Polysulfonylamine, CLV [1]. Starke und schwache Wasserstoffbrücken in den Kristallstrukturen von 2,2΄-Bipyridinium-, 1,10-Phenanthrolinium- und 1,8-Bis(dimethylamino)naphthalinium-dimesylamid / Polysulfonylamines, CLV [1]. Strong and Weak Hydrogen Bonding in the Crystal Structures of 2,2΄-Bipyridinium, 1,10-Phenanthrolinium and 1,8-Bis(dimethylamino)naphthalinium Dimesylamide

2002 ◽  
Vol 57 (7) ◽  
pp. 777-790 ◽  
Author(s):  
Dagmar Henschel ◽  
Oliver Moers ◽  
Ilona Lange ◽  
Armand Blaschette ◽  
Peter G Jones

As a sequel to prior reports on strong and weak hydrogen bonding in onium di(methanesulfonyl) amide crystals, low-temperature X-ray structures are described for three salts of general formula BH+(MeSO2)2N-, where BH+ is 2,2΄-bipyridinium (1; monoclinic, space group P21/n, Z΄ = 1), 1,10-phenanthrolinium (2; monoclinic, P21/c, Z΄ = 2), or 1,8-bis(dimethylamino) naphthalinium (3; orthorhombic, P212121, Z΄ = 1). Monoprotonation of the organic bases by (MeSO2)2NH results in the formation of an intra-cation N-H···N hydrogen bond, which is asymmetric in 1 and 2, but approximately symmetric in the proton-sponge cation of 3. Moreover, the acidic H atom is engaged in a cation-anion contact N-H···N- in 1 and 2 or H+···Oδ- in 3, thus conferring three-centre character upon the strong hydrogen bonding. Each structure displays a multitude of close interionic C-H···O/N contacts that are geometrically consistent with weak hydrogen bonding. A salient feature is provided by short S-CH2-H···O-S inter-anion contacts, which lead to layers in 1 and to catemers in 2, but are non-existent in structure 3. The cations of both 1 and 2 form π-stacks that are intercalated between the anion layers or surrounded by six anion catemers, whereas in structure 3 each cation is octahedrally coordinated by six anions and vice-versa. The heteroionic connectivity comprises the aforementioned branches of the strong three-centre hydrogen bonds (in 1-3), numerous Car-H···A bonds (1, 2: A = O; 3: A = O, N), S-CH2-H···Nring interactions (1, 2), and close N-CH2-H···O=S contacts (3; possibly destabilizing).

2002 ◽  
Vol 57 (5) ◽  
pp. 534-546 ◽  
Author(s):  
Dagmar Henschel ◽  
Oliver Moers ◽  
Karna Wijaya ◽  
Andreas Wirth ◽  
Armand Blaschette ◽  
...  

In order to study weak hydrogen bonds originating from inductively activated C(sp3)-H donor groups, low-temperature X-ray structures are reported for three onium salts of general formula BH+(MeSO2)2N-, where BH+ is Me3N+CH2CH2OH (1; orthorhombic, space group P212121, Z′ = 1), Me3N+CH2C(O)OH (2; orthorhombic, P212121, Z′ = 1), or Me2HN+CH2CH2NMe2 (3; monoclinic, P21/c, Z′ = 1). The asymmetric units consist of cationanion pairs assembled by an O-H···O=S hydrogen bond in 1, an O-H···N- bond in 2, and an N+-H ··· N- bond in 3. The packings display a plethora of short interionic C(sp3)-H···O/N contacts that are geometrically consistent with weak hydrogen bonding; those taken into consideration have normalized parameters d(H ··· O) ≤ 269 pm, d(H···N) ≤ 257 pm and θ(C-H···O/N) ≥ 127°. The roles of the weak hydrogen bonds are as follows: In structures 1 and 3, the anions are associated into corrugated layers, which intercalate catemers of cations (1) or stacks of discrete cations (3), whereas structure 2 involves cation catemers surrounded by four anion catemers and vice versa; moreover, all cations are linked to adjacent anions by several weak hydrogen bonds (and to one anion in particular by the strong H bond). Among the cation-anion interactions, the N+(CH2-H···)3O tripod pattern arising in 1 and 2 is of special interest.


2001 ◽  
Vol 56 (10) ◽  
pp. 1041-1051 ◽  
Author(s):  
Oliver Moers ◽  
Ilona Lange ◽  
Karna Wijaya ◽  
Armand Blaschette ◽  
Peter G. Jones

In order to study packing arrangements and hydrogen bonding networks, low-temperature X-ray structures were determined for pyH+(MeSO2)2N- (M, orthorhombic, space group P212121, Z′ = 1) and 4,4′-bipyH22+ ·(MeSO2)2N- (D, monoclinic, C2/c, Z′ = 0.5). The structures consist of ionic formula entities assembled by N+-H···N- hydrogen bonds; the dication in D displays crystallographic C2 symmetry and has its two pyridyl moieties twisted by 43.9°. According to the packing architectures, D represents a supramolecular dimer of the monomeric congener M. In particular, the (MeSO2)2N- ions of the M structure are associated via short C(sp3) - H···O contacts to form a diamondoid network, whereas in D a topologically congruent framework is constructed from weakly hydrogen-bonded [(MeSO2)N-]2 nodes. Hexagonal channels in the anion substructures each include two adjacent stacks of monomeric pyH+ or “dimeric” 4,4-bipyH22+ cations that are linked to the channel walls by the strong hydrogen bond(s) and a set of short Car-H···O contacts. All C - H···O taken into consideration have normalized parameters d(H···O) ≤ 270 pm and θ(C - H···O) ≥ 115°.


2000 ◽  
Vol 53 (6) ◽  
pp. 451 ◽  
Author(s):  
Murray S. Davies ◽  
Ronald R. Fenton ◽  
Fazlul Huq ◽  
Edwina C. H. Ling ◽  
Trevor W. Hambley

Two complexes, namely, chloro[N-(2-aminoethyl)-N-(2-ammonioethyl)ethane-1,2-diamine]platinum(II) chloride {[PtCl(tren+H)]Cl2} and dichloro[4,7-diaza-1-azoniacyclononane]platinum(II) tetrachloroplatinate(II)–water (1/2) {[PtCl2(tacn+H)]2[PtCl4]·2H2O}, have been prepared and structurally characterized by single-crystal X-ray diffractometry as part of a study of the nature and strength of Pt···H(–N) interactions. Crystals of [PtCl(tren+H)]Cl2 are monoclinic, space group P21/c, a 8.293(2), b 14.396(6), c 11.305(3) Å, β 107.34(2)º, Z 4, and the structure has been refined to a residual of 0.042 based on 1631 reflections. Crystals of [PtCl2(tacn+H)]2[PtCl4]·2H2O are monoclinic, space group P21/a, a 12.834(4), b 8.206(4), c 13.116(8) Å, β 93.01(4)˚, Z 2, and the structure has been refined to a residual of 0.035 based on 1974 reflections. In [PtCl(tren+H)]2+, the protonated amine forms hydrogen bonds with chloride anions and no close contacts with the metal ion. In [PtCl2(tacn+H)]+, a short intramolecular contact is observed between the metal and the protonated amine and the results of molecular mechanics modelling are consistent with there being a Pt···H hydrogen bond. Molecular mechanics modelling of [PtCl(tren+H)]2+ and [PtCl2(dien+H)]+ shows that the protonated amines could readily form close contacts with the metal. It is concluded that there is evidence for the formation of Pt···H(–N) hydrogen bonds but these bonds are very weak, being similar or lower in energy than Cl···H(–NPt) hydrogen bonds.


2007 ◽  
Vol 63 (3) ◽  
pp. 448-458 ◽  
Author(s):  
El-Eulmi Bendeif ◽  
Slimane Dahaoui ◽  
Nourredine Benali-Cherif ◽  
Claude Lecomte

The crystal structures of three similar guaninium salts, guaninium monohydrogenphosphite monohydrate, C5H6N5O+·H2O3P−·H2O, guaninium monohydrogenphosphite dihydrate, C5H6N5O+·H2O3P−·2H2O, and guaninium dihydrogenmonophosphate monohydrate, C5H6N5O+·H2O4P−·H2O, are described and compared. The crystal structures have been determined from accurate single-crystal X-ray data sets collected at 100 (2) K. The two phosphite salts are monoclinic, space group P21/c, with different packing and the monophosphate salt is also monoclinic, space group P21/n. An investigation of the hydrogen-bond network in these guaninium salts reveals the existence of two ketoamine tautomers, the N9H form and an N7H form.


2014 ◽  
Vol 69 (7) ◽  
pp. 839-843 ◽  
Author(s):  
Guido D. Frey ◽  
Wolfgang W. Schoeller ◽  
Eberhardt Herdtweck

The crystal structure of 1-(1H-pyrazol-4-yl)ethanone (commonly known as 4-acetylpyrazole; C5H6N2O) was determined from single-crystal X-ray data at 173 K: monoclinic, space group P21/n (no. 14), a = 3.865(1), b = 5.155(1), c = 26.105(8) Å, β = 91.13(1)°, V = 520.0(2) Å3 and Z = 4. The adjacent molecules assemble into a wave-like ribbon structure in the solid state, linked by strong intermolecular N-H...N hydrogen bonds between the pyrazole rings and a weak C-H...O=C hydrogen bond involving the carbonyl group. The ribbons are stacked in the solid state via weak π interactions between the pyrazole rings.


1999 ◽  
Vol 54 (11) ◽  
pp. 1420-1430 ◽  
Author(s):  
Oliver Moers ◽  
Karna Wijaya ◽  
Dagmar Henschel ◽  
Armand Blaschette ◽  
Peter G. Jones

In order to examine packing preferences and hydrogen bond patterns in secondary ammonium salts, low-temperature X-ray analyses were conducted for six compounds of general formula R2NH2+MeSO2)2 N-, where R2NH2+ = Me2NH2+ (1, triclinic, space group P1̄̄), MeEtNH2+,(2, monoclinic, P21/c), Et2NH2+ (3. triclinic, P1), pyrrolidinium (4, triclinic, P1), piperidinium (5, monoclinic, C2/c) or morpholinium (6, monoclinic, P21/c). Throughout the series, the constant anion retains a rigid conformation approximating to C2 symmetry and thus provides a geometrically reliable set of five potential hydrogen bond acceptors. Nevertheless, the six compounds exhibit a variety of unpredictable packing patterns, showing that, in unfavourable cases, the steric demands of molecular fragments not involved in hydrogen bonding can substantially alter the structure of a family of ionic crystals. In the present structures, the NH2+ donor groups form hydrogen bonds N+-H···N-/O to two (3-6) or three (1,2) adjacent anions. The occurrence of various two-, three- and four-centre hydrogen bonds leads to six different patterns, resulting in cation-anion layers (1, 2), discrete formula unit dimers (3, 4) or cation-anion chains (5, 6); in the morpholinium salt 6, these chains are associated into layers by a weak N+ - H ··· O(cation) interaction. In each of the crystal packings, short C-H···O contacts with H···O ≤ 270 pm and C-H ···O ≥ 130° are observed.


Author(s):  
Jan Vícha ◽  
Cina Foroutan-Nejad ◽  
Michal Straka

Illusive Au<sup>I/III</sup>···H hydrogen bonds and their effect on structure and dynamics of molecules have been a matter of debate. While a number of X-ray studies reported gold compounds with short Au<sup>I/III</sup>···H contacts, a solid spectroscopic evidence for Au<sup>I/III</sup>···H bonding has been missing. Recently<a></a><a>, Bakar <i>et al.</i></a> (NATURE COMMUNICATIONS 8:576) reported compound with four short Au···H contacts (2.61­–2.66 Å; X-ray determined). Assuming the central cluster be [Au<sub>6</sub>]<sup>2+</sup>and observing the <sup>1</sup>H (<sup>13</sup>C) NMR resonances at relevant H(C) nuclei deshielded with respect to precursor compound, the authors concluded with reservations that <i>“the present Au···H–C interaction is a kind of “hydrogen bond”, where the [Au<sub>6</sub>]<sup>2+</sup>serves as an acceptor”</i>. Here, we show that the Au<sub>6</sub>cluster in their compound bears negative charge and the Au···H contacts lead to a weak (~1 kcal/mol) auride···hydrogen bonding interactions, though unimportant for the overall stability of<b></b>the molecule. Additionally, computational analysis of NMR chemical shifts reveals that the deshielding effects at respective hydrogen nuclei are not directly related to Au···H–C hydrogen bonding .


2015 ◽  
Vol 71 (8) ◽  
pp. 733-741
Author(s):  
V. S. Minkov ◽  
V. V. Ghazaryan ◽  
E. V. Boldyreva ◽  
A. M. Petrosyan

L-Cysteine hydrogen fluoride, or bis(L-cysteinium) difluoride–L-cysteine–hydrogen fluoride (1/1/1), 2C3H8NO2S+·2F−·C3H7NO2S·HF or L-Cys+(L-Cys...L-Cys+)F−(F−...H—F), provides the first example of a structure with cations of the `triglycine sulfate' type,i.e.A+(A...A+) (whereAandA+are the zwitterionic and cationic states of an amino acid, respectively), without a doubly charged counter-ion. The salt crystallizes in the monoclinic system with the space groupP21. The dimeric (L-Cys...L-Cys+) cation and the dimeric (F−...H—F) anion are formedviastrong O—H...O or F—H...F hydrogen bonds, respectively, with very short O...O [2.4438 (19) Å] and F...F distances [2.2676 (17) Å]. The F...F distance is significantly shorter than in solid hydrogen fluoride. Additionally, there is another very short hydrogen bond, of O—H...F type, formed by a L-cysteinium cation and a fluoride ion. The corresponding O...F distance of 2.3412 (19) Å seems to be the shortest among O—H...F and F—H...O hydrogen bonds known to date. The single-crystal X-ray diffraction study was complemented by IR spectroscopy. Of special interest was the spectral region of vibrations related to the above-mentioned hydrogen bonds.


1988 ◽  
Vol 58 (2) ◽  
pp. 96-101 ◽  
Author(s):  
Stanley P. Rowland ◽  
Phyllis S. Howley

The extent of hydrogen bonding of O(3)H and O(6)H in “amorphous” regions, more specifically in accessible segments of fibrils, of the cotton fiber varied from near perfection to almost complete disorder in samples under examination. Perfection of hydrogen bonding in various samples and segments of cotton fibers decreased with decreasing crystallinity of the cellulose within the fibrils. For the most part, extents of O(3)H hydrogen bonding and O(6)H hydrogen bonding followed similar patterns with substantial differences in degrees of perfection, the O(3)H ranging from about 95% hydrogen bonding down to 8% and the O(6)H) from 92% down to 41%. Details of hydrogen bonds assessed in these chemical studies are discussed relative to crystallinities and assignments of hydrogen bond structures from x-ray diffraction studies.


2018 ◽  
Author(s):  
Jan Vícha ◽  
Cina Foroutan-Nejad ◽  
Michal Straka

Illusive Au<sup>I/III</sup>···H hydrogen bonds and their effect on structure and dynamics of molecules have been a matter of debate. While a number of X-ray studies reported gold compounds with short Au<sup>I/III</sup>···H contacts, a solid spectroscopic evidence for Au<sup>I/III</sup>···H bonding has been missing. Recently<a></a><a>, Bakar <i>et al.</i></a> (NATURE COMMUNICATIONS 8:576) reported compound with four short Au···H contacts (2.61­–2.66 Å; X-ray determined). Assuming the central cluster be [Au<sub>6</sub>]<sup>2+</sup>and observing the <sup>1</sup>H (<sup>13</sup>C) NMR resonances at relevant H(C) nuclei deshielded with respect to precursor compound, the authors concluded with reservations that <i>“the present Au···H–C interaction is a kind of “hydrogen bond”, where the [Au<sub>6</sub>]<sup>2+</sup>serves as an acceptor”</i>. Here, we show that the Au<sub>6</sub>cluster in their compound bears negative charge and the Au···H contacts lead to a weak (~1 kcal/mol) auride···hydrogen bonding interactions, though unimportant for the overall stability of<b></b>the molecule. Additionally, computational analysis of NMR chemical shifts reveals that the deshielding effects at respective hydrogen nuclei are not directly related to Au···H–C hydrogen bonding .


Sign in / Sign up

Export Citation Format

Share Document