Concomitant polymorphs of 1,3-bis(3-fluorophenyl)urea

2016 ◽  
Vol 72 (9) ◽  
pp. 692-696 ◽  
Author(s):  
Christina A. Capacci-Daniel ◽  
Jeffery A. Bertke ◽  
Shoaleh Dehghan ◽  
Rupa Hiremath-Darji ◽  
Jennifer A. Swift

Hydrogen bonding between urea functionalities is a common structural motif employed in crystal-engineering studies. Crystallization of 1,3-bis(3-fluorophenyl)urea, C13H10F2N2O, from many solvents yielded concomitant mixtures of at least two polymorphs. In the monoclinic form, one-dimensional chains of hydrogen-bonded urea molecules align in an antiparallel orientation, as is typical of many diphenylureas. In the orthorhombic form, one-dimensional chains of hydrogen-bonded urea molecules have a parallel orientation rarely observed in symmetrically substituted diphenylureas.

2017 ◽  
Vol 70 (5) ◽  
pp. 529 ◽  
Author(s):  
Hasti Iranmanesh ◽  
Kasun S. A. Arachchige ◽  
William A. Donald ◽  
Niamh Kyriacou ◽  
Chao Shen ◽  
...  

Four heteroleptic ruthenium(ii) complexes of 4′-functionalised 2,2′:6′,2′′-terpyridine are reported, along with their solid-state single-crystal X-ray structures. The complexes feature complementary hydrogen-bond donor (phenol) and acceptor (pyridyl) groups designed to assemble into one-dimensional polymers. In one example, the system obeys the programmed instructions to form a one-dimensional, self-complementary hydrogen-bonded polymer. In one other example, a water-bridged hydrogen-bonded polymer is formed. In the remaining two structures, aryl–aryl interactions dominate the intermolecular interactions, and outweigh the contribution of intermolecular hydrogen bonding.


2012 ◽  
Vol 2012 ◽  
pp. 1-10
Author(s):  
Makoto Tadokoro ◽  
Kyosuke Isoda ◽  
Yasuko Tanaka ◽  
Yuko Kaneko ◽  
Syoko Yamamoto ◽  
...  

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


2018 ◽  
Vol 74 (1) ◽  
pp. 21-30 ◽  
Author(s):  
Wilhelm Maximilian Hützler ◽  
Michael Bolte

Doubly and triply hydrogen-bonded supramolecular synthons are of particular interest for the rational design of crystal and cocrystal structures in crystal engineering since they show a high robustness due to their high stability and good reliability. The compound 5-methyl-2-thiouracil (2-thiothymine) contains an ADA hydrogen-bonding site (A = acceptor and D = donor) if the S atom is considered as an acceptor. We report herein the results of cocrystallization experiments with the coformers 2,4-diaminopyrimidine, 2,4-diamino-6-phenyl-1,3,5-triazine, 6-amino-3H-isocytosine and melamine, which contain complementary DAD hydrogen-bonding sites and, therefore, should be capable of forming a mixed ADA–DAD N—H...S/N—H...N/N—H...O synthon (denoted synthon 3s N·S;N·N;N·O), consisting of three different hydrogen bonds with 5-methyl-2-thiouracil. The experiments yielded one cocrystal and five solvated cocrystals, namely 5-methyl-2-thiouracil–2,4-diaminopyrimidine (1/2), C5H6N2OS·2C4H6N4, (I), 5-methyl-2-thiouracil–2,4-diaminopyrimidine–N,N-dimethylformamide (2/2/1), 2C5H6N2OS·2C4H6N4·C3H7NO, (II), 5-methyl-2-thiouracil–2,4-diamino-6-phenyl-1,3,5-triazine–N,N-dimethylformamide (2/2/1), 2C5H6N2OS·2C9H9N5·C3H7NO, (III), 5-methyl-2-thiouracil–6-amino-3H-isocytosine–N,N-dimethylformamide (2/2/1), (IV), 2C5H6N2OS·2C4H6N4O·C3H7NO, (IV), 5-methyl-2-thiouracil–6-amino-3H-isocytosine–N,N-dimethylacetamide (2/2/1), 2C5H6N2OS·2C4H6N4O·C4H9NO, (V), and 5-methyl-2-thiouracil–melamine (3/2), 3C5H6N2OS·2C3H6N6, (VI). Synthon 3s N·S;N·N;N·O was formed in three structures in which two-dimensional hydrogen-bonded networks are observed, while doubly hydrogen-bonded interactions were formed instead in the remaining three cocrystals whereby three-dimensional networks are preferred. As desired, the S atoms are involved in hydrogen-bonding interactions in all six structures, thus illustrating the ability of sulfur to act as a hydrogen-bond acceptor and, therefore, its value for application in crystal engineering.


2004 ◽  
Vol 82 (2) ◽  
pp. 386-398 ◽  
Author(s):  
Dominic Laliberté ◽  
Thierry Maris ◽  
James D Wuest

Tetraphenylmethane, tetraphenylsilane, and simple derivatives with substituents that do not engage in hydrogen bonding typically crystallize as close-packed structures with essentially no space available for the inclusion of guests. In contrast, derivatives with hydrogen-bonding groups are known to favor the formation of open networks that include significant amounts of guests. To explore this phenomenon, we synthesized six new derivatives 5a–5e and 6a of tetraphenylmethane and tetraphenylsilane with urethane and urea groups at the para positions, crystallized the compounds, and determined their structures by X-ray crystallography. As expected, all six compounds crystallize to form porous three-dimensional hydrogen-bonded networks. In the case of tetraurea 5e, 66% of the volume of the crystals is accessible to guests, and guests can be exchanged in single crystals without loss of crystallinity. Of special note are: (i) the use of tetrakis(4-isocyanatophenyl)methane (1f) as a precursor for making enantiomerically pure tetraurethanes and tetraureas, including compounds 5b, 5c; and (ii) their subsequent crystallization to give porous chiral hydrogen-bonded networks. Such materials promise to include chiral guests enantioselectively and to be useful in the separation of racemates, asymmetric catalysis, and other applications.Key words: crystal engineering, molecular tectonics, hydrogen bonding, networks, porosity, urethanes, ureas, tetraphenylmethane, tetraphenylsilane.


1996 ◽  
Vol 52 (6) ◽  
pp. 989-998 ◽  
Author(s):  
Q. Li ◽  
T. C. W. Mak

The new inclusion complexes tetraethylammonium nitrate–thiourea (1:3), (C2H5)4N+.NO3 −. 3(NH2)2CS (1), tetra-n-propylammonium nitrate–thiourea–water (1:3:1), (n-C3H7)4N+.NO3 −.3(NH2)2CS.H2O (2), tetramethylammonium nitrate–thiourea (1:1), (CH3)4N+.NO3 −.(NH2)2CS (3), tetra-n-propylammonium nitrate–thiourea (1:1), (n-C3H7)4N+.NO3 −. (NH2)2CS (4), and tetra-n-butylammonium nitrate–thiourea (1:1), (n-C4H9)4N+.NO3 −.(NH2)2CS (5) have been prepared and characterized by X-ray crystallography. Crystal data, Mo Kα: (1), space group P{\bar 1}, a = 10.300 (2), b = 14.704 (3), c = 15.784 (4) Å, α = 75.30 (3), β = 86.98 (3), γ = 72.25 (3)°, Z = 4 and RF = 0.039 for 5034 observed data; (2), space group P21/n, a = 8.433 (2), b = 9.369 (2), c = 34.361 (7) Å, β = 91.01 (3)°, Z = 4 and RF = 0.050 for 2475 observed data; (3), space group Pnma, a = 15.720 (3), b = 8.218 (2), c = 8.709 (2) Å, Z = 4 and RF = 0.073 for 579 observed data; (4), space group P21/n, a = 8.784 (2), b = 14.421 (3), c = 15.078 (3) Å, β = 92.31 (3)°, Z = 4 and RF = 0.046 for 2507 observed data; (5), space group Pna21, a = 19.934 (3), b = 12.680 (2), c = 9.092 (3) Å, Z = 4 and RF = 0.047 for 1646 observed data. In the crystal structure of (1) infinite chains each composed of an alternate arrangement of a twisted thiourea trimer and a nitrate ion are cross-linked to form a puckered layer and further hydrogen bonding between such layers leads to a channel host framework for accommodation of the cationic guests. In the crystal structure of (2) two independent thiourea molecules are used to construct a hydrogen-bonded puckered layer normal to the c axis, whereas the remaining thiourea molecule, together with the nitrate ion and water molecule, generate another puckered layer that is parallel to the first. Hydrogen bonding between these two types of layers gives rise to a channel system running parallel to the [100] direction and the cations are stacked regularly within each column. Compounds (3), (4) and (5) have closely related crystal structures in which the cations are separated by one-dimensional, infinitely extended thiourea–nitrate composite ribbons in a sandwich-like packing arrangement.


1993 ◽  
Vol 328 ◽  
Author(s):  
Mike Zaworotko ◽  
S. Subramanian ◽  
L. R. Macgillivray

ABSTRACTCrystal engineering has been invoked to design structural analogues of two prototypal SHG active solids, p-nitroaniline (pNA) and potassium dihydrogenphosphate (KDP). pNA exists as linear polar strands because of head-to-tail hydrogen bonding between adjacent molecules whereas KDP is a self-assembled hydrogen bonded diamondoid network that becomes polar when the hydrogen bonds align. We detail preparation and crystallographic characterization of two classes of multicomponent solid, organic cation hydrogen sulfates and cocrystals of the cubane cluster [M (CO)3(μ3-OH)]4, which structurally mimic pNA and KDP, respectively. Several of the Multi-component solids are polar and they represent a generic approach to designing new polar materials since one component can be changed without altering the basic architecture within the crystal.


Author(s):  
Sophia Lipstman ◽  
Israel Goldberg

Crystal engineering studies confirm that the zinc-tetra(4-pyridyl)porphyrin building block reveals versatile supramolecular chemistry. In this work, it was found to be reactive in the assembly of both (a) a 2D polymeric array by a unique combination of self-coordination and coordination through external zinc dichloride linkers and (b) an extended heteromolecular hydrogen-bonded network with mellitic acid sustained by multiple connectivity between the component species.


2018 ◽  
Vol 17 (5) ◽  
pp. 670-680 ◽  
Author(s):  
Gabriela Ortega ◽  
Jesús Hernández ◽  
Teresa González ◽  
Romano Dorta ◽  
Alexander Briceño

The crystal engineering of hydrogen bonded organic assemblies based on 1,2,4,5-benzenetetracarboxylic acid and stilbazole derivatives (1–10) is exploited in order to design novel photoreactive ionic assemblies.


2005 ◽  
Vol 60 (4) ◽  
pp. 363-372 ◽  
Author(s):  
Athanassios K. Boudalis ◽  
Vassilios Nastopoulos ◽  
Catherine P. Raptopoulou ◽  
Aris Terzis ◽  
Spyros P. Perlepes

In order to examine the possibility of using yttrium(III) in the crystal engineering of hydrogenbonded coordination complexes and to compare the molecular and supramolecular YIII/Cl3 or NO3-/DMU chemistry with the already well-developed LnIII/Cl− or NO3−/DMU chemistry (LnIII = lanthanide, DMU = N,N’-dimethylurea), compounds [Y(DMU)6][YCl6] (1) and [Y(NO3)3(DMU)3] (2) have been prepared. The structures of both compounds have been determined by single-crystal Xray diffraction. The structure of 1 consists of octahedral [Y(DMU)6]3+ and [YCl6]3− ions. The YIII ion in 2 is nine-coordinate and ligation is provided by three O-bonded DMU ligands and three bidentate chelating nitrato groups; the coordination polyhedron about the metal can be viewed as a distorted, monocapped square antiprism. The [Y(DMU)6]3+ cations and [YCl6]3− anions self-assemble to form a hydrogen-bonded 3D architecture in 1. Most of the hydrogen-bonding functionalities on the components of 2 create also a 3D network. Two motifs of interionic/intramolecular hydrogen-bonds have been observed: N-H···Cl in 1 and N-H···O(NO3−) in 2. The IR data are discussed in terms of the nature of bonding and the structures of the two complexes


Sign in / Sign up

Export Citation Format

Share Document