o&

Crystal structure, Hirshfeld surface analysis and inter­action energy and DFT studies of (S)-10-propargyl­pyrrolo­[2,1-c][1,4]benzodiazepine-5,11-dione

The title compound, C15H14N2O2, consists of pyrrole and benzodiazepine units linked to a propargyl moiety, where the pyrrole and diazepine rings adopt half-chair and boat conformations, respectively. The absolute configuration was assigned on the the basis of l-proline, which was used in the synthesis of benzodiazepine. In the crystal, weak C—HBnz⋯ODiazp and C—HProprg⋯ODiazp (Bnz = benzene, Diazp = diazepine and Proprg = proparg­yl) hydrogen bonds link the mol­ecules into two-dimensional networks parallel to the bc plane, enclosing R44(28) ring motifs, with the networks forming oblique stacks along the a-axis direction. The Hirshfeld surface analysis of the crystal structure indicates that the most important contributions for the crystal packing are from H⋯H (49.8%), H⋯C/C⋯H (25.7%) and H⋯O/O⋯H (20.1%) inter­actions. Hydrogen bonding and van der Waals inter­actions are the dominant inter­actions in the crystal packing. Computational chemistry indicates that in the crystal, C—H⋯O hydrogen-bond energies are 38.8 (for C—HBnz⋯ODiazp) and 27.1 (for C—HProprg⋯ODiazp) kJ mol−1. Density functional theory (DFT) optimized structures at the B3LYP/6–311 G(d,p) level are compared with the experimentally determined mol­ecular structure in the solid state. The HOMO–LUMO behaviour was elucidated to determine the energy gap.




o&

Bis(4-hy­droxy-N-isopropyl-N-methyl­trypt­ammo­nium) fumarate: a new crystalline form of miprocin

The title compound, bis­(4-hy­droxy-N-isopropyl-N-methyl­tryptammonium) (4-HO-MiPT) fumarate (systematic name: bis­{[2-(4-hy­droxy-1H-indol-3-yl)eth­yl](meth­yl)propan-2-yl­aza­nium} but-2-enedioate), 2C14H21N2O+·C4H2O42−, has a singly protonated tryptammonium cation and one half of a fumarate dianion in the asymmetric unit. The tryptammonium and fumarate ions are held together in one-dimensional chains by N—H⋯O and O—H⋯O hydrogen bonds. These chains are a combination of R42(20) rings, and C22(15) and C44(30) parallel chains along (110). They are further consolidated by N—H⋯π inter­actions. There are two two-component types of disorder impacting the tryptammonium fragment with a 0.753 (7):0.247 (7) occupancy ratio and one of the fumarate oxygen atoms with a 0.73 (8):0.27 (8) ratio.




o&

The first coordination compound of 6-fluoro­nicotinate: the crystal structure of a one-dimensional nickel(II) coordination polymer containing the mixed ligands 6-fluoro­nicotinate and 4,4'-bi­pyridine

A one-dimensional nickel(II) coordination polymer with the mixed ligands 6-fluoro­nicotinate (6-Fnic) and 4,4'-bi­pyridine (4,4'-bpy), namely, catena-poly[[di­aqua­bis­(6-fluoro­pyridine-3-carboxyl­ato-κO)nickel(II)]-μ-4,4'-bi­pyri­dine-κ2N:N'] trihydrate], {[Ni(6-Fnic)2(4,4'-bpy)(H2O)2]·3H2O}n, (1), was prepared by the reaction of nickel(II) sulfate hepta­hydrate, 6-fluoro­nicotinic acid (C6H4FNO2) and 4,4'-bi­pyridine (C10H8N2) in a mixture of water and ethanol. The nickel(II) ion in 1 is octa­hedrally coordinated by the O atoms of two water mol­ecules, two O atoms from O-monodentate 6-fluoro­nicotinate ligands and two N atoms from bridging 4,4'-bi­pyridine ligands, forming a trans isomer. The bridging 4,4'-bi­pyridine ligands connect symmetry-related nickel(II) ions into infinite one-dimensional polymeric chains running in the [1overline{1}0] direction. In the extended structure of 1, the polymeric chains and lattice water mol­ecules are connected into a three-dimensional hydrogen-bonded network via strong O—H⋯O and O—H⋯N hydrogen bonds, leading to the formation of distinct hydrogen-bond ring motifs: octa­meric R88(24) and hexa­meric R86(16) loops.




o&

Intra­molecular 1,5-S⋯N σ-hole inter­action in (E)-N'-(pyridin-4-yl­methyl­idene)thio­phene-2-carbohydrazide

The title compound, C11H9N3OS, (I), crystallizes in the monoclinic space group P21/n. The mol­ecular conformation is nearly planar and features an intra­molecular chalcogen bond between the thio­phene S and the imine N atoms. Within the crystal, the strongest inter­actions between mol­ecules are the N—H⋯O hydrogen bonds, which organize them into inversion dimers. The dimers are linked through short C—H⋯N contacts and are stacked into layers propagating in the (001) plane. The crystal structure features π–π stacking between the pyridine aromatic ring and the azomethine double bond. The calculated energies of pairwise inter­molecular inter­actions within the stacks are considerably larger than those found for the inter­actions between the layers.




o&

Crystal structure of 1,4,8,11-tetra­methyl-1,4,8,11-tetra­azonia­cyclo­tetra­decane bis­[chlorido­chromate(VI)] dichloride from synchrotron X-ray data

The crystal structure of title compound, (C14H36N4)[CrO3Cl]2Cl2, has been determined by synchrotron radiation X-ray crystallography at 220 K. The macrocyclic cation lies across a crystallographic inversion center and hence the asymmetric unit contains one half of the organic cation, one chloro­chromate anion and one chloride anion. Both the Cl− anion and chloro­chromate Cl atom are involved in hydrogen bonding. In the crystal, hydrogen bonds involving the 1,4,8,11-tetra­methyl-1,4,8,11-tetra­azonia­cyclo­tetra­decane (TMC) N—H groups and C—H groups as donor groups and three O atoms of the chloro­chromate and the chloride anion as acceptor groups link the components, giving rise to a three-dimensional network.




o&

Silver(I) nitrate two-dimensional coordination polymers of two new pyrazine­thio­phane ligands: 5,7-di­hydro-1H,3H-dithieno[3,4-b:3',4'-e]pyrazine and 3,4,8,10,11,13-hexa­hydro-1H,6H-bis­([1,4]di­thio­cino)[6,7-b:6',7'-e

The two new pyrazine­ophanes, 5,7-di­hydro-1H,3H-dithieno[3,4-b:3',4'-e]pyrazine, C8H8N2S2, L1, and 3,4,8,10,11,13-hexa­hydro-1H,6H-bis­([1,4]di­thio­cino)[6,7-b:6',7'-e]pyrazine, C12H16N2S4, L2, both crystallize with half a mol­ecule in the asymmetric unit; the whole mol­ecules are generated by inversion symmetry. The mol­ecule of L1, which is planar (r.m.s. deviation = 0.008 Å), consists of two sulfur atoms linked by a rigid tetra-2,3,5,6-methyl­ene­pyrazine unit, forming planar five-membered rings. The mol­ecule of L2 is step-shaped and consists of two S–CH2–CH2–S chains linked by the central rigid tetra-2,3,5,6-methyl­ene­pyrazine unit, forming eight-membered rings that have twist-boat-chair con­fig­urations. In the crystals of both compounds, there are no significant inter­molecular inter­actions present. The reaction of L1 with silver nitrate leads to the formation of a two-dimensional coordination polymer, poly[(μ-5,7-di­hydro-1H,3H-dithieno[3,4-b;3',4'-e]pyrazine-κ2S:S')(μ-nitrato-κ2O:O')silver(I)], [Ag(NO3)(C8H8N2S2)]n, (I), with the nitrato anion bridging two equivalent silver atoms. The central pyrazine ring is situated about an inversion center and the silver atom lies on a twofold rotation axis that bis­ects the nitrato anion. The silver atom has a fourfold AgO2S2 coordination sphere with a distorted shape. The reaction of L2 with silver nitrate also leads to the formation of a two-dimensional coordination polymer, poly[[μ33,4,8,10,11,13-hexa­hydro-1H,6H-bis­([1,4]di­thio­cino)[6,7-b;6',7'-e]pyrazine-κ3S:S':S''](nitrato-κO)silver(I)], [Ag(NO3)(C12H16N2S4)]n, (II), with the nitrate anion coordinating in a monodentate manner to the silver atom. The silver atom has a fourfold AgOS3 coordination sphere with a distorted shape. In the crystals of both complexes, the networks are linked by C—H⋯O hydrogen bonds, forming supra­molecular frameworks. There are additional C—H⋯S contacts present in the supra­molecular framework of II.




o&

Crystal structures of trans-acetyl­dicarbon­yl(η5-cyclo­penta­dien­yl)(1,3,5-tri­aza-7-phosphaadamantane)molybdenum(II) and trans-acetyl­di­carbon­yl(η5-cyclo­penta­dien­yl)(3,7-diacetyl-1,3,7-tr

The title compounds, [Mo(C5H5)(COCH3)(C6H12N3P)(CO)2], (1), and [Mo(C5H5)(COCH3)(C9H16N3O2P)(C6H5)2))(CO)2], (2), have been prepared by phosphine-induced migratory insertion from [Mo(C5H5)(CO)3(CH3)]. The mol­ecular structures of these complexes are quite similar, exhibiting a four-legged piano-stool geometry with trans-disposed carbonyl ligands. The extended structures of complexes (1) and (2) differ substanti­ally. For complex (1), the molybdenum acetyl unit plays a dominant role in the organization of the extended structure, joining the mol­ecules into centrosymmetrical dimers through C—H⋯O inter­actions with a cyclo­penta­dienyl ligand of a neighboring mol­ecule, and these dimers are linked into layers parallel to (100) by C—H⋯O inter­actions between the molybdenum acetyl and the cyclo­penta­dienyl ligand of another neighbor. The extended structure of (2) is dominated by C—H⋯O inter­actions involving the carbonyl groups of the acetamide groups of the DAPTA ligand, which join the mol­ecules into centrosymmetrical dimers and link them into chains along [010]. Additional C—H⋯O inter­actions between the molybdenum acetyl oxygen atom and an acetamide methyl group join the chains into layers parallel to (101).




o&

Synthesis and crystal structures of two 1,3-di(alk­yloxy)-2-(methyl­sulfan­yl)imidazolium tetra­fluorido­borates

Two salts were prepared by methyl­ation of the respective imidazoline-2-thione at the sulfur atom, using Meerwein's salt (tri­methyl­oxonium tetra­fluorido­borate) in CH2Cl2. 1,3-Dimeth­oxy-2-(methyl­sulfan­yl)imidazolium tetra­fluorido­borate (1), C6H11N2O2S+·BF4−, displays a syn conformation of its two meth­oxy groups relative to each other whereas the two benz­yloxy groups present in 1,3-dibenz­yloxy-2-(methyl­sulfan­yl)imidazolium tetra­fluorido­borate (2), C18H19N2O2S+·BF4−, adopt an anti conformation. In the mol­ecules of 1 and 2, the methyl­sulfanyl group is rotated out of the plane of the respective heterocyclic ring. In both crystal structures, inter­molecular inter­actions are dominated by C—H⋯F—B contacts, leading to three-dimensional networks. The tetra­fluorido­borate counter-ion of 2 is disordered over three orientations (occupancy ratio 0.42:0.34:0.24), which are related by rotation about one of the B—F bonds.




o&

Structural investigation of methyl 3-(4-fluoro­benzo­yl)-7-methyl-2-phenyl­indolizine-1-carboxyl­ate, an inhibitory drug towards Mycobacterium tuberculosis

The title compound, C24H18FNO3, crystallizes in the monoclinic centrosymmetric space group P21/n and its mol­ecular conformation is stabilized via C—H⋯O intra­molecular inter­actions. The supra­molecular network mainly comprises C—H⋯O, C—H⋯F and C—H⋯π inter­actions, which contribute towards the formation of the crystal structure. The different inter­molecular inter­actions have been further analysed via Hirshfeld surface analysis and fingerprint plots.




o&

Crystal structure, Hirshfeld surface and frontier mol­ecular orbital analysis of 10-benzyl-9-(3-eth­oxy-4-hy­droxy­phen­yl)-3,3,6,6-tetra­methyl-3,4,6,7,9,10-hexa­hydro­acridine-1,8(2H,5H)-dione

In the fused ring system of the title compound, C32H37NO4, the central di­hydro­pyridine ring adopts a flattened boat conformation, the mean and maximum deviations of the di­hydro­pyridine ring being 0.1429 (2) and 0.2621 (2) Å, respectively. The two cyclo­hexenone rings adopt envelope conformations with the tetra­substituted C atoms as flap atoms. The benzene and phenyl rings form dihedral angles of 85.81 (2) and 88.90 (2)°, respectively, with the mean plane of the di­hydro­pyridine ring. In the crystal, mol­ecules are linked via an O—H⋯O hydrogen bond, forming a helical chain along the b-axis direction. A Hirshfeld surface analysis indicates that the most important contributions to the crystal packing are from H⋯H (65.2%), O⋯H/H⋯O (18.8%) and C⋯H/H⋯C (13.9%) contacts. Quantum chemical calculations for the frontier mol­ecular orbitals were undertake to determine the chemical reactivity of the title compound.




o&

Crystal structure, Hirshfeld surface analysis and inter­action energy, DFT and anti­bacterial activity studies of ethyl 2-[(2Z)-2-(2-chloro­benzyl­idene)-3-oxo-3,4-di­hydro-2H-1,4-benzo­thia­zin-4-yl]acetate

The title compound, C19H16ClNO3S, consists of chloro­phenyl methyl­idene and di­hydro­benzo­thia­zine units linked to an acetate moiety, where the thia­zine ring adopts a screw-boat conformation. In the crystal, two sets of weak C—HPh⋯ODbt (Ph = phenyl and Dbt = di­hydro­benzo­thia­zine) hydrogen bonds form layers of mol­ecules parallel to the bc plane. The layers stack along the a-axis direction with inter­calation of the ester chains. The crystal studied was a two component twin with a refined BASF of 0.34961 (5). The Hirshfeld surface analysis of the crystal structure indicates that the most important contributions to the crystal packing are from H⋯H (37.5%), H⋯C/C⋯H (24.6%) and H⋯O/O⋯H (16.7%) inter­actions. Hydrogen-bonding and van der Waals inter­actions are the dominant inter­actions in the crystal packing. Computational chemistry indicates that in the crystal, C—HPh⋯ODbt hydrogen bond energies are 38.3 and 30.3 kJ mol−1. Density functional theory (DFT) optimized structures at the B3LYP/ 6–311 G(d,p) level are compared with the experimentally determined mol­ecular structure in the solid state. The HOMO–LUMO behaviour was elucidated to determine the energy gap. Moreover, the anti­bacterial activity of the title compound has been evaluated against gram-positive and gram-negative bacteria.




o&

Synthesis and crystal structure of a 6-chloro­nicotinate salt of a one-dimensional cationic nickel(II) coordination polymer with 4,4'-bi­pyridine

A 6-chloro­nicotinate (6-Clnic) salt of a one-dimensional cationic nickel(II) coordination polymer with 4,4'-bi­pyridine (4,4'-bpy), namely, catena-poly[[[tetra­aqua­nickel(II)]-μ-4,4'-bi­pyridine-κ2N:N'] bis­(6-chloro­nicotinate) tetra­hydrate], {[Ni(C10H8N2)(H2O)4](C6H3ClNO2)2·4H2O}n or {[Ni(4,4'-bpy)(H2O)4](6-Clnic)2·4H2O}n, (1), was prepared by the reaction of nickel(II) sulfate hepta­hydrate, 6-chloro­nicotinic acid and 4,4'-bi­pyridine in a mixture of water and ethanol. The mol­ecular structure of 1 comprises a one-dimensional polymeric {[Ni(4,4'-bpy)(H2O)4]2+}n cation, two 6-chloro­nicotinate anions and four water mol­ecules of crystallization per repeating polymeric unit. The nickel(II) ion in the polymeric cation is octa­hedrally coordinated by four water mol­ecule O atoms and by two 4,4'-bi­pyridine N atoms in the trans position. The 4,4'-bi­pyridine ligands act as bridges and, thus, connect the symmetry-related nickel(II) ions into an infinite one-dimensional polymeric chain extending along the b-axis direction. In the extended structure of 1, the polymeric chains of {[Ni(4,4'-bpy)(H2O)4]2+}n, the 6-chloro­nicotinate anions and the water mol­ecules of crystallization are assembled into an infinite three-dimensional hydrogen-bonded network via strong O—H⋯O and O—H⋯N hydrogen bonds, leading to the formation of the representative hydrogen-bonded ring motifs: tetra­meric R24(8) and R44(10) loops, a dimeric R22(8) loop and a penta­meric R45(16) loop.




o&

Crystal structures of (η4-cyclo­octa-1,5-diene)bis(1,3-di­methyl­imidazol-2-yl­idene)iridium(I) iodide and (η4-cyclo­octa-1,5-diene)bis­(1,3-di­ethyl­imidazol-2-yl­idene)iridium(I) iodide

The title complexes, (η4-cyclo­octa-1,5-diene)bis­(1,3-di­methyl­imidazol-2-yl­idene)iridium(I) iodide, [Ir(C5H8N2)2(C8H12)]I, (1) and (η4-cyclo­octa-1,5-di­ene)bis­(1,3-di­ethyl­imidazol-2-yl­idene)iridium(I) iodide, [Ir(C7H12N2)2(C8H12)]I, (2), were prepared using a modified literature method. After carrying out the oxidative addition of the amino acid l-proline to [Ir(COD)(IMe)2]I in water and slowly cooling the reaction to room temperature, a suitable crystal of 1 was obtained and analyzed by single-crystal X-ray diffraction at 100 K. Although this crystal structure has previously been reported in the Pbam space group, it was highly disordered and precise atomic coordinates were not calculated. A single crystal of 2 was also obtained by heating the complex in water and letting it slowly cool to room temperature. Complex 1 was found to crystallize in the monoclinic space group C2/m, while 2 crystallizes in the ortho­rhom­bic space group Pccn, both with Z = 4.




o&

Synthesis, crystal structure and Hirshfeld and thermal analysis of bis[benzyl 2-(heptan-4-yl­idene)hydrazine-1-carboxyl­ate-κ2N2,O]bis(thio­cyanato)­nickel(II)

The title centrosymmetric NiII complex, [Ni(NCS)2(C15H22N2O2)2], crystallizes with one half mol­ecule in the asymmetric unit of the monoclinic unit cell. The complex adopts an octa­hedral coordination geometry with two mutually trans benzyl-2-(heptan-4-yl­idene)hydrazine-1-carboxyl­ate ligands in the equatorial plane with the axial positions occupied by N-bound thio­cyanato ligands. The overall conformation of the mol­ecule is also affected by two, inversion-related, intra­molecular C—H⋯O hydrogen bonds. The crystal structure features N—H⋯S, C—H⋯S and C—H⋯N hydrogen bonds together with C—H⋯π contacts that stack the complexes along the b-axis direction. The packing was further explored by Hirshfeld surface analysis. The thermal properties of the complex were also investigated by simultaneous TGA–DTA analyses.




o&

Crystal structures of {1,1,1-tris­[(salicylaldimino)­meth­yl]ethane}­gallium as both a pyridine solvate and an aceto­nitrile 0.75-solvate and {1,1,1-tris[(salicylaldimino)­meth­yl]ethane}­indium di­chloro­

The sexa­dentate ligand 1,1,1-tris­[(salicyl­idene­amino)­meth­yl]ethane has been reported numerous times in its triply deprotonated form coordinated to transition metals and lanthanides, yet it has been rarely employed with main-group elements, including in substituted forms. Its structures with gallium and indium are reported as solvates, namely, ({[(2,2-bis­{[(2-oxido­benzyl­idene)amino-κ2N,O]meth­yl}prop­yl)imino]­meth­yl}phenololato-κ2N,O)gallium(III) pyridine monosolvate, [Ga(C26H24N3O3)]·C5H5N, the aceto­nitrile 0.75-solvate, [Ga(C26H24N3O3)]·0.75C2H3N, and ({[(2,2-bis­{[(2-oxido­benzyl­idene)amino-κ2N,O]meth­yl}prop­yl)imino]­meth­yl}phenololato-κ2N,O)indium(III) di­chloro­methane monosolvate, [In(C26H24N3O3)]·CH2Cl2. All three metal complexes are pseudo-octa­hedral and each structure contains multiple weak C—H⋯O and/or C—H⋯N inter­molecular hydrogen-bonding inter­actions. The syntheses and additional characterization in the forms of melting points, high-resolution mass spectra, infra-red (IR) spectra, and 1H and 13C NMR spectra are also reported.




o&

Crystal structure and Hirshfeld surface analysis of hexyl 1-hexyl-2-oxo-1,2-di­hydro­quinoline-4-carboxyl­ate

The asymmetric unit of the title compound, C22H31NO3, comprises of one mol­ecule. The mol­ecule is not planar, with the carboxyl­ate ester group inclined by 33.47 (4)° to the heterocyclic ring. Individual mol­ecules are linked by aromaticC—H⋯Ocarbon­yl hydrogen bonds into chains running parallel to [001]. Slipped π–π stacking inter­actions between quinoline moieties link these chains into layers extending parallel to (100). Hirshfeld surface analysis, two-dimensional fingerprint plots and mol­ecular electrostatic potential surfaces were used to qu­antify the inter­molecular inter­actions present in the crystal, indicating that the most important contributions for the crystal packing are from H⋯H (72%), O⋯H/H⋯O (14.5%) and C⋯H/H⋯C (5.6%) inter­actions.




o&

Synthesis, crystal structure, DFT calculations and Hirshfeld surface analysis of 3-butyl-2,6-bis­(4-fluoro­phen­yl)piperidin-4-one

The title compound, C21H23F2NO, consists of two fluoro­phenyl groups and one butyl group equatorially oriented on a piperidine ring, which adopts a chair conformation. The dihedral angle between the mean planes of the phenyl rings is 72.1 (1)°. In the crystal, N—H⋯O and weak C—H⋯F inter­actions, which form R22[14] motifs, link the mol­ecules into infinite C(6) chains propagating along [001]. A weak C—H⋯π inter­action is also observed. A Hirshfeld surface analysis of the crystal structure indicates that the most significant contributions to the crystal packing are from H⋯H (53.3%), H⋯C/C⋯H (19.1%), H⋯F/F⋯H (15.7%) and H⋯O/O⋯H (7.7%) contacts. Density functional theory geometry-optimized calculations were compared to the experimentally determined structure in the solid state and used to determine the HOMO–LUMO energy gap and compare it to the UV–vis experimental spectrum.




o&

Crystal structure of trans-di­chlorido­(1,4,8,11-tetra­aza­cyclo­tetra­decane-κ4N)chromium(III) bis­(form­amide-κO)(1,4,8,11-tetra­aza­cyclo­tetra­decane-κ4N)chromium(III) bis­[tetra­ch

The structure of the title compound, [CrCl2(C10H24N4)][Cr(HCONH2)2(C10H24N4)][ZnCl4]2 (C10H24N4 = 1,4,8,11-tetra­aza­cyclo­tetra­decane, cyclam; HCONH2 = formamide, fa), has been determined from synchrotron X-ray data. The asymmetric unit contains two independent halves of the [CrCl2(cyclam)]+ and [Cr(fa)(cyclam)]3+ cations, and one tetra­chlorido­zincate anion. In each complex cation, the CrIII ion is coordinated by the four N atoms of the cyclam ligand in the equatorial plane and two Cl ligands or two O-bonded formamide mol­ecules in a trans axial arrangement, displaying a distorted octa­hedral geometry with crystallographic inversion symmetry. The Cr—N(cyclam) bond lengths are in the range 2.061 (2) to 2.074 (2) Å, while the Cr—Cl and Cr—O(fa) bond distances are 2.3194 (7) and 1.9953 (19) Å, respectively. The macrocyclic cyclam moieties adopt the centrosymmetric trans-III conformation with six- and five-membered chelate rings in chair and gauche conformations. The crystal structure is stabilized by inter­molecular hydrogen bonds involving the NH or CH groups of cyclam and the NH2 group of coordinated formamide as donors, and Cl atoms of the ZnCl42− anion as acceptors.




o&

Crystal structure of bis­(1-mesityl-1H-imidazole-κN3)di­phenyl­boron tri­fluoro­methane­sulfonate

The solid-state structure of bis­(1-mesityl-1H-imidazole-κN3)di­phenyl­boron tri­fluoro­methane­sulfonate, C36H38BN4+·CF3SO3− or (Ph2B(MesIm)2OTf), is reported. Bis(1-mesityl-1H-imidazole-κN3)di­phenyl­boron (Ph2B(MesIm)2+) is a bulky ligand that crystallizes in the ortho­rhom­bic space group Pbcn. The asymmetric unit contains one Ph2B(MesIm)2+ cationic ligand and one tri­fluoro­methane­sulfonate anion that balances the positive charge of the ligand. The tetra­hedral geometry around the boron center is distorted as a result of the steric bulk of the phenyl groups. Weak inter­actions, such as π–π stacking are present in the crystal structure.




o&

Cis versus trans arrangement of di­thio­carbazate ligands in bis-chelated Ni and Cu complexes

The structures are described of two bis-chelated metal complexes of nickel(II) and copper(II) with S-n-hexyl 3-(1-phenyl­ethyl­idene)di­thio­carbazate Schiff bases in a cis configuration, namely, bis­[S-n-hexyl 3-(1-phenyl­ethyl­idene)di­thio­carbazato-κ2N3,S]nickel(II), [Ni(C15H21N2S2)2], and bis­[S-n-hexyl 3-(1-phenyl­ethyl­idene)di­thio­carbazato-κ2N3,S]copper(II), [Cu(C15H21N2S2)2]. In both complexes, the metals have distorted square-planar geometries. A search in the Cambridge Structural Database [Groom et al. (2016). Acta Cryst. B72, 171–179] for bis-chelated nickel(II) and copper(II) complexes with similar Schiff bases retrieved 55 and 36 hits for the two metals, respectively. An analysis of the geometrical parameters of complexes showing cis and trans configurations is reported and the values compared with those for the complexes described in this work.




o&

Synthesis and structure of ethyl 2-[(4-oxo-3-phenyl-3,4-di­hydro­quinazolin-2-yl)sulfan­yl]acetate

The title compound, C18H16N2O3S, was synthesized by reaction of 2-mercapto-3-phenyl­quinazolin-4(3H)-one with ethyl chloro­acetate. The quinazoline ring forms a dihedral angle of 86.83 (5)° with the phenyl ring. The terminal methyl group is disordered by a rotation of about 60° in a 0.531 (13): 0.469 (13) ratio. In the crystal, C—H⋯O hydrogen-bonding inter­actions result in the formation of columns running in the [010] direction. Two parallel columns further inter­act by C—H⋯O hydrogen bonds. The most important contributions to the surface contacts are from H⋯H (48.4%), C⋯H/H⋯C (21.5%) and O⋯H/H⋯O (18.7%) inter­actions, as concluded from a Hirshfeld analysis.




o&

Crystal structure, Hirshfeld surface analysis and DFT studies of 6-bromo-3-(12-bromo­dodec­yl)-2-(4-nitro­phen­yl)-4H-imidazo[4,5-b]pyridine

The title compound, C24H30Br2N4O2, consists of a 2-(4-nitro­phen­yl)-4H-imidazo[4,5-b]pyridine entity with a 12-bromo­dodecyl substituent attached to the pyridine N atom. The middle eight-carbon portion of the side chain is planar to within 0.09 (1) Å and makes a dihedral angle of 21.9 (8)° with the mean plane of the imidazolo­pyridine moiety, giving the mol­ecule a V-shape. In the crystal, the imidazolo­pyridine units are associated through slipped π–π stacking inter­actions together with weak C—HPyr⋯ONtr and C—HBrmdc­yl⋯ONtr (Pyr = pyridine, Ntr = nitro and Brmdcyl = bromo­dodec­yl) hydrogen bonds. The 12-bromo­dodecyl chains overlap with each other between the stacks. The terminal –CH2Br group of the side chain shows disorder over two resolved sites in a 0.902 (3):0.098 (3) ratio. Hirshfeld surface analysis indicates that the most important contributions for the crystal packing are from H⋯H (48.1%), H⋯Br/Br⋯H (15.0%) and H⋯O/O⋯H (12.8%) inter­actions. The optimized mol­ecular structure, using density functional theory at the B3LYP/ 6–311 G(d,p) level, is compared with the experimentally determined structure in the solid state. The HOMO–LUMO behaviour was elucidated to determine the energy gap.




o&

Structure of a push–pull olefin prepared by ynamine hydro­boration with a borandiol ester

N-[(Z)-2-(2H-1,3,2-Benzodioxaborol-2-yl)-2-phenyl­ethen­yl]-N-(propan-2-yl)aniline, C23H22BNO2, contains a C=C bond that is conjugated with a donor and an acceptor group. An analysis that included similar push–pull olefins revealed that bond lengths in their B—C=C—N core units correlate with the perceived acceptor and donor strength of the groups. The two phenyl groups in the mol­ecule are rotated with respect to the plane that contains the BCCN atoms, and are close enough for significant π-stacking. Definite characterization of the title compound demonstrates, for the first time in a reliable way, that hydro­boration of ynamines with borandiol esters is feasible. Compared to olefin hydro­boration with borane, the ynamine substrate is activated enough to undergo reaction with the less active hydro­boration reagent catecholborane.




o&

Crystal structure and DFT computational studies of (E)-2,4-di-tert-butyl-6-{[3-(tri­fluoro­meth­yl)benz­yl]imino­meth­yl}phenol

The title compound, C23H28F3NO, is an ortho-hy­droxy Schiff base compound, which adopts the enol–imine tautomeric form in the solid state. The mol­ecular structure is not planar and the dihedral angle between the planes of the aromatic rings is 85.52 (10)°. The tri­fluoro­methyl group shows rotational disorder over two sites, with occupancies of 0.798 (6) and 0.202 (6). An intra­molecular O—H⋯N hydrogen bonding generates an S(6) ring motif. The crystal structure is consolidated by C—H⋯π inter­actions. The mol­ecular structure was optimized via density functional theory (DFT) methods with the B3LYP functional and LanL2DZ basis set. The theoretical structure is in good agreement with the experimental data. The frontier orbitals and mol­ecular electrostatic potential map were also examined by DFT computations.




o&

Crystal structure and Hirshfeld surface analysis of 6-benzoyl-3,5-di­phenyl­cyclo­hex-2-en-1-one

In the title compound, C25H20O2, the central cyclo­hexenone ring adopts an envelope conformation. The mean plane of the cyclo­hexenone ring makes dihedral angles of 87.66 (11) and 23.76 (12)°, respectively, with the two attached phenyl rings, while it is inclined by 69.55 (11)° to the phenyl ring of the benzoyl group. In the crystal, the mol­ecules are linked by C—H⋯O and C—H⋯π inter­actions, forming a three-dimensional network.




o&

Hydrogen-bonding patterns in 2,2-bis­(4-methyl­phen­yl)hexa­fluoro­propane pyridinium and ethyl­enedi­ammonium salt crystals

The crystal structures of two salt crystals of 2,2-bis­(4-methyl­phen­yl)hexa­fluoro­propane (Bmphfp) with amines, namely, dipyridinium 4,4'-(1,1,1,3,3,3-hexa­fluoro­propane-2,2-di­yl)dibenzoate 4,4'-(1,1,1,3,3,3-hexa­fluoro­propane-2,2-di­yl)di­benzoic acid, 2C5H6N+·C17H8F6O42−·C17H10F6O4, (1), and a monohydrated ethyl­enedi­ammonium salt ethane-1,2-diaminium 4,4'-(1,1,1,3,3,3-hexa­fluoro­propane-2,2-di­yl)dibenzoate monohydrate, C2H10N22+·C17H8F6O42−·H2O, (2), are reported. Compounds 1 and 2 crystallize, respectively, in space group P21/c with Z' = 2 and in space group Pbca with Z' = 1. The crystals of compound 1 contain neutral and anionic Bmphfp mol­ecules, and form a one-dimensional hydrogen-bonded chain motif. The crystals of compound 2 contain anionic Bmphfp mol­ecules, which form a complex three-dimensional hydrogen-bonded network with the ethyl­enedi­amine and water mol­ecules.




o&

Synthesis and crystal structure of a penta­copper(II) 12-metallacrown-4: cis-di­aqua­tetra­kis­(di­methyl­formamide-κO)manganese(II) tetra­kis­(μ3-N,2-dioxido­benzene-1-carboximidate)penta­copper(II)

The title compound, [Mn(C3H7NO)4(H2O)2][Cu5(C7H4NO3)4]·C3H7NO or cis-[Mn(H2O)2(DMF)4]{Cu[12-MCCu(II)N(shi)-4]}·DMF, where MC is metallacrown, shi3− is salicyl­hydroximate, and DMF is N,N-di­methyl­formamide, crystallizes in the monoclinic space group P21/n. Two crystallographically independent metallacrown anions are present in the structure, and both anions exhibit minor main mol­ecule disorder by an approximate (non-crystallographic) 180° rotation with occupancy ratios of 0.9010 (9) to 0.0990 (9) for one anion and 0.9497 (8) to 0.0503 (8) for the other. Each penta­copper(II) metallacrown contains four CuII ions in the MC ring and a CuII ion captured in the central cavity. Each CuII ion is four-coordinate with a square-planar geometry. The anionic {Cu[12-MCCu(II)N(shi)-4]}2− is charged-balanced by the presence of a cis-[Mn(H2O)2(DMF)4]2+ cation located in the lattice. In addition, the octa­hedral MnII counter-cation is hydrogen bonded to both MC anions via the coordinated water mol­ecules of the MnII ion. The water mol­ecules form hydrogen bonds with the phenolate and carbonyl oxygen atoms of the shi3− ligands of the MCs.




o&

Synthesis, crystal structure and Hirshfeld surface analysis of N-(4-chloro­phen­yl)-5-cyclo­propyl-1-(4-meth­oxy­phen­yl)-1H-1,2,3-triazole-4-carboxamide

The title compound, C19H17ClN4O2, was obtained via a two-step synthesis involving the enol-mediated click Dimroth reaction of 4-azido­anisole with methyl 3-cyclo­propyl-3-oxo­propano­ate leading to the 5-cyclo­propyl-1-(4-meth­oxy­phen­yl)-1H-1,2,3-triazole-4-carb­oxy­lic acid and subsequent acid amidation with 4-chloro­aniline by 1,1'-carbonyl­diimidazole (CDI). It crystallizes in space group P21/n, with one mol­ecule in the asymmetric unit. In the extended structure, two mol­ecules arranged in a near coplanar fashion relative to the triazole ring planes are inter­connected by N—H⋯N and C—H⋯N hydrogen bonds into a homodimer. The formation of dimers is a consequence of the above inter­action and the edge-to-face stacking of aromatic rings, which are turned by 58.0 (3)° relative to each other. The dimers are linked by C—H⋯O inter­actions into ribbons. DFT calculations demonstrate that the frontier mol­ecular orbitals are well separated in energy and the HOMO is largely localized on the 4-chloro­phenyl amide motif while the LUMO is associated with aryl­triazole grouping. A Hirshfeld surface analysis was performed to further analyse the inter­molecular inter­actions.




o&

Structural changes during water-mediated amorphization of semiconducting two-dimensional thio­stannates

Owing to their combined open-framework structures and semiconducting properties, two-dimensional thio­stannates show great potential for catalytic and sensing applications. One such class of crystalline materials consists of porous polymeric [Sn3S72−]n sheets with molecular cations embedded in-between. The compounds are denoted R-SnS-1, where R is the cation. Dependent on the cation, some R-SnS-1 thio­stannates transition into amorphous phases upon dispersion in water. Knowledge about the fundamental chemical properties of the thio­stannates, including their water stability and the nature of the amorphous products, has not yet been established. This paper presents a time-resolved study of the transition from the crystalline to the amorphous phase of two violet-light absorbing thio­stannates, i.e. AEPz-SnS-1 [AEPz = 1-(2-amino­ethyl)­piperazine] and trenH-SnS-1 [tren = tris­(2-amino­ethyl)­amine]. X-ray total scattering data and pair distribution function analysis reveal no change in the local intralayer coordination during the amorphization. However, a rapid decrease in the crystalline domain sizes upon suspension in water is demonstrated. Although scanning electron microscopy shows no significant decrease of the micrometre-sized particles, transmission electron microscopy reveals the formation of small particles (∼200–400 nm) in addition to the larger particles. The amorphization is associated with disorder of the thio­stannate nanosheet stacking. For example, an average decrease in the interlayer distance (from 19.0 to 15.6 Å) is connected to the substantial loss of the organic components as shown by elemental analysis and X-ray photoelectron spectroscopy. Despite the structural changes, the light absorption properties of the amorphisized R-SnS-1 compounds remain intact, which is encouraging for future water-based applications of such materials.




o&

Three differently coloured polymorphs of 3,6-bis­(4-chloro­phenyl)-2,5-di­propyl-2,5-di­hydro­pyrrolo­[3,4-c]pyrrole-1,4-dione

In this paper, the conformational polymorphism of a chlorinated diketo­pyrrolo­pyrrole (DPP) dye having flexible substituents in a non-hydrogen-bonding system is reported. The propyl-substituted DPP derivative (PR3C) has three polymorphic forms, each showing a different colour (red, orange and yellow). All polymorphs could be obtained concomitantly under various crystallization conditions. The results of the crystal structure analysis indicate that PR3C adopts different conformations in each polymorph. The packing effect caused by the difference in the arrangement of neighbouring molecules was found to play an important role in the occurrence of the observed polymorphism. The thermodynamic stability relationship between the three polymorphs was identified by thermal analysis and indicates that the yellow polymorph is the thermally stable form. The results indicate that the yellow form and orange form are enantiotropically related, and the other polymorph is monotropically related to the others.




o&

Insight into the role of pre-assembly and desolvation in crystal nucleation: a case of p-nitro­benzoic acid

As one of the most important phenomena in crystallization, the crystal nucleation process has always been the focus of research. In this work, influences of pre-assembly species and the desolvation process on the crystal nucleation process were studied. p-Nitro­benzoic acid (PNBA) was taken as a model compound to investigate the relationship between solution chemistry and nucleation kinetics in seven different solvents. One unsolvated form and four solvates of PNBA were obtained and one of the solvates was newly discovered. The nucleation behaviours and nucleation kinetics of PNBA in the seven solvents were studied and analyzed. Density functional theory (DFT) and solvation energy calculation were adopted to evaluate the strength of solute–solvent interactions. Vibrational spectroscopy combined with molecular simulation was applied to reveal the pre-assembly species in the solution. Based on these results, a comprehensive understanding of the relationship between molecular structure, crystal structure, solution chemistry and nucleation dynamics was proposed and discussed. It was found that the structural similarity between solution chemistry and crystal structure, the interaction between specific sites and the overall strength of solvation will jointly affect the nucleation process.




o&

Orientational disorder of mono­methyl-quinacridone investigated by Rietveld refinement, structure refinement to the pair-distribution function and lattice-energy minimizations

The crystal structure of the nanocrystalline pigment mono­methyl-quinacridone was solved from X-ray powder data. The orientational disorder was investigated using Rietveld refinements, structure refinement to the pair-distribution function, and lattice-energy minimizations of various ordered structural models.




o&

Crystal structure and Hirshfeld surface analysis of 4,4'-(propane-1,3-diyl)bis(4H-1,2,4-triazol-1-ium) penta­fluorido­oxidovanadate(V)

In the structure of the title salt, second-order Jahn–Teller distortion of the coordination octa­hedra around V ions is reflected by coexistence of short V—O bonds and trans-positioned long V—F bonds, with four equatorial V—F distances being inter­mediate in magnitude. Hydrogen bonding of the anions is restricted to F-atom acceptors only, with particularly strong N–H⋯F inter­actions [N⋯F = 2.5072 (15) Å] established by axial and cis-positioned equatorial F atoms.




o&

Synthesis and crystal structure of (1,10-phenanthroline-κ2N,N')[2-(1H-pyrazol-1-yl)phenyl-κ2N2,C1]iridium(III) hexa­fluorido­phosphate with an unknown number of solvent mol­ecules

The cationic cyclo­metallated iridium(III) complex [Ir(C9H7N2)2(C12H8N2)](PF6) has been synthesized and crystallized by the inter-diffusion method. It contains an unknown number of solvent mol­ecules and has a different space-group symmetry (C2/c) structure than its solvatomorph (P21/c).




o&

Crystal structure and Hirshfeld surface analysis of 4-{2,2-di­chloro-1-[(E)-(4-fluoro­phen­yl)diazen­yl]ethen­yl}-N,N-di­methyl­aniline

The dihedral angle between the two aromatic rings of the title compound is 64.12 (14)°. The crystal structure is stabilized by a short Cl⋯H contact, C—Cl⋯π and van der Waals inter­actions.




o&

Co-crystal structure, Hirshfeld surface analysis and DFT studies of 3,4-ethyl­ene­dioxy­thio­phene solvated bis­[1,3-bis­(penta­fluoro­phen­yl)propane-1,3-dionato]copper(II)

The title complex, Cu(L)2 or [Cu(C15HF10O2)2], comprising one copper ion and two fully fluorinated ligands (L−), was crystallized with 3,4-ethyl­ene­dioxy­thio­phene (EDOT, C6H6O2S) as a guest mol­ecule to give in a di­chloro­methane solution a unique co-crystal, Cu(L)2·3C6H6O2S.




o&

Who's Who in Payments 2020 – Complete Overview of Key Payment Providers

Who’s Who in Payments – Complete Overview of Key Payment Providers encompasses key trends in the payments industry.




o&

Absolute structure of the chiral pyrrolidine derivative (2S)-methyl (Z)-5-(2-tert-but­oxy-1-cyano-2-oxo­ethyl­idene)pyrrolidine-2-carboxyl­ate, a com­pound with low resonant scattering

The enanti­opure monopyrrolidine derivative (2S)-methyl (Z)-5-(2-tert-but­oxy-1-cyano-2-oxo­ethyl­idene)pyrrolidine-2-carboxyl­ate, C13H18N2O4, (1), represents a potential ligand and an attractive inter­mediate for the synthesis of chiral metal com­plexes. At the mol­ecular level, the com­pound features an intra­molecular N—H⋯O hydrogen bond; neighbouring mol­ecules inter­act via N—H⋯N contacts to form chains along [100]. Due to its elemental com­position, resonant scattering of the target com­pound is entirely insignificant for diffraction experiments with Mo Kα and small even for Cu Kα radiation. A preliminary study with the harder radiation type confirmed the chiral space group and the suitability of the single crystal chosen; as expected, the results concerning the absolute structure remained com­pletely inconclusive. A second data collection with the longer wavelength gave satisfactory quality indicators for the correct handedness of the mol­ecule, albeit with high standard uncertainties. The absolute configuration has been assessed independently: CD spectra for both enanti­omers of the target mol­ecule were calculated and the spectrum for the S-configured stereoisomer was in agreement with the experiment. The Cotton effect of (1) may be ascribed to π–π* transitions from HOMO to LUMO and from HOMO to LUMO+1. As both independent techniques agree with respect to the handedness of the target mol­ecule, the absolute structure may be assigned with a high degree of confidence.




o&

Conversion of 3-amino-4-aryl­amino-1H-iso­chromen-1-ones to 1-aryl­iso­chromeno[3,4-d][1,2,3]triazol-5(1H)-ones: synthesis, spectroscopic characterization and the structures of four products and one ring-opened derivative

An efficient synthesis of 1-aryl­isochromeno[3,4-d][1,2,3]triazol-5(1H)-ones, involving the diazo­tization of 3-amino-4-aryl­amino-1H-isochromen-1-ones in weakly acidic solution, has been developed and the spectroscopic characterization and crystal structures of four examples are reported. The mol­ecules of 1-phenyl­isochromeno[3,4-d][1,2,3]triazol-5(1H)-one, C15H9N3O2, (I), are linked into sheets by a combination of C—H⋯N and C—H⋯O hydrogen bonds, while the structures of 1-(2-methyl­phen­yl)isochromeno[3,4-d][1,2,3]triazol-5(1H)-one, C16H11N3O2, (II), and 1-(3-chloro­phen­yl)isochromeno[3,4-d][1,2,3]triazol-5(1H)-one, C15H8ClN3O2, (III), each contain just one hydrogen bond which links the mol­ecules into simple chains, which are further linked into sheets by π-stacking inter­actions in (II) but not in (III). In the structure of 1-(4-chloro­phen­yl)isochromeno[3,4-d][1,2,3]triazol-5(1H)-one, (IV), isomeric with (III), a combination of C—H⋯O and C—H⋯π(arene) hydrogen bonds links the mol­ecules into sheets. When com­pound (II) was exposed to a strong acid in methanol, qu­anti­tative conversion occurred to give the ring-opened transesterification product methyl 2-[4-hy­droxy-1-(2-methyl­phen­yl)-1H-1,2,3-triazol-5-yl]benzoate, C17H15N3O3, (V), where the mol­ecules are linked by paired O—H⋯O hydrogen bonds to form centrosymmetric dimers.




o&

Synthesis of N-substituted 3-(2-aryl-2-oxoeth­yl)-3-hy­droxy­indolin-2-ones and their conversion to N-substituted (E)-3-(2-aryl-2-oxo­ethyl­idene)indolin-2-ones: synthetic sequence, spectroscopic characterization and structures of

An operationally simple and time-efficient approach has been developed for the synthesis of racemic N-substituted 3-(2-aryl-2-oxoeth­yl)-3-hy­droxy­indolin-2-ones by a piperidine-catalysed aldol reaction between aryl methyl ketones and N-alkyl­isatins. These aldol products were used successfully as strategic inter­mediates for the preparation of N-substituted (E)-3-(2-hetaryl-2-oxo­ethyl­idene)indolin-2-ones by a stereoselective dehydration reaction under acidic conditions. The products have all been fully characterized by 1H and 13C NMR spectroscopy, by mass spectrometry and, for a representative selection, by crystal structure analysis. In each of (RS)-1-benzyl-3-hy­droxy-3-[2-(4-meth­oxy­phen­yl)-2-oxoeth­yl]indolin-2-one, C24H21NO4, (Ic), and (RS)-1-benzyl-3-{2-[4-(di­methyl­amino)­phen­yl]-2-oxoeth­yl}-3-hy­droxy­indolin-2-one, C25H24N2O3, (Id), inversion-related pairs of mol­ecules are linked by O—H⋯O hydrogen bonds to form R22(10) rings, which are further linked into chains of rings by a combination of C—H⋯O and C—H⋯π(arene) hydrogen bonds in (Ic) and by C—H⋯π(arene) hydrogen bonds in (Id). The mol­ecules of (RS)-1-benzyl-3-hy­droxy-3-[2-oxo-2-(pyridin-4-yl)eth­yl]indolin-2-one, C22H18N2O3, (Ie), are linked into a three-dimensional framework structure by a combination of O—H⋯N, C—H⋯O and C—H⋯π(arene) hydrogen bonds. (RS)-3-[2-(Benzo[d][1,3]dioxol-5-yl)-2-oxoeth­yl]-1-benzyl-3-hy­droxy­indolin-2-one, C24H19NO5, (If), crystallizes with Z' = 2 in the space group Poverline{1} and the mol­ecules are linked into com­plex sheets by a combination of O—H⋯O, C—H⋯O and C—H⋯π(arene) hydro­gen bonds. In each of (E)-1-benzyl-3-[2-(4-fluoro­phen­yl)-2-oxo­ethyl­idene]indolin-2-one, C23H16FNO2, (IIa), and (E)-1-benzyl-3-[2-oxo-2-(thiophen-2-yl)ethylidene]indolin-2-one, C21H15NO2S, (IIg), the mol­ecules are linked into simple chains by a single C—H⋯O hydrogen bond, while those of (E)-1-benzyl-3-[2-oxo-2-(pyridin-4-yl)ethyl­idene]indolin-2-one, C22H16N2O2, (IIe), are linked by three C—H⋯O hydrogen bonds to form sheets which are further linked into a three-dimensional structure by C—H⋯π(arene) hydrogen bonds. There are no hydrogen bonds in the structures of either (E)-1-benzyl-3-[2-(4-meth­oxy­phen­yl)-2-oxo­ethyl­idene]indolin-2-one, C24H19NO3, (IIc), or (E)-1-benzyl-5-chloro-3-[2-(4-chloro­phen­yl)-2-oxo­ethyl­idene]indolin-2-one, C23H15Cl2NO2, (IIh), but the mol­ecules of (IIh) are linked into chains of π-stacked dimers by a combination of C—Cl⋯π(arene) and aromatic π–π stacking inter­actions.




o&

Synthesis, decom­position studies and crystal structure of a three-dimensional CuCN network structure with protonated N-methyl­ethano­lamine as the guest cation

The com­pound poly[2-hy­droxy-N-methyl­ethan-1-aminium [μ3-cyanido-κ3C:C:N-di-μ-cyanido-κ4C:N-dicuprate(I)]], {(C3H10NO)[Cu2(CN)3]}n or [meoenH]Cu2(CN)3, crystallizes in the tetra­gonal space group P43. The structure consists of a three-dimensional (3D) anionic CuICN network with noncoordinated protonated N-methyl­ethano­lamine cations providing charge neutrality. Pairs of cuprophilic Cu atoms are bridged by the C atoms of μ3-cyanide ligands, which link these units into a 43 spiral along the c axis. The spirals are linked together into a 3D anionic network by the two other cyanide groups. The cationic moieties are linked into their own 43 spiral via N—H⋯O and O—H⋯O hydrogen bonds, and the cations inter­act with the 3D network via an unusual pair of N—H⋯N hydrogen bonds to one of the μ2-cyanide groups. Thermogravimetric analysis indicates an initial loss of the base cation and one cyanide as HCN at temperatures in the range 130–250 °C to form CuCN. We show how loss of a specific cyanide group from the 3D CuCN structure could form the linear CuCN structure. Further heating leaves a residue of elemental copper, isolated as the oxide.




o&

Polymorphism and phase transformation in the dimethyl sulfoxide solvate of 2,3,5,6-tetra­fluoro-1,4-di­iodo­benzene

A new polymorph (form II) is reported for the 1:1 dimethyl sulfoxide solvate of 2,3,5,6-tetra­fluoro-1,4-di­iodo­benzene (TFDIB·DMSO or C6F4I2·C2H6SO). The structure is similar to that of a previously reported polymorph (form I) [Britton (2003). Acta Cryst. E59, o1332–o1333], containing layers of TFDIB mol­ecules with DMSO mol­ecules between, accepting I⋯O halogen bonds from two TFDIB mol­ecules. Re-examination of form I over the temperature range 300–120 K shows that it undergoes a phase transformation around 220 K, where the DMSO mol­ecules undergo re-orientation and become ordered. The unit cell expands by ca 0.5 Å along the c axis and contracts by ca 1.0 Å along the a axis, and the space-group symmetry is reduced from Pnma to P212121. Refinement of form I against data collected at 220 K captures the (average) structure of the crystal prior to the phase transformation, with the DMSO mol­ecules showing four distinct disorder com­ponents, corresponding to an overlay of the 297 and 120 K structures. Assessment of the inter­molecular inter­action energies using the PIXEL method indicates that the various orientations of the DMSO mol­ecules have very similar total inter­action energies with the molecules of the TFDIB framework. The phase transformation is driven by inter­actions between DMSO mol­ecules, whereby re-orientation at lower temperature yields significantly closer and more stabilizing inter­actions between neighbouring DMSO mol­ecules, which lock in an ordered arrangement along the shortened a axis.




o&

Anna Mastro's debut 'Walter' epitomizes Palm Springs Film Festival

Andrew J. West stars in Anna Mastro's "Walter"; Credit: "Walter"

R.H. Greene

It's always dicey to characterize a major film festival based on the movies you personally see there, because no matter how diligent you try to be, your impression will always be statistically anecdotal.

I'll see perhaps 10 percent of the films at this year's Palm Springs International Film Festival by the time they roll up the red carpets for the final time, added to the 25 or so I'd watched before I got here, owing to the festival's unique programming policies.

Not bad considering there are 190 movies being screened. So I think I've got the feel of things here. I wouldn't want my doctor to diagnose me based on a test with a 35 to 40 percent chance of accuracy, but I'm not a doctor. Instead of "Do no harm," I quote Spencer Tracy to myself. He said the secret to the creative process is to "just look 'em in the eye and tell 'em the truth."

And the truth is, with the exception of a couple of documentaries and a horror movie, virtually every film I've seen at Palm Springs so far shared some obvious characteristics: the Palm Springs International Film Festival loves it some poignancy and affirmation.

I've already commented on "Match," the Patrick Stewart acting showcase, and "Cowboys," a very funny Croatian comedy with cross-currents of seriousness. I may comment later about "Today," Iran's Oscar submission. (It's terrific by the way, a deeply affecting story about a burnt out cab driver who gets yanked into the world of a battered, unwed mother who steps into his cab.)

(Still from "Today” (Emrooz) by Iranian filmmaker Reza Mirkarimi)

I also saw an Anne Hathaway passion project called "Song One" here. I'm not going to write about it because I'm not in the mood to stomp on somebody else's butterfly. Plus the dramedy "1001 Grams" by the splendiferous-ly named Norwegian Bent Hamer, whose deadpan satire is routinely compared to Jacques Tati.

WATCH the official trailer for "1001 Grams," which includes some foreign languages

At their best, these are all movies that want to move the audience to tears before bouncing a ray of hope off the screen at them. At their worst, these movies are about pain in the same way Novocain is. They acknowledge its reality, in order to neutralize it.

Filmmaker Anna Mastro's debut film "Walter" (one of the Palm Springs premieres) fits what seems to be the festival's programming model, too, and is, I think, a really quite appealing little indie film, with the by now familiar mildly magical realist bent.

It's is a story about grief, though one with a screwball premise so that it doesn't quite present that way at first. Walter (portrayed with charisma and nuance by Andrew J. West) is a 20-something slacker, but a very uptight one, with a soldier's commitment to dress and routine.

He still lives with mom (Virginia Madsen, now shifting toward the character actress portion of her career with ease and grace) and has a job one rung above fast food worker on the ladder of success: He's a ticket taker at the local multiplex.

But what the world surely sees as failure, Walter knows to be his cover for a far more important vocation. Walter's father died when he was just 10 years old; ever since the funeral, Walter has realized something we don't: His real job in life is to decide where people go after they die.

His snap judgments secretly send people to heaven or hell ... until a dead guy from Walter's past shows up and demands that Walter determine his fate, and then all hell breaks loose.

It's an odd premise, bordering on the labored, but Mastro and her extremely appealing cast pull it off, in part by wearing their influences on their sleeves. The fingerprints of Wes Anderson are all over this picture, especially in terms of the way shots are framed and music is used, and I was able to identify the pivotal contribution of "Beasts of the Southern Wild" co-composer Dan Romer by ear, long before I noticed his screen credit.

I suppose that's supposed to be a damning criticism of a first-timer, but I don't see it that way. Tarantino aped Scorsese for years and virtually remade a minor Hong Kong gangster picture when he debuted with "Reservoir Dogs."

Spielberg acknowledges his debt to David Lean. Hitchcock's apprenticeship at Germany's UFA film studio resulted in a lifelong visual and thematic debt to the great Expressionist master Fritz Lang.

The question is, what do you do with your influences, how do you make them your own? And Mastro — who has a real gift for casting, pacing a scene and maneuvering her actors easily between farce and seriousness — has her own talents. She understands how Anderson's visual syntax has become a cinematic shorthand for quirk, and she deploys it to that effect, then tells the story at hand.

There are some issues with that story, though. There's a girl in concessions (Leven Rambin) Walter likes, and there's a bully at work. For all its surface oddity, the mechanical underpinnings of "Walter" frequently feel like they belong in an "American Pie" sequel.

And yet this movie won me over. I liked its faith in the movie palace as a place that still vibrates with the marvelous. I found a dream sequence, where Rambin undresses to camera while sprawled on a rich yellow bed of movie house popcorn hilarious and deeply expressive.

But I think my affection for this picture is mostly centered on Mastro and her cast, which includes a standout performance by Justin Kirk as a very grounded ghost and a broad but successful cameo from William H. Macy as Walter's psychiatrist. They're all groping toward something rather grim and real about loss, while doing their best to serve up some laughs and wonder along the way.

It touched me, because it feels kind of wise.

Off-Ramp contributor R.H. Greene, former editor of Boxoffice Magazine, is in Palm Spring this week to cover the 26th Annual Palm Springs International Film Festival. Look for his missives here, and listen Saturday at noon to Off-Ramp, when he'll interview Chaz Ebert about her late husband Roger Ebert's contributions to the film festival circuit.

This content is from Southern California Public Radio. View the original story at SCPR.org.




o&

Patt's Hats: Disney sells Tonto's headdress from 'The Lone Ranger'

Patt Morrison models a headdress from the movie "The Lone Ranger."; Credit: Michelle Lanz/KPCC

Patt Morrison

Trust me – you’re going to be seeing a lot of these between now and Halloween.

I went to “The Lone Ranger” premiere last month, and outside the theater, Disneyland began selling a version of the Tonto headdress dreamed up by Johnny Depp and his folks for his role in the film, which I found to be a rollicking, ironic version of the classic action adventure with some very sober scenes evoking Native Americans’ tragic history.

The inspiration, Depp says, was artist Kirby Sattler’s interpretive 2006 painting “I Am Crow.”

Depp himself has claimed Native American ancestry, and the bird atop his bean plays a substantial if silent role in the proceedings. It is an interpretive painting, as I said, not a literal rendering of any tribal makeup. In the Sattler painting, the bird is flying above the figure’s head, not perched on it.

But the movie’s invested in storytelling, not the fine points of accuracy. If it had been, it wouldn’t have made the historical solecisms of relocating both Monument Valley and the transcontinental railroad to … Texas.

This content is from Southern California Public Radio. View the original story at SCPR.org.




o&

Very high CO<sub>2</sub> levels decrease yield and antioxidant content of some green vegetables

Increases in the amount of CO<sub>2</sub> in the atmosphere can be beneficial to crops, by providing a source of carbon for growth. However, very high levels of CO<sub>2</sub> have the reverse effect, decreasing the yield and quality of vegetable crops, a new study has shown. The researchers say atmospheric CO<sub>2</sub> concentration should be kept below 5 000 ppm to enhance the yield of leafy vegetables such as cabbage and lettuce.




o&

Composting green waste saves as much CO<sub>2</sub> as energy recovery

Recycling green waste as compost could match the environmental benefits of converting it into renewable energy, in terms of CO2 savings, according to new German research. It suggests that the two forms of waste management should be seen as complementary and both should receive subsidies.





o&

Yahoo&#39;s work-from-home ban sparks recruitment frenzy

Many companies are using Twitter to send out an invitation to Yahoo employees affected by Marissa Mayer's decision to take away remote-working privileges.




o&

Beyonce spends $2,261 on Nando&#39;s chicken. So what?

After a concert in England, the singer treated her crew to chicken wings and much more from a South African chicken chain. Why is this news?




o&

Trying to conceive? The do&#39;s and don&#39;ts to boosting fertility

There are a number of things you can do that will not only boost your fertility but also help ensure that you're in tip-top condition.



  • Babies & Pregnancy