Search results

Search for "hydrogen bond" in Full Text gives 390 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Small anion-assisted electrochemical potential splitting in a new series of bistriarylamine derivatives: organic mixed valency across a urea bridge and zwitterionization

  • Keishiro Tahara,
  • Tetsufumi Nakakita,
  • Alyona A. Starikova,
  • Takashi Ikeda,
  • Masaaki Abe and
  • Jun-ichi Kikuchi

Beilstein J. Org. Chem. 2019, 15, 2277–2286, doi:10.3762/bjoc.15.220

Graphical Abstract
  • . Such N–H···F hydrogen-bond formation was also reported for other urea derivatives with PF6− as counteranion in the solid state [54][55]. The N···F distance of 2.85 Å in 1b+–PF6− is slightly longer than that observed in the crystal structure of a silver complex having a pyridyl urea ligand (2.67 and
PDF
Album
Supp Info
Full Research Paper
Published 24 Sep 2019

Aggregation-induced emission effect on turn-off fluorescent switching of a photochromic diarylethene

  • Luna Kono,
  • Yuma Nakagawa,
  • Ayako Fujimoto,
  • Ryo Nishimura,
  • Yohei Hattori,
  • Toshiki Mutai,
  • Nobuhiro Yasuda,
  • Kenichi Koizumi,
  • Satoshi Yokojima,
  • Shinichiro Nakamura and
  • Kingo Uchida

Beilstein J. Org. Chem. 2019, 15, 2204–2212, doi:10.3762/bjoc.15.217

Graphical Abstract
  • because the proton transfer occurs through an intramolecular hydrogen bond. This is also an origin of a large Stokes shift (8,000–11,000 cm−1) in the emission from the ESIPT state. Consequently, yellow luminescence was observed by UV-excitation [17][18]. Mutai et al. reported an ESIPT luminescence of an
PDF
Album
Supp Info
Full Research Paper
Published 20 Sep 2019

Click chemistry towards thermally reversible photochromic 4,5-bisthiazolyl-1,2,3-triazoles

  • Chenxia Zhang,
  • Kaori Morinaka,
  • Mahmut Kose,
  • Takashi Ubukata and
  • Yasushi Yokoyama

Beilstein J. Org. Chem. 2019, 15, 2161–2169, doi:10.3762/bjoc.15.213

Graphical Abstract
  • , differences can still be observed. In general, the thermal back reactions are faster in the more polar solvents, and faster in MeCN than in EtOH among these two solvents. However, for 2c, it is faster in EtOH than in MeCN. This may have originated from the hydrogen-bond formation between EtOH molecules and
PDF
Album
Supp Info
Letter
Published 13 Sep 2019

1,2,3-Triazolium macrocycles in supramolecular chemistry

  • Mastaneh Safarnejad Shad,
  • Pulikkal Veettil Santhini and
  • Wim Dehaen

Beilstein J. Org. Chem. 2019, 15, 2142–2155, doi:10.3762/bjoc.15.211

Graphical Abstract
  • selectivity of this receptor for chloride anion is due to the cavity size (see Figure 3) [37]. 2.2. Optical anion sensing by 1,2,3-triazolium macrocycles within porphyrin cages Various porphyrin-based host supramolecules containing hydrogen-bond donor groups such as integrated amide, urea, pyrrole, ammonium
  • donors: The benefits of combining the triazolium motif with other hydrogen bond donor motifs has been explored by Sessler et al. and they have synthesized a pyrrole based tetra-1,2,3-triazolium macrocycle 10 (Figure 10) via the tetraalkylation of a triazole macrocycle originally prepared via click
  • chemistry. Different hydrogen bond donors such as pyrrole N–H and benzene C–H along with the triazolium C5–H have been successfully incorporated within a single macrocyclic framework. The complex formation of anions with the hydrogen bond donors varied according to the polarity of the solvent. It was found
PDF
Album
Review
Published 12 Sep 2019

1,2,3,4-Tetrahydro-1,4,5,8-tetraazaanthracene revisited: properties and structural evidence of aromaticity loss

  • Arnault Heynderickx,
  • Sébastien Nénon,
  • Olivier Siri,
  • Vladimir Lokshin and
  • Vladimir Khodorkovsky

Beilstein J. Org. Chem. 2019, 15, 2059–2068, doi:10.3762/bjoc.15.203

Graphical Abstract
  • ) and N–H···Cl (6a). The experimental (averaged) and calculated selected bond lengths are listed in Table 1. There are two independent molecules of 3 in the cell, bound with the short =N···H–N< hydrogen bond of 2.184 Å (Figure 2). Examination of the bond distances within the N2–C1–C14–C13–N12, N5–C6–C7
PDF
Album
Supp Info
Full Research Paper
Published 28 Aug 2019

Synthesis of 1-azaspiro[4.4]nonan-1-oxyls via intramolecular 1,3-dipolar cycloaddition

  • Yulia V. Khoroshunova,
  • Denis A. Morozov,
  • Andrey I. Taratayko,
  • Polina D. Gladkikh,
  • Yuri I. Glazachev and
  • Igor A. Kirilyuk

Beilstein J. Org. Chem. 2019, 15, 2036–2042, doi:10.3762/bjoc.15.200

Graphical Abstract
  • nitroxides [26], implying free rotation of ethyl groups. In the EPR spectra in toluene the nitroxides 11a–c show 0.035–0.04 mT lower hfs constants on nitrogen atom compared to nitroxides 12a–c, presumably due to intramolecular hydrogen bond formation. The difference in aN between 11a–c and 12a–c is almost
PDF
Album
Supp Info
Full Research Paper
Published 27 Aug 2019
Graphical Abstract
  • conclusion comes from a comparison of complexes in methanol and DMSO. One would expect that due to the dual hydrogen bond donor/acceptor character of methanol this solvent should be more destructive to ion-pair complexes based on charged hydrogen bonds than DMSO that can only serve as a hydrogen bond
PDF
Album
Supp Info
Full Research Paper
Published 12 Aug 2019

Tautomerism as primary signaling mechanism in metal sensing: the case of amide group

  • Vera Deneva,
  • Georgi Dobrikov,
  • Aurelien Crochet,
  • Daniela Nedeltcheva,
  • Katharina M. Fromm and
  • Liudmil Antonov

Beilstein J. Org. Chem. 2019, 15, 1898–1906, doi:10.3762/bjoc.15.185

Graphical Abstract
  • can be seen that a hydrogen bond is formed between the nitrogen proton of one keto tautomer and the carbonyl group of another neighboring molecule. Probably, the process of associate formation is facilitated by the position of the chromophore part in the isolated K form (Figure 3, left). Obviously
  • , based on 4-(phenyldiazenyl)naphthalen-1-ol. According to the theoretical calculations the enol form stabilization could be achieved through a strong intramolecular hydrogen bond formed between the tautomeric hydroxy group and the carbonyl group from the tautomeric backbone. However, intermolecular
PDF
Album
Supp Info
Full Research Paper
Published 08 Aug 2019

Complexation of 2,6-helic[6]arene and its derivatives with 1,1′-dimethyl-4,4′-bipyridinium salts and protonated 4,4'-bipyridinium salts: an acid–base controllable complexation

  • Jing Li,
  • Qiang Shi,
  • Ying Han and
  • Chuan-Feng Chen

Beilstein J. Org. Chem. 2019, 15, 1795–1804, doi:10.3762/bjoc.15.173

Graphical Abstract
  • with larger counteranions of the guests and in less polar solvents. Furthermore, the switchable complexation between the helic[6]arene and protonated 4,4'-bipyridinium salt could be efficiently controlled by acids and bases. Keywords: 4,4'-bipyridinium salts; complexation; helic[6]arene; hydrogen bond
  • protons of G1 and the aromatic rings of H1 with distances of 2.683 for A, 2.845 for B, 2.788 for C, 2.802 for D, and 2.868 Å for E, respectively. There also exist π–π stacking interactions between the pyridinium of G1 and the aromatic ring of H1 with the distance of 3.854 Å for F, a CH···O hydrogen bond
PDF
Album
Supp Info
Full Research Paper
Published 26 Jul 2019

Water inside β-cyclodextrin cavity: amount, stability and mechanism of binding

  • Stiliyana Pereva,
  • Valya Nikolova,
  • Silvia Angelova,
  • Tony Spassov and
  • Todor Dudev

Beilstein J. Org. Chem. 2019, 15, 1592–1600, doi:10.3762/bjoc.15.163

Graphical Abstract
  • /narrow belt) can be arranged in two alternative ways: (i) facing inward and creating a hydrogen-bond girdle (“head–tail” arrangement) that greatly decreases the size of the cavity vestibule (Figure 1A), and (ii) facing outward thus widening the aperture (“open” configuration, Figure 1B). Notably, in the
  • “open” configuration the primary hydroxy groups do not participate in intramolecular hydrogen bond interactions. Relatively weak hydrogen bonds are formed between the secondary hydroxy groups in the wider/lower rim of the molecule. The calculations reveal that the “head–tail” arrangement of the narrow
  • rim is the preferred mode of coordination (n = 2; Figure 2, structure a). In this construct the water dimer not only connects two oppositely located OH groups from the rim, but additionally interacts with four other OH groups and forms an elaborate hydrogen-bond network that almost occludes the
PDF
Album
Supp Info
Full Research Paper
Published 17 Jul 2019

Synthesis and conformational preferences of short analogues of antifreeze glycopeptides (AFGP)

  • Małgorzata Urbańczyk,
  • Michał Jewgiński,
  • Joanna Krzciuk-Gula,
  • Jerzy Góra,
  • Rafał Latajka and
  • Norbert Sewald

Beilstein J. Org. Chem. 2019, 15, 1581–1591, doi:10.3762/bjoc.15.162

Graphical Abstract
  • the formation of an intramolecular hydrogen bond are characterized by chemical shifts that are independent on the temperature. When the amide proton is involved in a hydrogen bond, its temperature factor value is lower than 4.0 ppb/K [29][30]. Values of the measured temperature coefficient factors
  • (see Table 1) suggest that the conformation of each of the investigated pentapeptides is stabilized by one hydrogen bond, created by the amide proton of Ala4. In the case of peptide 4 we could observe the strongest hydrogen bond (Table 1). Previous studies pointed out the role of specific hydrogen
  • bonds between the peptide backbone and GalNAc moiety, which are responsible for creating the extended conformation [31][32]. However, further NMR investigations in this area found no evidence of an intramolecular hydrogen bond, or the existence of a very weak interaction, between the carbonyl group of
PDF
Album
Supp Info
Full Research Paper
Published 16 Jul 2019

Enantioselective PCCP Brønsted acid-catalyzed aza-Piancatelli rearrangement

  • Gabrielle R. Hammersley,
  • Meghan F. Nichol,
  • Helena C. Steffens,
  • Jose M. Delgado,
  • Gesine K. Veits and
  • Javier Read de Alaniz

Beilstein J. Org. Chem. 2019, 15, 1569–1574, doi:10.3762/bjoc.15.160

Graphical Abstract
  • hydrogen bond group afforded the best selectivity (7k). To simplify the purification process, all acid products were transformed into the corresponding methyl ester in situ using (trimethylsilyl)diazomethane. A slight drop in enantioselectivity is observed when the benzoic acid group was moved to the meta
  • furan ring required to initiate the cascade sequence. This highlights that a balance between nucleophilicity and hydrogen bond capabilities are required to obtain an efficient and selective reaction. Compared to the rearrangement of furylcarbinol 1a, sterically bulky aryl groups attached to the
PDF
Album
Supp Info
Letter
Published 12 Jul 2019

A heteroditopic macrocycle as organocatalytic nanoreactor for pyrroloacridinone synthesis in water

  • Piyali Sarkar,
  • Sayan Sarkar and
  • Pradyut Ghosh

Beilstein J. Org. Chem. 2019, 15, 1505–1514, doi:10.3762/bjoc.15.152

Graphical Abstract
  • environmentally oriented efficient organocatalysts with prominence “on-water conditions” [30][31]. It is known that an important condition for an organic reaction in water is the aggregation of the reactants and the catalyst by hydrophobic forces. Further, the potential of hydrogen-bond donation may be a vital
PDF
Album
Supp Info
Full Research Paper
Published 08 Jul 2019

Synthesis and biological evaluation of truncated derivatives of abyssomicin C as antibacterial agents

  • Leticia Monjas,
  • Peter Fodran,
  • Johanna Kollback,
  • Carlo Cassani,
  • Thomas Olsson,
  • Maja Genheden,
  • D. G. Joakim Larsson and
  • Carl-Johan Wallentin

Beilstein J. Org. Chem. 2019, 15, 1468–1474, doi:10.3762/bjoc.15.147

Graphical Abstract
  • conformations and with expected interaction points in the binding site, providing suitable docking scores (see Supporting Information File 1). The key interactions include a hydrogen bond to the backbone of Arg-45 and lipophilic interactions in the deep pocket defined by Phe-241 and Leu-34 (Figure 2). Our
PDF
Album
Supp Info
Letter
Published 02 Jul 2019

2,3-Dibutoxynaphthalene-based tetralactam macrocycles for recognizing precious metal chloride complexes

  • Li-Li Wang,
  • Yi-Kuan Tu,
  • Huan Yao and
  • Wei Jiang

Beilstein J. Org. Chem. 2019, 15, 1460–1467, doi:10.3762/bjoc.15.146

Graphical Abstract
  • parallel orientation. Three of the four NH protons are directed into the cavity, and the fourth one flipped outward and forms a hydrogen bond with the oxygen atom of H2O (H···O distance: 1.97 Å). Two CH3CN molecules were trapped in the cavities by the amide groups through N–H···N hydrogen bonds (H···N
PDF
Album
Supp Info
Full Research Paper
Published 02 Jul 2019

Selenophene-containing heterotriacenes by a C–Se coupling/cyclization reaction

  • Pierre-Olivier Schwartz,
  • Sebastian Förtsch,
  • Astrid Vogt,
  • Elena Mena-Osteritz and
  • Peter Bäuerle

Beilstein J. Org. Chem. 2019, 15, 1379–1393, doi:10.3762/bjoc.15.138

Graphical Abstract
  • (Figure 2b), because a completely flat geometry of the isolated molecule DST 3 (in the gas phase) was obtained from theoretical calculations (vide infra). Molecules of DST 3 order in a typical herringbone fashion, where the terminal hydrogen atoms form hydrogen bond-like C–H heteroatom interactions (2.819
PDF
Album
Supp Info
Full Research Paper
Published 24 Jun 2019

Stereo- and regioselective hydroboration of 1-exo-methylene pyranoses: discovery of aryltriazolylmethyl C-galactopyranosides as selective galectin-1 inhibitors

  • Alexander Dahlqvist,
  • Axel Furevi,
  • Niklas Warlin,
  • Hakon Leffler and
  • Ulf J. Nilsson

Beilstein J. Org. Chem. 2019, 15, 1046–1060, doi:10.3762/bjoc.15.102

Graphical Abstract
  • the anomeric position of the minimal D-galactose ligand allows for an additional affinity-enhancing hydrogen bond according to structural and affinity analyses [20][21]. A complementary strategy has been to, instead of a second saccharide unit, add non-natural structural elements to a monogalactoside
PDF
Album
Supp Info
Full Research Paper
Published 07 May 2019

Catalytic asymmetric oxo-Diels–Alder reactions with chiral atropisomeric biphenyl diols

  • Chi-Tung Yeung,
  • Wesley Ting Kwok Chan,
  • Wai-Sum Lo,
  • Ga-Lai Law and
  • Wing-Tak Wong

Beilstein J. Org. Chem. 2019, 15, 955–962, doi:10.3762/bjoc.15.92

Graphical Abstract
  • : asymmetric organocatalysis; axial chirality; biaryls; hydrogen bond; oxo-Diels–Alder reaction; Introduction The Diels–Alder (DA) reaction is a useful and easy-to-perform method for the synthesis of six-membered rings through the direct formation of C–C bonds between a diene and a dienophile (a substituted
  • hydrogen bonding organocatalyst for the reaction between 1-dimethylamino-3-tert-butyldimethylsilyloxy-1,3-butadiene (Rawal’s diene) and aldehydes with excellent enantioselectivities (aromatic aldehyde: up to 86–98% ee) in 2003 [27]. Activation via a single-point hydrogen bond between one of the hydroxy
  • other biphenyl hybrid diols with different steric bulky substituents were incorporated (3 and 4). Since it is known that the catalyst acidity has significant influence on hydrogen-bond-catalyzed reactions [32], we also incorporated a CF3 group in our molecular scaffold to investigate how it would affect
PDF
Album
Supp Info
Full Research Paper
Published 18 Apr 2019

Halogen bonding and host–guest chemistry between N-alkylammonium resorcinarene halides, diiodoperfluorobutane and neutral guests

  • Fangfang Pan,
  • Mohadeseh Dashti,
  • Michael R. Reynolds,
  • Kari Rissanen,
  • John F. Trant and
  • Ngong Kodiah Beyeh

Beilstein J. Org. Chem. 2019, 15, 947–954, doi:10.3762/bjoc.15.91

Graphical Abstract
  • inclusion guest (1,4-dioxane), the lattice solvent molecules, regardless of their polarity, insert between two N-hexyl arms using an OH/CH···Br− hydrogen bond (Figure 2) [32]. To better understand whether this is fundamental to these systems, or a result of the enforced geometry caused by the included guest
  • , which are ca. 41 and 53 Å3, respectively [35]. Hence, the resorcinarene has to deform to adapt to the small guests by maximizing the contacts between the host and guest (Figure 3). In detail, in the MeCN occupied systems, the N atom is stabilized by an NH···N hydrogen bond and a weak pnictogen bond to a
  • were positioned at the same level as the (HNH···Cl−)4 HB circle. The O···Cl− distances of 3.13(3) and 3.22(4) Å suggest OH···Cl− hydrogen bonds. Meanwhile, an OH···O hydrogen bond should also exist between the two water molecules with the O···O distance of 3.01(5) Å. It is surprising to find that the
PDF
Album
Supp Info
Full Research Paper
Published 18 Apr 2019

Mechanochemistry of supramolecules

  • Anima Bose and
  • Prasenjit Mal

Beilstein J. Org. Chem. 2019, 15, 881–900, doi:10.3762/bjoc.15.86

Graphical Abstract
  • . In 2008, they have established a [2 + 2] photodimerization through solid-state grinding either in neat or liquid-assisted conditions [93]. To achieve 100% stereospecific products they considered resorcinol derivatives as hydrogen-bond donors for the photodimerization of 1,2-di(pyridin-4-yl)ethylene
  • (Figure 21a). However, 1,8-dipyridylnaphthalene was used as hydrogen-bond acceptor for the [2 + 2] cycloaddition of fumaric acid derivatives (Figure 21b). In 2017 Purse and co-workers reported the host–guest chemistry of pyrogallo[4]arene (39) hexamers under mechano-milling conditions [94]. A hexameric
  • , supramolecular catalysis proposes a much simpler model to understand the catalytic activity of enzymes. In 2010, MacGillivray and co-workers have demonstrated the concept of “supramolecular catalysis” in a hydrogen-bond-assisted self-assembled formation of a [2 + 2]-cycloaddition product. The reaction was found
PDF
Album
Review
Published 12 Apr 2019

Coordination chemistry and photoswitching of dinuclear macrocyclic cadmium-, nickel-, and zinc complexes containing azobenzene carboxylato co-ligands

  • Jennifer Klose,
  • Tobias Severin,
  • Peter Hahn,
  • Alexander Jeremies,
  • Jens Bergmann,
  • Daniel Fuhrmann,
  • Jan Griebel,
  • Bernd Abel and
  • Berthold Kersting

Beilstein J. Org. Chem. 2019, 15, 840–851, doi:10.3762/bjoc.15.81

Graphical Abstract
  • lengths reveal no anomalies and are similar to those in [Zn2L(μ-OAc)]+. There are no π–π stacking interactions between the azobenzene moieties. However, the [Zn2L(μ-azo-OH)]+ and [Zn2L(μ-azo-O)] complexes are connected by a OH···O hydrogen bond of length 2.46 Å (O3a···O3b, not shown in Figure 4). [Zn2L(μ
PDF
Album
Supp Info
Full Research Paper
Published 03 Apr 2019

Catalyst-free assembly of giant tris(heteroaryl)methanes: synthesis of novel pharmacophoric triads and model sterically crowded tris(heteroaryl/aryl)methyl cation salts

  • Rodrigo Abonia,
  • Luisa F. Gutiérrez,
  • Braulio Insuasty,
  • Jairo Quiroga,
  • Kenneth K. Laali,
  • Chunqing Zhao,
  • Gabriela L. Borosky,
  • Samantha M. Horwitz and
  • Scott D. Bunge

Beilstein J. Org. Chem. 2019, 15, 642–654, doi:10.3762/bjoc.15.60

Graphical Abstract
  • distance between the quinolone carbonyl and the OH of hydroxycoumarin (H(1)–O(4) 1.740 Å). The DFT-optimized structure of 9{4,7,1} (Figure 4) confirms the formation of a highly stable hydrogen bond between the quinolone carbonyl and the OH of hydroxycoumarin, with a O···H bond distance of 1.603 Å. It
PDF
Album
Supp Info
Full Research Paper
Published 12 Mar 2019

Synthesis of the aglycon of scorzodihydrostilbenes B and D

  • Katja Weimann and
  • Manfred Braun

Beilstein J. Org. Chem. 2019, 15, 610–616, doi:10.3762/bjoc.15.56

Graphical Abstract
  • transition metal-mediated, regioselective, chelation-directed activation of an aryl–hydrogen bond. A promising approach was seen in an application of Murai’s elegant ruthenium-catalyzed ortho-functionalization of aryl alkyl ketones and their addition to olefins [11][12][13][14][15]. This route appeared
PDF
Album
Supp Info
Full Research Paper
Published 06 Mar 2019

Dirhodium(II)-catalyzed [3 + 2] cycloaddition of N-arylaminocyclopropane with alkyne derivatives

  • Wentong Liu,
  • Yi Kuang,
  • Zhifan Wang,
  • Jin Zhu and
  • Yuanhua Wang

Beilstein J. Org. Chem. 2019, 15, 542–550, doi:10.3762/bjoc.15.48

Graphical Abstract
  • 3,5-dimethyl compound 1d afforded the desired product 3d with an excellent yield of 91% (Table 2, entry 3). We reasoned the electron-withdrawing groups on the arene increased the nitrogen–hydrogen bond dissociation energy (BDEN–H) of compound 1 [21], decreasing the rate of ring opening. Substrates 1i
PDF
Album
Supp Info
Full Research Paper
Published 25 Feb 2019

Study on the regioselectivity of the N-ethylation reaction of N-benzyl-4-oxo-1,4-dihydroquinoline-3-carboxamide

  • Pedro N. Batalha,
  • Luana da S. M. Forezi,
  • Maria Clara R. Freitas,
  • Nathalia M. de C. Tolentino,
  • Ednilsom Orestes,
  • José Walkimar de M. Carneiro,
  • Fernanda da C. S. Boechat and
  • Maria Cecília B. V. de Souza

Beilstein J. Org. Chem. 2019, 15, 388–400, doi:10.3762/bjoc.15.35

Graphical Abstract
  • ), and anticancer activity for three cancer cell lines (Figure 1) [16]. Although it is not a general rule for achieving a bioactive profile, any groups attached to C-3 of the 4-oxoquinoline moiety, especially those containing a hydrogen bond donor group, such as a carboxyl, an acyl hydrazide or a
  • carboxamide group, may contribute to enhance the bioactivity. This fact could be explained by the coplanarity induced by the C-4 carbonyl hydrogen bond interactions with biological targets [3] or complexation with physiological metal cations such as magnesium and zinc [23]. Besides the derivatives 3a and 3b
  • , accomplished through temperatures above 200 °C, any derivatization afterwards should maintain the N–H carboxamide group intact, in order to provide the hydrogen bond donor group attached to C-3, which is usually related to the bioactivity of such compounds (Figure 2) [3][23]. In order to obtain an N1-alkylated
PDF
Album
Supp Info
Full Research Paper
Published 12 Feb 2019
Other Beilstein-Institut Open Science Activities