ine trans-Dichloridobis[(S)-(−)-1-(4-methylphenyl)ethylamine-κN]palladium(II) By journals.iucr.org Published On :: 2024-01-12 The title complex, [PdCl2(C9H13N)2], comprises a single molecule in the asymmetric unit. The PdII atom is tetracoordinated by two N atoms from two trans-aligned organic ligands and two Cl ligands, forming a square-planar metal coordination environment. The distances from the ortho-H atoms on the phenyl ring to the central PdII atom fall within the range 4.70–5.30 Å, precluding any significant intramolecular Pd⋯H interactions. Full Article text
ine Dichlorido(4,7-dimethoxy-1,10-phenanthroline-κ2N,N')zinc(II) By journals.iucr.org Published On :: 2024-01-12 In the title complex, [ZnCl2(C14H12N2O2)], the ZnII atom is located on a twofold rotation axis and is fourfold coordinated by two chlorido ligands and a bidentate 4,7-methoxy-1,10-phenanthroline ligand in a distorted tetrahedral environment. Weak π–π stacking interactions between adjacent 4,7-dimethoxy-1,10-phenanthroline rings [centroid-to-centroid distances = 3.5969 (11) and 3.7738 (11) Å] contribute to the alignment of the complexes in layers parallel to (overline{2}01). Full Article text
ine Bis[2,6-bis(benzimidazol-2-yl)pyridine-κ3N,N',N'']nickel(II) bis(trifluoromethanesulfonate) diethyl ether monosolvate By journals.iucr.org Published On :: 2024-01-31 In the title complex, [Ni(C19H13N5)2](CF3SO3)2·(CH3CH2)2O, the central NiII atom is sixfold coordinated by three nitrogen atoms of each 2,6-bis(2-benzimidazolyl)pyridine ligand in a distorted octahedral geometry with two trifluoromethanesulfonate ions and a molecule of diethyl ether completing the outer coordination sphere of the complex. Hydrogen bonding contributes to the organization of the asymmetric units in columns along the a axis generating a porous supramolecular structure. The structure was refined as a two-component twin with a refined BASF value of 0.4104 (13). Full Article text
ine (2,2'-Bipyridine-κ2N,N')(4,4'-dimethoxy-2,2'-bipyridine-κ2N,N')palladium(II) bis(trifluoromethanesulfonate) By journals.iucr.org Published On :: 2024-02-08 In the title complex salt, [Pd(C10H8N2)(C12H12N2O2)](CF3SO3)2, the palladium(II) atom is fourfold coordinated by two chelating ligands, 2,2'-bipyridine and 4,4'-dimethoxy-2,2'-bipyridine, in a distorted square-planar environment. In the crystal, weak π–π stacking interactions between the 2,2'-bipyridine rings [centroid-to-centroid distances = 3.8984 (19) Å] and between the 4,4'-dimethoxy-2,2'-bipyridine rings [centroid-to-centroid distances = 3.747 (18) Å] contribute to the alignment of the complex cations in columns parallel to the b-axis direction. Full Article text
ine Aquabis(2,2'-bipyridine-κ2N,N')(isonicotinamide-κN)ruthenium(II) bis(trifluoromethanesulfonate) By journals.iucr.org Published On :: 2024-02-08 In the title complex, [Ru(C10H8N2)2(C6H6N2O)(H2O)](CF3SO3)2, the central RuII atom is sixfold coordinated by two bidentate 2,2'-bipyridine, an isonicotinamide ligand, and a water molecule in a distorted octahedral environment with trifluoromethanesulfonate ions completing the outer coordination sphere of the complex. Hydrogen bonding involving the water molecule and weak π–π stacking interactions between the pyridyl rings in adjacent molecules contribute to the alignment of the complexes in columns parallel to the c axis. Full Article text
ine 1-Ethyl-3,3-dimethylspiro[indoline-2,8'-phenaleno[1,9-fg]chromene] By journals.iucr.org Published On :: 2024-02-13 The title pyrene-fused spiropyran derivative, C30H25NO, crystallizes with two molecules in the asymmetric unit with dihedral angles between their fused-ring sub units of 76.20 (8) and 89.38 (9)°. In the crystal, weak C—H⋯π interactions link the molecules into a three-dimensional network. Full Article text
ine cis,cis,cis-Dichloridobis(N4,N4-dimethylpyridin-4-amine-κN1)bis(dimethyl sulfoxide-κS)ruthenium(II) By journals.iucr.org Published On :: 2024-03-06 The structure of the title compound, [RuCl2(C7H10N2)2(C2H6OS)2], has monoclinic (P21/n) symmetry. The Ru—N distances of the coordination compound are influenced by the trans chloride or dimethylsulfoxide-κS ligands. The molecular structure exhibits disorder for two of the terminal methyl groups of a dimethyl sulfoxide ligand. Full Article text
ine 10-Bromo-N,N-diphenylanthracen-9-amine By journals.iucr.org Published On :: 2024-03-12 In the title compound, C26H18BrN, the dihedral angles between the anthracene ring system and the phenyl rings are 89.51 (14) and 74.03 (15)°. In the extended structure, a weak C—H⋯Br interaction occurs, which generates [100] chains, but no significant π–π or C—H⋯π interactions are observed. Full Article text
ine 4-Fluorobenzyl (Z)-2-(2-oxoindolin-3-ylidene)hydrazine-1-carbodithioate By journals.iucr.org Published On :: 2024-03-19 The title compound, C16H12FN3OS, a fluorinated dithiocarbazate imine derivative, was synthesized by the one-pot, multi-component condensation reaction of hydrazine hydrate, carbon disulfide, 4-fluorobenzyl chloride and isatin. The compound demonstrates near-planarity across much of the molecule in the solid state and a Z configuration for the azomethine C=N bond. The Z form is further stabilized by the presence of an intramolecular N—H⋯O hydrogen bond. In the extended structure, molecules are linked into dimers by N—H⋯O hydrogen bonds and further connected into chains along either [2overline{1}0] or [100] by weak C—H⋯S and C—H⋯F hydrogen bonds, which further link into corrugated sheets and in combination form the overall three-dimensional network. Full Article text
ine Poly[(μ-2,3-diethyl-7,8-dimethylquinoxaline-κ2N:N)(2,3-diethyl-7,8-dimethylquinoxaline-κN)-μ-nitrato-κ2O:O'-nitrato-κ2O,O'-disilver(I)] By journals.iucr.org Published On :: 2024-03-21 The structure of the title compound, [C14H18N2)2Ag2](NO3)2, contains subtle differences in ligand, metal, and counter-anion coordination. One quinoxaline ligand uses one of its quinoxaline N atoms to bond to one silver cation. That silver cation is bound to a second quinoxaline which, in turn, is bound to a second silver atom; thereby using both of its quinoxaline N atoms. A nitrate group bonds with one of its O atoms to the first silver and uses the same oxygen to bond to a silver atom (related by symmetry to the second), thereby forming an extended network. The second nitrate group on the other silver bonds via two nitrate O atoms; one silver cation therefore has a coordination number of three whereas the second has a coordination number of four. One of the quinoxaline ligands has a disordered ethyl group. Full Article text
ine Bis[2,6-bis(1H-benzimidazol-2-yl)pyridine]ruthenium(II) bis(hexafluoridophosphate) diethyl ether trisolvate By journals.iucr.org Published On :: 2024-03-28 The title compound, [Ru(C19H13N5)2](PF6)2·3C4H10O, was obtained from the reaction of Ru(bimpy)Cl3 [bimpy is 2,6-bis(1H-benzimidazol-2-yl)pyridine] and bimpy in refluxing ethanol followed by recrystallization from diethyl ether/acetonitrile. At 125 K the complex has orthorhombic (Pca21) symmetry. It is remarkable that the structure is almost centrosymmetric. However, refinement in space group Pbcn leads to disorder and definitely worse results. It is of interest with respect to potential catalytic reduction of CO2. The structure displays N—H⋯O, N—H⋯F hydrogen bonding and significant π–π stacking and C—H⋯π stacking interactions. Full Article text
ine 13-Nitrobenzo[a][1,4]benzothiazino[3,2-c]phenoxazine By journals.iucr.org Published On :: 2024-04-26 In the title compound, C22H11N3O3S, dihedral angle between the phenyl rings on the periphery of the molecule is 8.05 (18)°. In the crystal, aromatic π–π stacking distance and short C—H⋯O contacts are observed. The maximum absorption occurs at 688 nm. Full Article text
ine Bis[2,3-bis(thiophen-2-yl)pyrido[3,4-b]pyrazine]silver(I) perchlorate methanol disolvate By journals.iucr.org Published On :: 2024-04-26 The title compound, [Ag(C15H9N3S2)2]ClO4·2CH3OH, is monoclinic. The AgI atom is coordinated by pyrido N atoms and is two-coordinate; however, the AgI atom has nearby O atoms that can be assumed to be weakly bonded – one from the perchlorate anion and one from the methanol solvate molecule. One of the thienyl groups on a 2,3-bis(thiophen-2-yl)pyrido[3,4-b]pyrazine is flipped disordered and was refined to occupancies of 68.4 (6) and 31.6 (6)%. Full Article text
ine mer-Bis(quinoline-2-carboxaldehyde 4-ethylthiosemicarbazonato)nickel(II) methanol 0.33-solvate 0.67-hydrate By journals.iucr.org Published On :: 2024-04-26 In the title compound, [Ni(C13H13N4S)2]·0.33CH3OH·0.67H2O, the NiII atom is coordinated by two tridentate quinoline-2-carboxaldehyde 4-ethylthiosemicarbazonate ligands in a distorted octahedral shape. At 100 K, the crystal symmetry is monoclinic (space group P21/n). A mixture of water and methanol crystallizes with the title complex, and one of the ethyl groups in the coordinating ligands is disordered over two positions, with an occupancy ratio of 58:42. There is intermolecular hydrogen bonding between the solvent molecules and the amine and thiolate groups in the ligands. No other significant interactions are present in the crystal packing. Full Article text
ine Benzo[a][1,4]benzothiazino[3,2-c]phenothiazine By journals.iucr.org Published On :: 2024-04-30 The title compound, C22H12N2S2, crystallizes in space group P21/c with four molecules in the asymmetric unit. The heterocyclic molecule is quasi-planar with a dihedral angle between the phenyl rings on the periphery of the molecule of 1.73 (19)°. Short H⋯S (2.92 Å) and C—H⋯π [2.836 (3) Å] contacts are observed in the crystal with shorted π–π stacking distances of 3.438 (3) Å along the b axis. Surprisingly, and unlike a closely related material, this molecule readily forms large crystals by sublimation and by slow evaporation from dichloromethane. The maximum absorbance in the UV-Vis spectrum is at 533 nm. Emission was measured upon excitation at 533 nm with a fluorescence λmax of 658 nm and cutoff of 900 nm. Full Article text
ine 2-(10-Bromoanthracen-9-yl)-N-phenylaniline By journals.iucr.org Published On :: 2024-05-31 In the title compound, C26H18BrN, the central benzene ring makes dihedral angles with its adjacent anthracene ring system and pendant benzene ring of 87.49 (13) and 62.01 (17)°, respectively. The N—H moiety is sterically blocked from forming a hydrogen bond, but weak C—H⋯π interactions occur in the extended structure. Full Article text
ine Poly[[{μ2-5-[(dimethylamino)(thioxo)methoxy]benzene-1,3-dicarboxylato-κ4O1,O1':O3,O3'}(μ2-4,4'-dipyridylamine-κ2N4:N4')cobalt(II)] dimethylformamide hemisolvate monohydrate] By journals.iucr.org Published On :: 2024-06-04 In the crystal structure of the title compound, {[Co(C11H9NSO5)(C10H9N3)]0.5C3H7NO·H2O}n or {[Co(dmtb)(dpa)]·0.5DMF·H2O}n (dmtb2– = 5-[(dimethylamino)thioxomethoxy]-1,3-benzenedicarboxylate and dpa = 4,4'-dipyridylamine), an assembly of periodic [Co(C11H9NSO5)(C10H9N3)]n layers extending parallel to the bc plane is present. Each layer is constituted by distorted [CoO4N2] octahedra, which are connected through the μ2-coordination modes of both dmtb2– and dpa ligands. Occupationally disordered water and dimethylformamide (DMF) solvent molecules are located in the voids of the network to which they are connected through hydrogen-bonding interactions. Full Article text
ine Chlorido[5,10,15,20-tetrakis(quinoline-7-carboxamido)porphinato]iron(III) By journals.iucr.org Published On :: 2024-06-04 The title compound, [Fe(C84H52N12O4)Cl], crystallizes in space group C2/c. The central FeIII cation (site symmetry 2) is coordinated in a fivefold manner, with four pyrrole N atoms of the porphyrin core in the basal sites and one Cl atom (site symmetry 2) in the apical position, which completes a slightly distorted square-pyramidal environment. The porphyrin macrocycle shows a characteristic ruffled-shape distortion and the iron atom is displaced out of the porphyrin plane by 0.42 Å with the average Fe—N distance being 2.054 (4) Å; the Fe—Cl bond length is 2.2042 (7) Å. Intermolecular C—H⋯N and C—H⋯O hydrogen bonds occur in the crystal structure. Full Article text
ine (2,5-Dimethylimidazole){N,N',N'',N'''-[porphyrin-5,10,15,20-tetrayltetra(2,1-phenylene)]tetrakis(pyridine-3-carboxamide)}manganese(II) chlorobenzene disolvate By journals.iucr.org Published On :: 2024-06-04 In the title compound, [Mn(C68H44N12O4)(C5H8N2)]·2C6H5Cl, the central MnII ion is coordinated by four pyrrole N atoms of the porphyrin core in the basal sites and one N atom of the 2,5-dimethylimidazole ligand in the apical site. Two chlorobenzene solvent molecules are also present in the asymmetric unit. Due to the apical imidazole ligand, the Mn atom is displaced out of the 24-atom porphyrin mean plane by 0.66 Å. The average Mn—Np (p = porphyrin) bond length is 2.143 (8) Å, and the axial Mn—NIm (Im = 2,5-dimethylimidazole) bond length is 2.171 (8) Å. The structure displays intermolecular and intramolecular N—H⋯O, N—H⋯N, C—H⋯O and C—H⋯N hydrogen bonding. The crystal studied was refined as a two-component inversion twin. Full Article text
ine Bis{(S)-(−)-N-[(2-biphenyl)methylidene]-1-(4-methoxyphenyl)ethylamine-κN}dichloridopalladium(II) By journals.iucr.org Published On :: 2024-06-16 The PdII complex bis{(S)-(−)-N-[(biphenyl-2-yl)methylidene]1-(4-methoxyphenyl)ethanamine-κN}dichloridopalladium(II), [PdCl2(C22H21NO)2], crystallizes in the monoclinic Sohncke space group P21 with a single molecule in the asymmetric unit. The coordination environment around the palladium is slightly distorted square planar. The N—Pd—Cl bond angles are 91.85 (19), 88.10 (17), 89.96 (18), and 90.0 (2)°, while the Pd—Cl and Pd—N bond lengths are 2.310 (2) and 2.315 (2) Å and 2.015 (2) and 2.022 (6) Å, respectively. The crystal structure features intermolecular N—H⋯Cl and intramolecular C—H⋯Pd interactions, which lead to the formation of a supramolecular framework structure. Full Article text
ine 2-(Pyridin-4-yl)-2,3-dihydro-1H-naphtho[1,8-de][1,3,2]diazaborinine By journals.iucr.org Published On :: 2024-06-28 The title compound, C15H12BN3, is a type of diazaborinane featuring substitution at 1, 2, and 3 positions in the nitrogen–boron six-membered heterocycle. It is comprised of two almost planar units, the pyridyl ring and the Bdan (dan = 1,8-diaminonaphtho) group, which subtend a dihedral angle of 24.57 (5)°. In the crystal, the molecules are linked into R44(28) hydrogen-bonding networks around the fourfold inversion axis, giving cyclic tetramers. The molecules form columnar stacks along the c axis. Full Article text
ine Bis(ethylenediammonium) μ-ethylenediaminetetraacetato-1κ3O,N,O':2κ3O'',N',O'''-bis[trioxidomolybdate(VI)] tetrahydrate By journals.iucr.org Published On :: 2024-07-12 The title compound, (C2H10N2)2[(C10H12N2O8)(MoO3)2]·4H2O, which crystallizes in the monoclinic C2/c space group, was obtained by mixing molybdenum oxide, ethylenediamine and ethylenediaminetetraacetic acid (H4edta) in a 2:4:1 ratio. The complex anion contains two MoO3 units bridged by an edta4− anion. The midpoint of the central C—C bond of the edta4− anion is located on a crystallographic inversion centre. The independent Mo atom is tridentately coordinated by a nitrogen atom and two carboxylate groups of the edta4− ligand, together with the three oxo ligands, producing a distorted octahedral coordination environment. In the three-dimensional supramolecular crystal structure, the dinuclear anions, the organoammonium counter-ions and the solvent water molecules are linked by N—H⋯Ow, N—H⋯Oedta and O—H⋯O hydrogen bonds. Full Article text
ine Redetermined structure of 4-(benzyloxy)benzoic acid By journals.iucr.org Published On :: 2024-08-06 In the title compound, C14H14O3, the dihedral angle between the aromatic rings is 39.76 (9)°. In the crystal, the molecules associate to form centrosymmetric acid–acid dimers linked by pairwise O—H⋯O hydrogen bonds. The precision of the geometric parameters in the present single-crystal study is about an order of magnitude better than the previous powder diffraction study [Chattopadhyay et al. (2013). CrystEngComm, 15, 1077–1085]. Full Article text
ine Dichloridotetrakis(3-methoxyaniline)nickel(II) By journals.iucr.org Published On :: 2024-08-13 The reaction of nickel(II) chloride with 3-methoxyaniline yielded dichloridotetrakis(3-methoxyaniline)nickel(II), [NiCl2(C7H9NO)4], as yellow crystals. The NiII ion is pseudo-octahedral with the chloride ions trans to each other. The four 3-methoxyaniline ligands differ primarily due to different conformations about the Ni—N bond, which also affect the hydrogen bonding. Intermolecular N—H⋯ Cl hydrogen bonds and short Cl⋯Cl contacts between molecules link them into chains parallel to the b axis. Full Article text
ine Methyl 2-[(Z)-5-bromo-2-oxoindolin-3-ylidene]hydrazinecarbodithioate By journals.iucr.org Published On :: 2024-08-16 The title compound, C10H8BrN3OS2, a brominated dithiocarbazate imine derivative, was obtained from the condensation reaction of S-methyldithiocarbazate (SMDTC) and 5-bromoisatin. The essentially planar molecule exhibits a Z configuration, with the dithiocarbazate and 5-bromoisatin fragments located on the same sides of the C=N azomethine bond, which allows for the formation of an intramolecular N—H⋯Ob (b = bromoisatin) hydrogen bond generating an S(6) ring motif. In the crystal, adjacent molecules are linked by pairs of N—H⋯O hydrogen bonds, forming dimers characterized by an R22(8) loop motif. In the extended structure, molecules are linked into a three-dimensional network by C—H⋯S and C—H⋯Br hydrogen bonds, C—Br⋯S halogen bonds and aromatic π–π stacking. Full Article text
ine catena-Poly[[(8-aminoquinoline)cobalt(II)]-di-μ-azido] By journals.iucr.org Published On :: 2024-09-06 The title coordination polymer, [Co(N3)2(C9H8N2)]n, was synthesized solvothermally. The CoII atom exhibits a distorted octahedral [CoN6] coordination geometry with a bidentate 8-aminoquinoline ligand and four azide ligands. Bridging azide ligands result in chains extending along [100]. N—H⋯N hydrogen bonds join the chains to give an extended structure with sheets parallel to (002). Full Article text
ine Bis[2-(isoquinolin-1-yl)phenyl-κ2N,C1](2-phenyl-1H-imidazo[4,5-f][1,10]phenanthroline-κ2N,N')iridium(III) hexafluoridophosphate methanol monosolvate By journals.iucr.org Published On :: 2024-09-06 The title compound, [Ir(C15H10N)2(C19H12N4)]PF6·CH3OH, crystallizes in the C2/c space group with one monocationic iridium complex, one hexafluoridophosphate anion, and one methanol solvent molecule of crystallization in the asymmetric unit, all in general positions. The anion and solvent are linked to the iridium complex cation via hydrogen bonding. All bond lengths and angles fall into expected ranges compared to similar compounds. Full Article text
ine Redetermined structure of methyl 3-{4,4-difluoro-2-[2-(methoxycarbonyl)ethyl]-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacen-6-yl}propionate By journals.iucr.org Published On :: 2024-09-17 In the title compound, C21H27BF2N2O4, a highly fluorescent boron–dipyrromethene dye, the methylpropionate moieties have different conformations. In the crystal, weak C—H⋯F and C—H⋯O interactions link the molecules. Some optical properties are presented. Full Article text
ine Δ-Bis[(S)-2-(4-isopropyl-4,5-dihydrooxazol-2-yl)phenolato-κ2N,O1](1,10-phenanthroline-κ2N,N')ruthenium(III) hexafluoridophosphate By journals.iucr.org Published On :: 2024-09-17 The title compound, [Ru(C12H14NO2)2(C12H8N2)]PF6 crystallizes in the tetragonal Sohnke space group P41212. The two bidentate chiral salicyloxazoline ligands and the phenanthroline co-ligand coordinate to the central RuIII atom through N,O and N,N atom pairs to form bite angles of 89.76 (15) and 79.0 (2)°, respectively. The octahedral coordination of the bidentate ligands leads to a propeller-like shape, which induces metal-centered chirality onto the complex, with a right-handed (Δ) absolute configuration [the Flack parameter value is −0.003 (14)]. Both the complex cation and the disordered PF6− counter-anion are located on twofold rotation axes. Apart from Coulombic forces, the crystal cohesion is ensured by non-classical C—H⋯O and C—H⋯F interactions. Full Article text
ine Di-μ-adipato-κ4O1,O1':O6,O6'-bis[(2,2'-dipyridylamine-κ2N,N')zinc(II)] trihydrate By journals.iucr.org Published On :: 2024-09-20 The title compound, [Zn2(C6H8O4)2(C10H9N3)2]·3H2O or {Zn2[(C5H4N)2NH]2[μ-(CH2)4(COO)2]2}·3H2O, was separated from the solvothermal reaction of zinc(II) sulfate heptahydrate, 2,2'-dipyridylamine and sodium adipate. The dinuclear metal complex has a centrosymmetric structure, with the ZnII atom adopting a highly distorted octahedral coordination sphere composed of four oxygen atoms from bridging adipato ligands and two pyridine nitrogen atoms. In the crystal, the title compound aggregates into a tri-periodic supramolecular structure through intermolecular hydrogen-bonding networks of the form O—H⋯O and N—H⋯O. Full Article text
ine 1,4-Dimethylpiperazine-2,3-dione By journals.iucr.org Published On :: 2024-10-04 In the title compound, C6H10N2O2, the piperazine-2,3-dione ring adopts a half-chair conformation. In the crystal, the molecules are linked by weak C—H⋯O hydrogen bonds, forming (010) sheets. Full Article text
ine Methyl 2-[(Z)-5-methyl-2-oxoindolin-3-ylidene]hydrazinecarbodithioate By journals.iucr.org Published On :: 2024-10-08 The title dithiocarbazate imine, C11H11N3OS2, was obtained from the condensation reaction of S-methyldithiocarbazate (SMDTC) and 5-methylisatin. It shows a Z configuration about the imine C=N bond, which is associated with an intramolecular N—H⋯O hydrogen bond that closes an S(6) ring. In the crystal, inversion dimers linked by pairwise N—H⋯O hydrogen bonds generate R22(8) loops. The extended structure features C—H⋯S contacts as well as reciprocal carbonyl–carbonyl (C=O⋯C=O) interactions. Full Article text
ine (Z)-N-(2,6-Diisopropylphenyl)-1-[(2-methoxyphenyl)amino]methanimine oxide By journals.iucr.org Published On :: 2024-10-21 The molecular structure of the title compound, C20H26N2O2 reveals non-co-planarity between the central formamidine backbone and each of the outer methoxy- and i-propyl- substituted benzene rings with dihedral angles of 7.88 (15) and 81.17 (15)°, respectively, indicating significant twists in the molecule. In the crystal, intermolecular C—H⋯O interactions, forming an R34(30) graph set, occur within a two-dimensional layer that extends along the ac plane. Full Article text
ine (Z)-N-(2,6-Dimethylphenyl)-1-[(2-methoxyphenyl)amino]methanimine oxide methanol monosolvate By journals.iucr.org Published On :: 2024-10-21 In the title solvate, C16H18N2O2·CH4O, the dihedral angles between the formamidine backbone and the pendant 2-methoxyphenyl and 2,6-dimethylphenyl groups are 14.84 (11) and 81.61 (12)°, respectively. In the crystal, the components are linked by C—H⋯O, O—H⋯O and C—H⋯ π hydrogen bonds, generating a supramolecular chain that extends along the crystallographic a-axis direction. Full Article text
ine (1H-Benzodiazol-2-ylmethyl)diethylamine By journals.iucr.org Published On :: 2024-10-31 In the crystal of the title compound, C12H17N3, the molecules are linked by N—H⋯N hydrogen bonds, generating a C(4) chain extending along the c-axis direction. One of the ethyl groups is disordered over two sets of sites with a refined occupancy ratio of 0.582 (15):0.418 (15). Full Article text
ine Structural insights into 1,4-bis(neopentyloxy)pillar[5]arene and the pyridine host–guest system By journals.iucr.org Published On :: 2024-11-08 The crystal structure of 1,4-bis(neopentyloxy)pillar[5]arene, C95H140N2O10 (TbuP), featuring two encapsulated pyridine molecules, reveals significant host–guest interactions. Interestingly, the pyridine guests are positioned near the neopentyloxy substituents instead of the electron-rich aromatic core of the pillar[5]arene. This spatial arrangement suggests a preference for the pyridine molecules to engage with the aliphatic regions of the host. Detailed analysis of the structural characteristics of this host–guest system (TbuP·2Py), as well as its packing pattern within the crystal network, is presented and discussed. Full Article text
ine α-d-2'-Deoxyadenosine, an irradiation product of canonical DNA and a component of anomeric nucleic acids: crystal structure, packing and Hirshfeld surface analysis By journals.iucr.org Published On :: 2024-01-22 α-d-2'-Deoxyribonucleosides are products of the γ-irradiation of DNA under oxygen-free conditions and are constituents of anomeric DNA. They are not found as natural building blocks of canonical DNA. Reports on their conformational properties are limited. Herein, the single-crystal X-ray structure of α-d-2'-deoxyadenosine (α-dA), C10H13N5O3, and its conformational parameters were determined. In the crystalline state, α-dA forms two conformers in the asymmetric unit which are connected by hydrogen bonds. The sugar moiety of each conformer is arranged in a `clamp'-like fashion with respect to the other conformer, forming hydrogen bonds to its nucleobase and sugar residue. For both conformers, a syn conformation of the nucleobase with respect to the sugar moiety was found. This is contrary to the anti conformation usually preferred by α-nucleosides. The sugar conformation of both conformers is C2'-endo, and the 5'-hydroxyl groups are in a +sc orientation, probably due to the hydrogen bonds formed by the conformers. The formation of the supramolecular assembly of α-dA is controlled by hydrogen bonding and stacking interactions, which was verified by a Hirshfeld and curvedness surface analysis. Chains of hydrogen-bonded nucleobases extend parallel to the b direction and are linked to equivalent chains by hydrogen bonds involving the sugar moieties to form a sheet. A comparison of the solid-state structures of the anomeric 2'-deoxyadenosines revealed significant differences of their conformational parameters. Full Article text
ine Structure and absolute configuration of natural fungal product beauveriolide I, isolated from Cordyceps javanica, determined by 3D electron diffraction By journals.iucr.org Published On :: 2024-02-27 Beauveriolides, including the main beauveriolide I {systematic name: (3R,6S,9S,13S)-9-benzyl-13-[(2S)-hexan-2-yl]-6-methyl-3-(2-methylpropyl)-1-oxa-4,7,10-triazacyclotridecane-2,5,8,11-tetrone, C27H41N3O5}, are a series of cyclodepsipeptides that have shown promising results in the treatment of Alzheimer's disease and in the prevention of foam cell formation in atherosclerosis. Their crystal structure studies have been difficult due to their tiny crystal size and fibre-like morphology, until now. Recent developments in 3D electron diffraction methodology have made it possible to accurately study the crystal structures of submicron crystals by overcoming the problems of beam sensitivity and dynamical scattering. In this study, the absolute structure of beauveriolide I was determined by 3D electron diffraction. The cyclodepsipeptide crystallizes in the space group I2 with lattice parameters a = 40.2744 (4), b = 5.0976 (5), c = 27.698 (4) Å and β = 105.729 (6)°. After dynamical refinement, its absolute structure was determined by comparing the R factors and calculating the z-scores of the two possible enantiomorphs of beauveriolide I. Full Article text
ine Relationship between synthesis method–crystal structure–melting properties in cocrystals: the case of caffeine–citric acid By journals.iucr.org Published On :: 2024-05-07 The influence of the crystal synthesis method on the crystallographic structure of caffeine–citric acid cocrystals was analyzed thanks to the synthesis of a new polymorphic form of the cocrystal. In order to compare the new form to the already known forms, the crystal structure of the new cocrystal (C8H10N4O2·C6H8O7) was solved by powder X-ray diffraction thanks to synchrotron experiments. The structure determination was performed using `GALLOP', a recently developed hybrid approach based on a local optimization with a particle swarm optimizer, particularly powerful when applied to the structure resolution of materials of pharmaceutical interest, compared to classical Monte-Carlo simulated annealing. The final structure was obtained through Rietveld refinement, and first-principles density functional theory (DFT) calculations were used to locate the H atoms. The symmetry is triclinic with the space group Poverline{1} and contains one molecule of caffeine and one molecule of citric acid per asymmetric unit. The crystallographic structure of this cocrystal involves different hydrogen-bond associations compared to the already known structures. The analysis of these hydrogen bonds indicates that the cocrystal obtained here is less stable than the cocrystals already identified in the literature. This analysis is confirmed by the determination of the melting point of this cocrystal, which is lower than that of the previously known cocrystals. Full Article text
ine Molecular structure and selective theophylline complexation by conformational change of diethyl N,N'-(1,3-phenylene)dicarbamate By journals.iucr.org Published On :: 2024-05-07 The receptor ability of diethyl N,N'-(1,3-phenylene)dicarbamate (1) to form host–guest complexes with theophylline (TEO) and caffeine (CAF) by mechanochemistry was evaluated. The formation of the 1–TEO complex (C12H16N2O4·C7H8N4O2) was preferred and involves the conformational change of one of the ethyl carbamate groups of 1 from the endo conformation to the exo conformation to allow the formation of intermolecular interactions. The formation of an N—H⋯O=C hydrogen bond between 1 and TEO triggers the conformational change of 1. CAF molecules are unable to form an N—H⋯O=C hydrogen bond with 1, making the conformational change and, therefore, the formation of the complex impossible. Conformational change and selective binding were monitored by IR spectroscopy, solid-state 13C nuclear magnetic resonance and single-crystal X-ray diffraction. The 1–TEO complex was characterized by IR spectroscopy, solid-state 13C nuclear magnetic resonance, powder X-ray diffraction and single-crystal X-ray diffraction. Full Article text
ine Synthesis, characterization and structural analysis of complexes from 2,2':6',2''-terpyridine derivatives with transition metals By journals.iucr.org Published On :: 2024-05-16 The synthesis and structural characterization of three families of coordination complexes synthesized from 4'-phenyl-2,2':6',2''-terpyridine (8, Ph-TPY), 4'-(4-chlorophenyl)-2,2':6',2''-terpyridine (9, ClPh-TPY) and 4'-(4-methoxyphenyl)-2,2':6',2''-terpyridine (10, MeOPh-TPY) ligands with the divalent metals Co2+, Fe2+, Mn2+ and Ni2+ are reported. The compounds were synthesized from a 1:2 mixture of the metal and ligand, resulting in a series of complexes with the general formula [M(R-TPY)2](ClO4)2 (where M = Co2+, Fe2+, Mn2+ and Ni2+, and R-TPY = Ph-TPY, ClPh-TPY and MeOPh-TPY). The general formula and structural and supramolecular features were determinated by single-crystal X-ray diffraction for bis(4'-phenyl-2,2':6',2''-terpyridine)nickel(II) bis(perchlorate), [Ni(C21H15N3)2](ClO4)2 or [Ni(Ph-TPY)2](ClO4)2, bis[4'-(4-methoxyphenyl)-2,2':6',2''-terpyridine]manganese(II) bis(perchlorate), [Mn(C22H17N3O)2](ClO4)2 or [Mn(MeOPh-TPY)2](ClO4)2, and bis(4'-phenyl-2,2':6',2''-terpyridine)manganese(II) bis(perchlorate), [Mn(C21H15N3)2](ClO4)2 or [Mn(Ph-TPY)2](ClO4)2. In all three cases, the complexes present distorted octahedral coordination polyhedra and the crystal packing is determined mainly by weak C—H⋯π interactions. All the compounds (except for the Ni derivatives, for which FT–IR, UV–Vis and thermal analysis are reported) were fully characterized by spectroscopic (FT–IR, UV–Vis and NMR spectroscopy) and thermal (TGA–DSC, thermogravimetric analysis–differential scanning calorimetry) methods. Full Article text
ine Using cocrystals as a tool to study non-crystallizing molecules: crystal structure, Hirshfeld surface analysis and computational study of the 1:1 cocrystal of (E)-N-(3,4-difluorophenyl)-1-(pyridin-4-yl)methanimine and acetic By journals.iucr.org Published On :: 2024-07-05 Using a 1:1 cocrystal of (E)-N-(3,4-difluorophenyl)-1-(pyridin-4-yl)methanimine with acetic acid, C12H8F2N2·C2H4O2, we investigate the influence of F atoms introduced to the aromatic ring on promoting π–π interactions. The cocrystal crystallizes in the triclinic space group P1. Through crystallographic analysis and computational studies, we reveal the molecular arrangement within this cocrystal, demonstrating the presence of hydrogen bonding between the acetic acid molecule and the pyridyl group, along with π–π interactions between the aromatic rings. Our findings highlight the importance of F atoms in promoting π–π interactions without necessitating full halogenation of the aromatic ring. Full Article text
ine TAAM refinement on high-resolution experimental and simulated 3D ED/MicroED data for organic molecules By journals.iucr.org Published On :: 2024-06-27 3D electron diffraction (3D ED), or microcrystal electron diffraction (MicroED), has become an alternative technique for determining the high-resolution crystal structures of compounds from sub-micron-sized crystals. Here, we considered l-alanine, α-glycine and urea, which are known to form good-quality crystals, and collected high-resolution 3D ED data on our in-house TEM instrument. In this study, we present a comparison of independent atom model (IAM) and transferable aspherical atom model (TAAM) kinematical refinement against experimental and simulated data. TAAM refinement on both experimental and simulated data clearly improves the model fitting statistics (R factors and residual electrostatic potential) compared to IAM refinement. This shows that TAAM better represents the experimental electrostatic potential of organic crystals than IAM. Furthermore, we compared the geometrical parameters and atomic displacement parameters (ADPs) resulting from the experimental refinements with the simulated refinements, with the periodic density functional theory (DFT) calculations and with published X-ray and neutron crystal structures. The TAAM refinements on the 3D ED data did not improve the accuracy of the bond lengths between the non-H atoms. The experimental 3D ED data provided more accurate H-atom positions than the IAM refinements on the X-ray diffraction data. The IAM refinements against 3D ED data had a tendency to lead to slightly longer X—H bond lengths than TAAM, but the difference was statistically insignificant. Atomic displacement parameters were too large by tens of percent for l-alanine and α-glycine. Most probably, other unmodelled effects were causing this behaviour, such as radiation damage or dynamical scattering. Full Article text
ine The influence of the axial group on the crystal structures of boron subphthalocyanines By journals.iucr.org Published On :: 2024-09-04 The crystal structures of 16 boron subphthalocyanines (BsubPcs) with structurally diverse axial groups were analyzed and compared to elucidate the impact of the axial group on the intermolecular π–π interactions, axial-group interactions, axial bond length and BsubPc bowl depth. π–π interactions between the isoindole units of adjacent BsubPc molecules most often involve concave–concave packing, whereas axial-group interactions with adjacent BsubPc molecules tend to favour the convex side of the BsubPc bowl. Furthermore, axial groups that contain O and/or F atoms tend to have significant hydrogen-bonding interactions, while axial groups containing arene site(s) can participate in π–π interactions with the BsubPc bowl, both of which can strongly influence the crystal packing. Bulky axial groups did tend to disrupt the π–π interactions and/or axial-group interactions, preventing some of the close packing that is seen in BsubPcs with less bulky axial groups. The atomic radius of the heteroatom bonded to boron directly influences the axial bond length, whereas the axial group has minimal impact on the BsubPc bowl depth. Finally, the crystal growth method did not generally appear to have a significant impact on the solid-state arrangement, with the exception of water occasionally being incorporated into crystal structures when hygroscopic solvents were used. These insights can help with the design and fine-tuning of the solid-state structures of BsubPcs as they continue to be developed as functional materials in organic electronics. Full Article text
ine Crystal structures of two unexpected products of vicinal diamines left to crystallize in acetone By journals.iucr.org Published On :: 2024-08-23 Herein we report the crystal structures of two benzodiazepines obtained by reacting N,N'-(4,5-diamino-1,2-phenylene)bis(4-methylbenzenesulfonamide) (1) or 4,5-(4-methylbenzenesulfonamido)benzene-1,2-diaminium dichloride (1·2HCl) with acetone, giving 2,2,4-trimethyl-8,9-bis(4-methylbenzenesulfonamido)-2,3-dihydro-5H-1,5-benzodiazepine, C26H30N4O4S2 (2), and 2,2,4-trimethyl-8,9-bis(4-methylbenzenesulfonamido)-2,3-dihydro-5H-1,5-benzodiazepin-1-ium chloride 0.3-hydrate, C26H31N4O4S2+·Cl−·0.3H2O (3). Compounds 2 and 3 were first obtained in attempts to recrystallize 1 and 1·2HCl using acetone as solvent. This solvent reacted with the vicinal diamines present in the molecular structures, forming a 5H-1,5-benzodiazepine ring. In the crystal structure of 2, the seven-membered ring of benzodiazepine adopts a boat-like conformation, while upon protonation, observed in the crystal structure of 3, it adopts an envelope-like conformation. In both crystalline compounds, the tosylamide N atoms are not in resonance with the arene ring, mainly due to hydrogen bonds and steric hindrance caused by the large vicinal groups in the aromatic ring. At a supramolecular level, the crystal structure is maintained by a combination of hydrogen bonds and hydrophobic interactions. In 2, amine-to-tosyl N—H⋯O and amide-to-imine N—H⋯N hydrogen bonds can be observed. In contrast, in 3, the chloride counter-ion and water molecule result in most of the hydrogen bonds being of the amide-to-chloride and ammonium-to-chloride N—H⋯Cl types, while the amine interacts with the tosyl group, as seen in 2. In conclusion, we report the synthesis of 1, 1·2HCl and 2, as well as their chemical characterization. For 2, two synthetic methods are described, i.e. solvent-mediated crystallization and synthesis via a more efficient and cleaner route as a polycrystalline material. Salt 3 was only obtained as presented, with only a few crystals being formed. Full Article text
ine Coordination structure and intermolecular interactions in copper(II) acetate complexes with 1,10-phenanthroline and 2,2'-bipyridine By journals.iucr.org Published On :: 2024-08-23 The crystal structures of two coordination compounds, (acetato-κO)(2,2'-bipyridine-κ2N,N')(1,10-phenanthroline-κ2N,N')copper(II) acetate hexahydrate, [Cu(C2H3O2)(C10H8N2)(C12H8N2)](C2H3O2)·6H2O or [Cu(bipy)(phen)Ac]Ac·6H2O, and (acetato-κO)bis(2,2'-bipyridine-κ2N,N')copper(II) acetate–acetic acid–water (1/1/3), [Cu(C2H3O2)(C10H8N2)2](C2H3O2)·C2H4O2·3H2O or [Cu(bipy)2Ac]Ac·HAc·3H2O, are reported and compared with the previously published structure of [Cu(phen)2Ac]Ac·7H2O (phen is 1,10-phenanthroline, bipy for 2,2'-bipyridine, ac is acetate and Hac is acetic acid). The geometry around the metal centre is pentacoordinated, but highly distorted in all three cases. The coordination number and the geometric distortion are both discussed in detail, and all complexes belong to the space group Poverline{1}. The analysis of the geometric parameters and the Hirshfeld surface properties dnorm and curvedness provide information about the metal–ligand interactions in these complexes and allow comparison with similar systems. Full Article text
ine Revisiting a natural wine salt: calcium (2R,3R)-tartrate tetrahydrate By journals.iucr.org Published On :: 2024-09-04 The crystal structure of the salt calcium (2R,3R)-tartrate tetrahydrate {systematic name: poly[[diaqua[μ4-(2R,3R)-2,3-dihydroxybutanedioato]calcium(II)] dihydrate]}, {[Ca(C4H8O8)(H2O)2]·2H2O}n, is reported. The absolute configuration of the crystal was established unambiguously using anomalous dispersion effects in the diffraction patterns. High-quality data also allowed the location and free refinement of all the H atoms, and therefore to a careful analysis of the hydrogen-bond interactions. Full Article text
ine Deep residual networks for crystallography trained on synthetic data By journals.iucr.org Published On :: 2024-01-01 The use of artificial intelligence to process diffraction images is challenged by the need to assemble large and precisely designed training data sets. To address this, a codebase called Resonet was developed for synthesizing diffraction data and training residual neural networks on these data. Here, two per-pattern capabilities of Resonet are demonstrated: (i) interpretation of crystal resolution and (ii) identification of overlapping lattices. Resonet was tested across a compilation of diffraction images from synchrotron experiments and X-ray free-electron laser experiments. Crucially, these models readily execute on graphics processing units and can thus significantly outperform conventional algorithms. While Resonet is currently utilized to provide real-time feedback for macromolecular crystallography users at the Stanford Synchrotron Radiation Lightsource, its simple Python-based interface makes it easy to embed in other processing frameworks. This work highlights the utility of physics-based simulation for training deep neural networks and lays the groundwork for the development of additional models to enhance diffraction collection and analysis. Full Article text
ine The High-Pressure Freezing Laboratory for Macromolecular Crystallography (HPMX), an ancillary tool for the macromolecular crystallography beamlines at the ESRF By journals.iucr.org Published On :: 2024-01-24 This article describes the High-Pressure Freezing Laboratory for Macromolecular Crystallography (HPMX) at the ESRF, and highlights new and complementary research opportunities that can be explored using this facility. The laboratory is dedicated to investigating interactions between macromolecules and gases in crystallo, and finds applications in many fields of research, including fundamental biology, biochemistry, and environmental and medical science. At present, the HPMX laboratory offers the use of different high-pressure cells adapted for helium, argon, krypton, xenon, nitrogen, oxygen, carbon dioxide and methane. Important scientific applications of high pressure to macromolecules at the HPMX include noble-gas derivatization of crystals to detect and map the internal architecture of proteins (pockets, tunnels and channels) that allows the storage and diffusion of ligands or substrates/products, the investigation of the catalytic mechanisms of gas-employing enzymes (using oxygen, carbon dioxide or methane as substrates) to possibly decipher intermediates, and studies of the conformational fluctuations or structure modifications that are necessary for proteins to function. Additionally, cryo-cooling protein crystals under high pressure (helium or argon at 2000 bar) enables the addition of cryo-protectant to be avoided and noble gases can be employed to produce derivatives for structure resolution. The high-pressure systems are designed to process crystals along a well defined pathway in the phase diagram (pressure–temperature) of the gas to cryo-cool the samples according to the three-step `soak-and-freeze method'. Firstly, crystals are soaked in a pressurized pure gas atmosphere (at 294 K) to introduce the gas and facilitate its interactions within the macromolecules. Samples are then flash-cooled (at 100 K) while still under pressure to cryo-trap macromolecule–gas complexation states or pressure-induced protein modifications. Finally, the samples are recovered after depressurization at cryo-temperatures. The final section of this publication presents a selection of different typical high-pressure experiments carried out at the HPMX, showing that this technique has already answered a wide range of scientific questions. It is shown that the use of different gases and pressure conditions can be used to probe various effects, such as mapping the functional internal architectures of enzymes (tunnels in the haloalkane dehalogenase DhaA) and allosteric sites on membrane-protein surfaces, the interaction of non-inert gases with proteins (oxygen in the hydrogenase ReMBH) and pressure-induced structural changes of proteins (tetramer dissociation in urate oxidase). The technique is versatile and the provision of pressure cells and their application at the HPMX is gradually being extended to address new scientific questions. Full Article text
ine A service-based approach to cryoEM facility processing pipelines at eBIC By journals.iucr.org Published On :: 2024-02-20 Electron cryo-microscopy image-processing workflows are typically composed of elements that may, broadly speaking, be categorized as high-throughput workloads which transition to high-performance workloads as preprocessed data are aggregated. The high-throughput elements are of particular importance in the context of live processing, where an optimal response is highly coupled to the temporal profile of the data collection. In other words, each movie should be processed as quickly as possible at the earliest opportunity. The high level of disconnected parallelization in the high-throughput problem directly allows a completely scalable solution across a distributed computer system, with the only technical obstacle being an efficient and reliable implementation. The cloud computing frameworks primarily developed for the deployment of high-availability web applications provide an environment with a number of appealing features for such high-throughput processing tasks. Here, an implementation of an early-stage processing pipeline for electron cryotomography experiments using a service-based architecture deployed on a Kubernetes cluster is discussed in order to demonstrate the benefits of this approach and how it may be extended to scenarios of considerably increased complexity. Full Article text