distorted octahedral geometry being coordinated by two O atoms of two cis water molecules and four O atoms from two chelating oxalate dianions. The cis-diaquadioxalatochromate(III) anions, 2-amino-5-chloropyridinium cations and uncoordinated water molecules are linked into a three-dimensional supramolecular array by O—H...O and N—H...O hydrogen-bonding interactions. One of the two independent lattice water molecules is situated on a twofold rotation axis.
APA, Harvard, Vancouver, ISO, and other styles
35

Chou, Chang-Chuan, Chia-Chi Yang, Hao-Ching Chang, Way-Zen Lee, and Ting-Shen Kuo. "Weaving an infinite 3-D supramolecular network via AuI⋯AuIII aurophilicity and C–H⋯Cl hydrogen bonding." New Journal of Chemistry 40, no. 3 (2016): 1944–47. http://dx.doi.org/10.1039/c5nj02860d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Zhao, Pu, Xian Wang, Fang Jian, Jun Zhang, and Lian Xiao. "Crystal engineered acid-base complexes with 2d and 3d hydrogen bonding systems using p-hydroxybenzoic acid as the building block." Journal of the Serbian Chemical Society 75, no. 4 (2010): 459–73. http://dx.doi.org/10.2298/jsc090416011z.

Full text
Abstract:
p-Hydroxybenzoic acid (p-HOBA) was selected as the building block for self-assembly with five bases, i.e., diethylamine, tert-butyl amine, cyclohexylamine, imidazole and piperazine, and generate the corresponding acid-base complexes 1-5. Crystal structure analyses suggest that proton-transfer from the carboxyl hydrogen to the nitrogen atom of the bases can be observed in 1-4; while only in 5 does a solvent water molecule co-exists with p-HOBA and piperazine. With the presence of O-H?O hydrogen bonds in 1-4, the deprotonated p-hydroxybenzoate anions (p-HOBAA-) are simply connected each other in a head-to-tail motif to form one-dimensional (1D) arrays, which are further extended to distinct two-dimensional (2D) (for 1 and 4) and three-dimensional (3D) (for 2 and 3 ) networks via N-H?O interactions. While in 5, neutral acid and base are combined pair wise by O-H?N and N-H?O bonds to form a 1D tape and then the 1D tapes are sequentially combined by water molecules to create a 3D network. Some interlayer or intralayer C-H?O, C-H?? and ??? interactions help to stabilize the supramolecular buildings. Melting point determination analyses indicate that the five acidbase complexes are not the ordinary superposition of the reactants and they are more stable than the original reactants.
APA, Harvard, Vancouver, ISO, and other styles
37

Hökelek, Tuncer, Gizem Sertkaya, Ezgi Ay, Safiye Özkaya, and Hacali Necefoğlu. "Crystal structure and Hirshfeld surface analysis of diaquabis(isonicotinamide-κN)bis(2,4,6-trimethylbenzoato-κO 1)nickel(II) dihydrate." Acta Crystallographica Section E Crystallographic Communications 73, no. 8 (July 21, 2017): 1237–41. http://dx.doi.org/10.1107/s205698901701060x.

Full text
Abstract:
In the title NiII complex, [Ni(C10H11O2)2(C6H6N2O)2(H2O)2]·2H2O, the divalent Ni ion occupies a crystallographically imposed centre of symmetry and is coordinated by two O atoms from the carboxylate groups of two 2,4,6-trimethylbenzoate (TMB) ligands [Ni—O = 2.0438 (12) Å], two N atoms from the pyridyl groups of two isonicotinamide (INA) ligands [Ni—N = 2.1506 (15) Å] and two water molecules [Ni—O = 2.0438 (12) Å] in a slightly distorted octahedral geometry. The coordinating water molecules are hydrogen bonded to the non-coordinating carboxylate O atom of the TMB ligand [O...O = 2.593 (3) Å], enclosing an S(6) hydrogen-bonding motif. Two solvent water molecules are also present in the formula unit. In the crystal, a network of intermolecular N—H...O and O—H...O hydrogen bonds link the complexes into a three-dimensional array. Hirshfeld surface analysis indicates that the most important contributions for the crystal packing are from H...H (59.8%), O...H/H...O (20.2%) and C...H/H...C (13.7%) interactions.
APA, Harvard, Vancouver, ISO, and other styles
38

Okeke, Ugochukwu, Yilma Gultneh, and Ray J. Butcher. "(Acetonitrile-κN)aqua[N,N′-bis(pyridin-2-ylmethyl)ethane-1,2-diamine-κ4N,N′,N′′,N′′′]zinc(II) perchlorate." Acta Crystallographica Section E Crystallographic Communications 73, no. 10 (September 29, 2017): 1568–71. http://dx.doi.org/10.1107/s2056989017013603.

Full text
Abstract:
The structure of the title compound, [Zn(C14H18N4)(C2H3N)(H2O)](ClO4)2, contains a six-coordinate cation consisting of the tetradentate bispicen ligand, coordinated water, and coordinated acetonitrile, with the latter two ligands adopting acisconfiguration. There are two formula units in the asymmetric unit. Both cations show almost identical structural features with the bispicen ligand adopting the more commoncis-β conformation. One of the four perchlorate anions is disordered over two positions, with occupancies of 0.9090 (15) and 0.0910 (15). There is extensive inter-ionic hydrogen bonding between the perchlorate anions and O—H and N—H groups in the cations, including a bifurcated hydrogen bond between an N—H group and two O atoms of one perchlorate anion. As a result of this extended hydrogen-bond network, the ions are linked into a complex three-dimensional array.
APA, Harvard, Vancouver, ISO, and other styles
39

Wölper, Christoph, Peter G. Jones, and Armand Blaschette. "Polysulfonylamine, CLXXX [1]. Bis(triphenylphosphoranyliden)ammonium-di(methansulfonyl)amid: Eine Kristallstruktur mit CH/π-Wechselwirkungen zwischen Methylund Phenylgruppen / Polysulfonylamines, CLXXX [1]. Bis(triphenylphosphoranylidene)ammonium Di(methanesulfonyl)amide:A Crystal Structure Featuring Methyl-Phenyl CH/π Interactions." Zeitschrift für Naturforschung B 62, no. 9 (September 1, 2007): 1167–73. http://dx.doi.org/10.1515/znb-2007-0910.

Full text
Abstract:
The structure of [Ph3PNPPh3][(MeSO2)2N] (monoclinic, P21/c, Z ' = 1; single-crystal X-ray diffraction at −140 °C) displays chains of relatively small anions, surrounded by four parallel columns of bulky cations, giving a rectangular array. In contrast to a previously reported series of onium dimesylamides with smaller cations, interanion hydrogen bonding CMe-H· · · O, originating from the inductively activated methyl groups, is not observed in the present structure. In consequence, the packing arrangement is determined by multiple phenyl embraces between cations, weak cation-anion hydrogen bonds CPh-H· · ·O/N and, most notably, by methyl-phenyl CH/π interactions between anions and cations. One methyl group forms two C-H· · ·π contacts to different phenyl rings (hydrogen-to-centroid distances: 272 and 285 pm), the other lies in a tripodal mode above an aromatic ring plane (hydrogen-to-centroid distances: ca. 330 pm).
APA, Harvard, Vancouver, ISO, and other styles
40

Sabbaghi, Fahimeh, Mehrdad Pourayoubi, Abolghasem Farhadipour, Nazila Ghorbanian, and Pavel V. Andreev. "A novel tubular hydrogen-bond pattern in a new diazaphosphole oxide: a combination of X-ray crystallography and theoretical study of hydrogen bonds." Acta Crystallographica Section C Structural Chemistry 73, no. 7 (June 8, 2017): 508–16. http://dx.doi.org/10.1107/s205322961700794x.

Full text
Abstract:
In the structure of 2-(4-chloroanilino)-1,3,2λ4-diazaphosphol-2-one, C12H11ClN3OP, each molecule is connected with four neighbouring molecules through (N—H)2...O hydrogen bonds. These hydrogen bonds form a tubular arrangement along the [001] direction built from R 3 3(12) and R 4 3(14) hydrogen-bond ring motifs, combined with a C(4) chain motif. The hole constructed in the tubular architecture includes a 12-atom arrangement (three P, three N, three O and three H atoms) belonging to three adjacent molecules hydrogen bonded to each other. One of the N—H groups of the diazaphosphole ring, not co-operating in classical hydrogen bonding, takes part in an N—H...π interaction. This interaction occurs within the tubular array and does not change the dimension of the hydrogen-bond pattern. The energies of the N—H...O and N—H...π hydrogen bonds were studied by NBO (natural bond orbital) analysis, using the experimental hydrogen-bonded cluster of molecules as the input file for the chemical calculations. In the 1H NMR experiment, the nitrogen-bound proton of the diazaphosphole ring has a high value of 17.2 Hz for the 2 J H–P coupling constant.
APA, Harvard, Vancouver, ISO, and other styles
41

Ferguson, G., C. Glidewell, R. M. Gregson, and P. R. Meehan. "Crystal Engineering Using Tris-phenols. Cross-Linked, Pairwise-Interwoven Two-Dimensional Nets in the 2:1 Adduct of 1,1,1-Tris(4-hydroxyphenyl)ethane with 1,2-Diaminoethane." Acta Crystallographica Section B Structural Science 54, no. 3 (June 1, 1998): 330–38. http://dx.doi.org/10.1107/s010876819701495x.

Full text
Abstract:
In 1,1,1-tris(4-hydroxyphenyl)ethane–1,2-diaminoethane (2/1), [CH3C(C6H4OH)3]2.H2NCH2CH2NH2 (1), triclinic, P1¯, with Z = 2, a = 10.9430 (12), b = 11.1075 (12), c = 15.249 (2) Å, α = 98.672 (15), β = 96.312 (10), γ = 98.377 (13)°, the tris-phenol units form continuous two-dimensional nets, built from pseudo-hexagonal R^4_4(38) rings, interwoven pairs of which are cross-linked by the 1,2-diaminoethane units. Each tris-phenol unit acts as a triple donor, forming two O—H...O and one O—H...N hydrogen bonds, and as a double acceptor in two O—H...O hydrogen bonds: the diamine unit, in which the CH2 groups are disordered over two sets of sites with site-occupation factors of 0.740 (5) and 0.260 (5), respectively, acts as a double acceptor only and the N—H bonds play no role in the hydrogen bonding. The O...O distances in the O—H...O hydrogen bonds are 2.642 (2), 2.690 (2), 2.810 (2) and 2.835 (2) Å, and the two independent O...N distances are both 2.665 (3) Å. Adjacent bilayers are connected into a continuous three-dimensional array by C—H...O hydrogen bonds, all having a C...O distance of 3.468 (4) Å.
APA, Harvard, Vancouver, ISO, and other styles
42

Hökelek, Tuncer, Elif Özbek, Mustafa Sertçelik, Çiğdem Şahin Yenice, and Hacali Necefoğlu. "Crystal structure and Hirshfeld surface analysis of hexakis(μ-benzoato-κ2O:O′)bis(pyridine-3-carbonitrile-κN1)trizinc(II)." Acta Crystallographica Section E Crystallographic Communications 73, no. 12 (November 28, 2017): 1966–70. http://dx.doi.org/10.1107/s2056989017016899.

Full text
Abstract:
The asymmetric unit of the title complex, [Zn3(C7H5O2)6(C6H4N2)2], contains one half of the complex molecule,i.e.one and a half ZnIIcations, three benzoate (Bnz) and one pyridine-3-carbonitrile (Cpy) molecule; the Bnz anions act as bidentate ligands through the carboxylate O atoms, while the Cpy ligand acts as a monodentate N(pyridine)-bonding ligand. The complete centrosymmetric trinuclear complex thus comprises a linear array of three ZnIIcations. The central ZnIIcation shows an octahedral coordination and is bridged to each of the terminal ZnIIcations by three Bnz anions. By additional coordination of the CPy ligand, the terminal ZnIIcations adopt a trigonal–pyramidal coordination environment. In the crystal, the Bnz anions link to the Cpy N atomsviaweak C—H...N hydrogen bonds, forming a two-dimensional network. C—H...π and π–π interactions [between the benzene and pyridine rings of adjacent molecules with an intercentroid distance of 3.850 (4) Å] help to consolidate a three-dimensional architecture. The Hirshfeld surface analysis confirms the role of H-atom contacts in establishing the packing.
APA, Harvard, Vancouver, ISO, and other styles
43

Okeke, Ugochukwu, Raymond Otchere, Yilma Gultneh, and Ray J. Butcher. "Crystal structure of tetrakis(μ2-(E)-2,4-dibromo-6-{[2-(pyridin-2-yl)ethyl]iminomethyl}phenolato)trizinc bis(perchlorate) acetonitrile disolvate." Acta Crystallographica Section E Crystallographic Communications 74, no. 9 (August 31, 2018): 1380–83. http://dx.doi.org/10.1107/s2056989018012100.

Full text
Abstract:
The title compound, [Zn3(C14H11Br2N2O)4](ClO4)2·2CH3CN, crystallizes as a symmetrical trinuclear cation with all three metal atoms being located on a twofold rotation axis. It contains a tetrahedral ZnII atom that bridges two six-coordinate ZnII atoms. The complex contains N- and O-donor atoms of four tridentate 2,4-dibromo-6-{[2-(pyridin-2-yl)ethyl]iminomethyl}phenolate ligands. The ratio of ZnII atoms to ligands is 3:4. The two terminal ZnII cations adopt distorted octahedral geometries and the central ZnII cation adopts a distorted tetrahedral geometry. In the cation there are π–π interactions between the dibromophenyl rings, as well as halogen-bonding interactions between the dibromophenyl rings in the cation, which stabilize its conformation. In addition, there are C—H...O interactions between the anions and both the cations and solvent molecules as well as C—H...N interactions between the cation and solvent molecules. These interspecies interactions link the cations, anions and solvent molecules into a complex three-dimensional array
APA, Harvard, Vancouver, ISO, and other styles
44

Thoelen, Felix, and Walter Frank. "Crystal engineering with short-chained amphiphiles: decasodium octa-n-butanesulfonate di-μ-chlorido-bis[dichloridopalladate(II)] tetrahydrate, a layered inorganic–organic hybrid material." Acta Crystallographica Section E Crystallographic Communications 75, no. 5 (April 2, 2019): 557–61. http://dx.doi.org/10.1107/s2056989019004201.

Full text
Abstract:
In the course of crystal-engineering experiments, crystals of the hydrated title salt, Na10[Pd2Cl6](C4H9SO3)8·4H2O, were obtained from a water/2-propanol solution of sodium n-butanesulfonate and sodium tetrachloridopalladate(II). In the crystal, sodium n-butanesulfonate anions and water molecules are arranged in an amphiphilic inverse bilayered cationic array represented by the formula {[Na10(C4H9SO3)8(H2O)4]2+} n . Within this lamellar array: (i) a hydrophilic layer region parallel to the bc plane is established by the Na+ cations, the H2O molecules (as aqua ligands in κNa,κNa′-bridging coordination mode) and the O3S– groups of the sulfonate ions, and (ii) hydrophobic regions are present containing all the n-butyl groups in an almost parallel orientation, with the chain direction approximately perpendicular to the aforementioned hydrophilic layer. Unexpectedly, the flat centrosymmetric [Pd2Cl6]2− anion in the structure is placed between the butyl groups, within the hydrophobic regions, but due to its appropriate length primarily bonded to the hydrophilic `inorganic' layer regions above and below the hydrophobic area via Pd—Clt...Na- and Pd—Clt...H—O(H)—Na-type (Clt is terminal chloride) interactions. In addition to these hydrogen-bonding interactions, both aqua ligands are engaged in charge-supported S—O...H—O hydrogen bonds of a motif characterized by the D 4 3(9) graph-set descriptor within the hydrophilic region. The crystal structure of the title compound is the first reported for a metal n-butanesulfonate.
APA, Harvard, Vancouver, ISO, and other styles
45

Langford, Steven J., and Clint P. Woodward. "Synthesis, characterisation and X-ray structure of a novel porphyrin array employing Zn–O and O–H…O bonding motifs." Polyhedron 26, no. 2 (January 2007): 338–43. http://dx.doi.org/10.1016/j.poly.2006.06.002.

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

Hijji, Yousef, Ellis Benjamin, Jerry P. Jasinski, and Ray J. Butcher. "Crystal structure of the thalidomide analog (3aR*,7aS*)-2-(2,6-dioxopiperidin-3-yl)hexahydro-1H-isoindole-1,3(2H)-dione." Acta Crystallographica Section E Crystallographic Communications 74, no. 11 (October 16, 2018): 1595–98. http://dx.doi.org/10.1107/s2056989018014317.

Full text
Abstract:
The title compound, C13H16N2O4, crystallizes in the monoclinic centrosymmetric space group, P21/c, with four molecules in the asymmetric unit, thus there is no crystallographically imposed symmetry and it is a racemic mixture. The structure consists of a six-membered unsaturated ring bound to a five-membered pyrrolidine-2,5-dione ring N-bound to a six-membered piperidine-2,6-dione ring and thus has the same basic skeleton as thalidomide, except for the six-membered unsaturated ring substituted for the aromatic ring. In the crystal, the molecules are linked into inversion dimers by R 2 2(8) hydrogen bonding involving the N—H group. In addition, there are bifurcated C—H...O interactions involving one of the O atoms on the pyrrolidine-2,5-dione with graph-set notation R 1 2(5). These interactions along with C—H...O interactions involving one of the O atoms on the piperidine-2,6-dione ring link the molecules into a complex three-dimensional array. There is pseudomerohedral twinning present which results from a 180° rotation about the [100] reciprocal lattice direction and with a twin law of 1 0 0 0 \overline{1} 0 0 0 \overline{1} [BASF 0.044 (1)].
APA, Harvard, Vancouver, ISO, and other styles
47

Sangeetha, Ramalingam, Kasthuri Balasubramani, Kaliyaperumal Thanigaimani, and Savaridasson Jose Kavitha. "Crystal structure and Hirshfeld surface analysis of 2,4-diamino-6-phenyl-1,3,5-triazin-1-ium 4-methylbenzenesulfonate." Acta Crystallographica Section E Crystallographic Communications 74, no. 8 (July 27, 2018): 1159–62. http://dx.doi.org/10.1107/s2056989018010368.

Full text
Abstract:
In the title molecular salt, C9H10N5 +·C7H7O3S−, the asymmetric unit consists of a 2,4-diamino-6-phenyl-1,3,5-triazin-1-ium cation and a 4-methylbenzenesulfonate anion. The cation is protonated at the N atom lying between the amine and phenyl substituents. The protonated N and amino-group N atoms are involved in hydrogen bonding with the sulfonate O atoms through a pair of intermolecular N—H...O hydrogen bonds, giving rise to a hydrogen-bonded cyclic motif with R 2 2(8) graph-set notation. The inversion-related molecules are further linked by four N—H...O intermolecular interactions to produce a complementary DDAA (D = donor, A = acceptor) hydrogen-bonded array, forming R 2 2(8), R 4 2(8) and R 2 2(8) ring motifs. The centrosymmetrically paired cations form R 2 2(8) ring motifs through base-pairing via N—H...N hydrogen bonds. In addition, another R 3 3(10) motif is formed between centrosymetrically paired cations and a sulfonate anion via N—H...O hydrogen bonds. The crystal structure also features weak S=O...π and π–π interactions. Hirshfeld surface and fingerprint plots were employed in order to further study the intermolecular interactions.
APA, Harvard, Vancouver, ISO, and other styles
48

Liu, Chun-Sen, Min Hu, Song-Tao Ma, Qiang Zhang, Li-Ming Zhou, Li-Jun Gao, and Shao-Ming Fang. "Coordination Polymers with a Bulky Perylene-Based Tetracarboxylate Ligand: Syntheses, Crystal Structures, and Luminescent Properties." Australian Journal of Chemistry 63, no. 3 (2010): 463. http://dx.doi.org/10.1071/ch09411.

Full text
Abstract:
To explore the coordination possibilities of perylene-based ligands with a larger conjugated π-system, four ZnII, MnII, and CoII coordination polymers with perylene-3,4,9,10-tetracarboxylate (ptc) and the chelating 1,10-phenanthroline (phen) ligands were synthesized and characterized: {[Zn2(ptc)(phen)2](H2O)10}∞ (1), {[Zn3(ptc)(OH)2(phen)2](H2O)3}∞ (2), {[Mn(ptc)0.5(phen)(H2O)2](H2O)1.5}∞ (3), and {[Co(ptc)0.5(phen)(H2O)2](H2O)2.5}∞ (4). Structural analysis reveals that complexes 1 and 2 both take one-dimensional polymeric chain structures with dinuclear and trinuclear units as nodes, respectively, which are further extended via the accessorial secondary interchain interactions, such as C–H···O H-bonding or aromatic π···π stacking interactions, to give rise to the relevant higher-dimensional frameworks. Compound 3 has a two-dimensional sheet structure that is further assembled to form a three-dimensional framework by interlayer π···π stacking interactions. Complex 4 is a one-dimensional ribbon-like array structure that is interlinked by the co-effects of intermolecular π···π stacking and C–H···π supramolecular interactions, resulting in a higher-dimensional framework from the different crystallographic directions. Moreover, complexes 1–4 exhibit strong solid-state luminescence emissions at room temperature, which mainly originate from intraligand π→π* transitions of ptc.
APA, Harvard, Vancouver, ISO, and other styles
49

Egekenze, Rita, Yilma Gultneh, and Ray J. Butcher. "Bis(3,5-dimethoxy-2-{[2-(pyridin-2-yl)ethylimino-κN]methyl}phenolato-κO)bis(dimethyl sulfoxide)manganese(III) perchlorate methanol 0.774-solvate." Acta Crystallographica Section E Crystallographic Communications 73, no. 10 (September 15, 2017): 1479–82. http://dx.doi.org/10.1107/s205698901701204x.

Full text
Abstract:
The title compound, [Mn(C16H17N2O3)2(C2H6OS)2]ClO4·0.774CH3OH, comprises a central octahedrally coordinated MnIIIcation, with two bidentate Schiff base ligands occupying the equatorial positions and two dimethyl sulfoxide (DMSO) ligands occupying the axial positions. There are two independant cations in the asymmetric unit, with the MnIIIatoms of both cations being positioned on crystallographic centers of inversion. The perchlorate anion is disordered over two equivalent conformations, with occupancies of 0.744 (3) and 0.226 (3). In addition, there is a methanol solvent molecule in the crystal lattice that is too close to the minor component of the perchlorate anion to be present simultaneously and thus it was refined to have the same occupancy as the major component of this anion. There is a Jahn–Teller distortion which results in Mn—ODMSOaxial bond lengths of 2.2365 (12) and 2.2368 (12) Å in the two cations. In the crystal, intermolecular π–π stacking between the non-coordinating pyridine rings of each cation is observed. This π–π stacking, along with extensive O—H...O hydrogen bonding and C—H...O interactions, link the components into a complex three-dimensional array.
APA, Harvard, Vancouver, ISO, and other styles
50

Trella, Thomas, and Walter Frank. "Hexaaquaaluminium(III) tris(methanesulfonate)." Acta Crystallographica Section E Structure Reports Online 68, no. 8 (July 28, 2012): m1136—m1137. http://dx.doi.org/10.1107/s1600536812033235.

Full text
Abstract:
The title compound, [Al(H2O)6](CH3SO3)3(common name: aluminium methanesulfonate hexahydrate), was crystallized from an aqueous solution prepared by the precipitation reaction of aluminium sulfate and barium methanesulfonate. Its crystal structure is the first of the boron group methanesulfonates to be determined. The characteristic building block is a centrosymmetric unit containing two hexaaquaaluminium cations that are connected to each other by two O atoms of the –SO3groups in an O—H...O...H—O sequence. Further O—H...O hydrogen bonding links these blocks in orthogonal directions – along [010] forming a double chain array, along [10-1] forming a layered arrangement of parallel chains and along [101] forming a three-dimensional network. As indicated by the O...O distances of 2.600 (3)–2.715 (3) Å, the hydrogen bonds are from medium–strong to strong. A further structural feature is the arrangement of two and four methyl groups, respectively, establishing `hydrophobic islands' of different size, all positioned in a layer-like region perpendicular to [101]. The only other building block within this region is one of the –SO3groups giving a local connection between the hydrophilic structural regions on both sides of the `hydrophobic' one. Thermal analysis indicates that a stepwise dehydration process starts at about 413 K and proceedsviathe respective penta- and dihydrate until the compound completely decomposes at about 688 K.
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