Journal articles on the topic 'Dihydrogen Ligands'

To see the other types of publications on this topic, follow the link: Dihydrogen Ligands.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Dihydrogen Ligands.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Jacobsen, Heiko. "Localized-orbital locator (LOL) profiles of transition-metal hydride and dihydrogen complexes,." Canadian Journal of Chemistry 87, no. 7 (July 2009): 965–73. http://dx.doi.org/10.1139/v09-060.

Full text
Abstract:
A bond descriptor based on the kinetic-energy density, the localized-orbital locator (LOL), is used to characterize the nature of the chemical bond in transition-metal hydride and dihydrogen complexes. Cationic complexes of the iron triad [MH3(PMe3)4]+ (M = Fe, Ru, Os) serve as model compounds for transition-metal hydrogen bonding, since these complexes not only present examples for hydride as well as dihydrogen complexes, but for certain representatives, the two different types of metal–hydrogen bonds are realized within the same molecule. Both types of ligands show characteristic LOL profiles: (3,–3) Γ attractors in close vicinity to the H-atom for hydride ligands, and (3,–3) Γ attractors located between the two atoms for a dihydrogen ligand with νΓ-values of 0.8 and 0.9, respectively. In-between structures combine elements of the hydride and dihydrogen ligands. Relativistic effects on the relative stability of various isomers for the set of model compounds have been evaluated.
APA, Harvard, Vancouver, ISO, and other styles
2

Li, Gang, Deven P. Estes, Jack R. Norton, Serge Ruccolo, Aaron Sattler, and Wesley Sattler. "Dihydrogen Activation by Cobaloximes with Various Axial Ligands." Inorganic Chemistry 53, no. 19 (September 18, 2014): 10743–47. http://dx.doi.org/10.1021/ic501975r.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Jagirdar, Balajir, and Nisha Mathew. "Chemistry of dihydrogen complexes containing only phosphorus co-ligands." Journal of Chemical Sciences 114, no. 4 (August 2002): 285–89. http://dx.doi.org/10.1007/bf02703821.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Bardají, Manuel, Anne-Marie Caminade, Jean-Pierre Majoral, and Bruno Chaudret. "Ruthenium Hydride and Dihydrogen Complexes with Dendrimeric Multidentate Ligands." Organometallics 16, no. 15 (July 1997): 3489–97. http://dx.doi.org/10.1021/om970092+.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Cano, Israel, Luis M. Martínez-Prieto, and Piet W. N. M. van Leeuwen. "Heterolytic cleavage of dihydrogen (HCD) in metal nanoparticle catalysis." Catalysis Science & Technology 11, no. 4 (2021): 1157–85. http://dx.doi.org/10.1039/d0cy02399j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Ghosh, Shishir, Katherine B. Holt, Shariff E. Kabir, Michael G. Richmond, and Graeme Hogarth. "Electrocatalytic proton reduction catalysed by the low-valent tetrairon-oxo cluster [Fe4(CO)10(κ2-dppn)(μ4-O)]2− [dppn = 1,1′-bis(diphenylphosphino)naphthalene]." Dalton Transactions 44, no. 11 (2015): 5160–69. http://dx.doi.org/10.1039/c4dt03323j.

Full text
Abstract:
[Fe4(CO)102-dppn)(μ4-O)]2− reduces protons and DFT calculations support the sequential formation of hydride and dihydrogen ligands at the unique iron centre.
APA, Harvard, Vancouver, ISO, and other styles
7

Avdeeva, Varvara V., Anna V. Vologzhanina, Elena A. Malinina, and Nikolai T. Kuznetsov. "Dihydrogen Bonds in Salts of Boron Cluster Anions [BnHn]2− with Protonated Heterocyclic Organic Bases." Crystals 9, no. 7 (June 28, 2019): 330. http://dx.doi.org/10.3390/cryst9070330.

Full text
Abstract:
Dihydrogen bonds attract much attention as unconventional hydrogen bonds between strong donors of H-bonding and polyhedral (car)borane cages with delocalized charge density. Salts of closo-borate anions [B10H10]2− and [B12H12]2− with protonated organic ligands 2,2’-dipyridylamine (BPA), 1,10-phenanthroline (Phen), and rhodamine 6G (Rh6G) were selectively synthesized to investigate N−H...H−B intermolecular bonding. It was found that the salts contain monoprotonated and/or diprotonated N-containing cations at different ratios. Protonation of the ligands can be implemented in an acidic medium or in water because of hydrolysis of metal cations resulting in the release of H3O+ cations into the reaction solution. Six novel compounds were characterized by X-ray diffraction and FT-IR spectroscopy. It was found that strong dihydrogen bonds manifest themselves in FT-IR spectra that allows one to use this technique even in the absence of crystallographic data.
APA, Harvard, Vancouver, ISO, and other styles
8

Freitag, Kerstin, Mariusz Molon, Paul Jerabek, Katharina Dilchert, Christoph Rösler, Rüdiger W. Seidel, Christian Gemel, Gernot Frenking, and Roland A. Fischer. "Zn⋯Zn interactions at nickel and palladium centers." Chemical Science 7, no. 10 (2016): 6413–21. http://dx.doi.org/10.1039/c6sc02106a.

Full text
Abstract:
Zinc–zinc interactions on nickel and palladium centers are highly dependent on the co-ligands. These dependencies are also found for the formation of dihydrogen vs. dihydride complexes and underline the analogy [Zn2Cp*2] ↔ H2.
APA, Harvard, Vancouver, ISO, and other styles
9

Barthazy, Peter, Diego Broggini, and Antonio Mezzetti. "Making a 16-electron bromo (or iodo) complex of ruthenium(II) and a C—F bond in one pot." Canadian Journal of Chemistry 79, no. 5-6 (May 1, 2001): 904–14. http://dx.doi.org/10.1139/v01-049.

Full text
Abstract:
The 16e– bromo or iodo complexes [RuX(dppp)2]+ (dppp = 1,3-bis(diphenylphosphino)propane, X = Br (1c), I (1d)) and [RuX(dppe)2]+ (dppe = 1,2-bis(diphenylphosphino)ethane, X = Br (2c), I (2d)) have been prepared exploiting the reaction of the fluoro complexes [RuF(dppp)2]+ (1a) and [Tl(µ-F)2Ru(dppe)2]+ (3) with activated alkyl bromides or iodides. The X-ray structures of 1c, 1d, 2c, and 2d suggest that the distortion of the Y-shaped trigonal-bipyramidal structure of [MX(P∩P)2]+ is possibly related to the formation of intramolecular hydrogen bonds between the halide ligand and the ortho-hydrogen atoms of the neighbouring phenyl rings. The five-coordinate species 1c, 1d, 2c, and 2d react with H2 to form the dihydrogen complexes [RuX(η2-H2)(P∩P)2]+. The reaction of the dppp derivatives 1c and 1d with H2 (P = 1 atm, 1 atm = 101.322 kPa) is an equilibrium. Quantitative formation of [RuBr(η2-H2)(dppp)2] (4c) is obtained under H2 pressure (100 bar, 1 bar = 100 kPa), whereas the iodo analogue is not stable under analogous conditions. The less crowded dppe derivatives 2c and 2d react quantitatively with H2 under ambient pressure. The iodo and bromo derivatives [RuX(η2-H2)(P∩P)2]+ contain elongated dihydrogen ligands, as indicated by their transverse relaxation times T1 (min). The present data suggest that Cl, Br, and I have similar donor properties in these dihydrogen complexes, and that the different chemical behaviour in the Cl, Br, I series is mainly a result of steric effects.Key words: 16e– complexes, ruthenium, fluoro complexes, bromo complexes, iodo complexes, dihydrogen complexes.
APA, Harvard, Vancouver, ISO, and other styles
10

Eckert, J. "Interconversion of dihydrogen and hydride ligands in transition metal complexes." Acta Crystallographica Section A Foundations of Crystallography 58, s1 (August 6, 2002): c219. http://dx.doi.org/10.1107/s0108767302093765.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Polezhaev, Alexander V., Mariam G. Ezernitskaya, and Avthandil A. Koridze. "Dihydrogen and dinitrogen rhodium complexes bearing metallocene-based pincer ligands." Inorganica Chimica Acta 496 (October 2019): 118844. http://dx.doi.org/10.1016/j.ica.2019.03.039.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Ringenberg, Mark R., Mark J. Nilges, Thomas B. Rauchfuss, and Scott R. Wilson. "Oxidation of Dihydrogen by Iridium Complexes of Redox-Active Ligands." Organometallics 29, no. 8 (April 26, 2010): 1956–65. http://dx.doi.org/10.1021/om9010593.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Hu, Mao-Lin, Hong-Ping Xiao, Shun Wang, and Xin-Hua Li. "catena-Poly[[(1,10-phenanthroline-κ2 N,N′)copper(II)]-μ-(dihydrogen benzene-1,2,4,5-tetracarboxylato)-κ2 O 1:O 4]." Acta Crystallographica Section C Crystal Structure Communications 59, no. 11 (October 11, 2003): m454—m455. http://dx.doi.org/10.1107/s0108270103021097.

Full text
Abstract:
In the title compound, [Cu(C10H4O8)(C12H8N2)] n , the CuII cation has a four-coordination environment completed by two N atoms from one 1,10-phenanthroline (phen) ligand and two O atoms belonging to two dihydrogen benzene-1,2,4,5-tetracarboxylate anions (H2TCB2−). There is a twofold axis passing through the CuII cation and the centre of the phen ligand. The [Cu(phen)]2+ moieties are bridged by H2TCB2− anions to form an infinite one-dimensional coordination polymer with a zigzag chain structure along the c axis. A double-chain structure is formed by hydrogen bonds between adjacent zigzag chains. Furthermore, there are π–π stacking interactions between the phen ligands, with an average distance of 3.64 Å, resulting in a two-dimensional network structure.
APA, Harvard, Vancouver, ISO, and other styles
14

Lauricella, Marco, Letizia Chiodo, Giovanni Ciccotti, and Alberto Albinati. "Ab initio accelerated molecular dynamics study of the hydride ligands in the ruthenium complex: Ru(H2)2H2(P(C5H9)3)2." Physical Chemistry Chemical Physics 21, no. 45 (2019): 25247–57. http://dx.doi.org/10.1039/c9cp03776d.

Full text
Abstract:
The dihydrogen complex Ru(H2)2H2(P(C5H9)3)2 has been investigated, via ab initio accelerated molecular dynamics, to elucidate the H ligands dynamics and possible reaction paths for H2/H exchange.
APA, Harvard, Vancouver, ISO, and other styles
15

Bautista, Maria Teresa, Kelly Anne Earl, Patricia Anne Maltby, Robert Harold Morris, and Caroline Theresia Schweitzer. "New dihydrogen complexes: the synthesis and spectroscopic properties of iron(II), ruthenium(II), and osmium(II) complexes containing the meso-tetraphos-1 ligand." Canadian Journal of Chemistry 72, no. 3 (March 1, 1994): 547–60. http://dx.doi.org/10.1139/v94-078.

Full text
Abstract:
The synthesis and properties of dihydrogen complexes trans-[MH(H2)L]+, M = Fe, Ru, Os, which contain the ligand meso-tetraphos-1, S,R-Ph2PCH2CH2P(Ph)CH2CH2P(Ph)CH2CH2PPh2 (L) are described. There are interesting possibilities of isomerism in such trans complexes because the axial binding sites at the metal are different, one being surrounded by four phenyl groups and the other by two phenyl groups. The osmium complex is prepared in an unusual reaction of cis-β-Os(Cl)2L with H2 (1 atm) and NaBPh4 (1 mol) in THF or by the reaction of trans-OsH(Cl)L with NaBPh4 and H2. The iron and ruthenium complexes were made by a reaction of HBF4 with complexes trans-M(H)2L that have inequivalent trans hydrides. The ruthenium complex was also prepared starting from isomers of trans-RuH(Cl)L. The H—H distance in the rapidly spinning dihydrogen ligand has been calculated from T1(min) data to be 0.88, 0.89, and 0.99 Å for the complexes of Fe, Ru, Os, respectively. The presence of the H—D bond in the isotopomers trans-[MH(HD)L]+ and trans-[MD(HD)L]+ is also confirmed by the observation of 1JHD coupling constants of 32, 33.5, and 26.4 Hz for Fe, Ru, and Os, respectively. There is no rapid intramolecular H atom exchange in these complexes in contrast to those with di-tert-phosphine ligands like [MH(H2)(dppe)2]+ or to the trihydride Re(H)3L. Described also are the properties of the precursor complexes including cis-β- and trans-Ru(Cl)2L and derivatives of the dihydrogen complexes trans-[MH(L′)L]+, L′ = CH3CN (on Ru and Os), PMe2Ph (on Ru), and CO (on Os). Trends in the NMR properties of isostructural complexes are reported.
APA, Harvard, Vancouver, ISO, and other styles
16

Grellier, Mary, Laure Vendier, and Sylviane Sabo-Etienne. "Ruthenium Complexes Carrying Hydride, Dihydrogen, and Phosphine Ligands: Reversible Hydrogen Release." Angewandte Chemie 119, no. 15 (March 6, 2007): 2667–69. http://dx.doi.org/10.1002/ange.200605038.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Grellier, Mary, Laure Vendier, and Sylviane Sabo-Etienne. "Ruthenium Complexes Carrying Hydride, Dihydrogen, and Phosphine Ligands: Reversible Hydrogen Release." Angewandte Chemie International Edition 46, no. 15 (April 2, 2007): 2613–15. http://dx.doi.org/10.1002/anie.200605038.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Morris, R. H. "ChemInform Abstract: The Chemistry of the Dihydrogen Ligand in Transition-Metal Compounds with Sulfur Donor Ligands." ChemInform 30, no. 14 (June 16, 2010): no. http://dx.doi.org/10.1002/chin.199914318.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Donald, Steven M. A., Anton Vidal-Ferran, and Feliu Maseras. "A DFT/MM analysis of the effect of ligand substituents on asymmetric hydrogenation catalyzed by rhodium complexes with phosphine–phosphinite ligands." Canadian Journal of Chemistry 87, no. 10 (October 2009): 1273–79. http://dx.doi.org/10.1139/v09-051.

Full text
Abstract:
DFT and DFT/MM calculations are carried out on the rate-determining step of the addition of dihydrogen to methyl-(N)-acetylaminoacrylate catalyzed by a rhodium catalyst containing a bidentate phosphine–phosphinite ligand. DFT calculations reproduce the experimental results, while DFT/MM calculations do not. The failure of DFT/MM methods for this particular problem is analyzed through a series of calculations with different partitions between the DFT and MM regions, which show that electronic effects of all ligand substituents considered are critical. The analysis of these electronic effects provides key information on the role of each of the substituents in the outcome of the overall catalytic process.
APA, Harvard, Vancouver, ISO, and other styles
20

Cherni, S., F. Zid, and A. Driss. "Crystal structure of a [(dihydrogen pyrophosphato-K2O,O') bis(1,10-phenanthroline-N,N')nickel(II)]2.5-hydrate [Ni(H2P2O7)(C12H8N2)2] · 2.5H2O." Журнал структурной химии 57, no. 8 (2016): 1769. http://dx.doi.org/10.26902/jsc20160825.

Full text
Abstract:
Coordination nickel pyrophosphate [Ni(H2P2O7)(C12H8N2)2]×2.5H2O (I) is hydrothermally synthesized and characterized by single crystal X-ray diffraction. The title compound crystallizes in the triclinic system, space group P-1, with cell parameters M = 640.07, a = 10.285(2) Å, b = 10.510(3) Å, c = 12.775(3) Å, α = 88.06(2)°, β = 77.87(2)°, γ = 89.26(2)°, V = 1349.2(5) Å3, Z = 2, R1[I > 2σ(I)] = 0.0438, wR2[I > 2σ(I)] = 0.1244. This compound displays a new structure of ladder-like 2D layers parallel to (010) consisting of [Ni(H2P2O7)(phen)2] entities with the distorted octahedral NiN4O2 coordination geometry arising from two chelating 1,10-phe­nanthroline ligands and diphosphate [H2P2O7] ligand bridged through π⋯π stacking interactions between the neighboring 1,10-phen ligands with interplanar distances of 4.425 Å and 4.525 Å. In the compound, the phen ligands bind in a bidentate fashion to the metal atoms and the ladder-like structure of the compound extends into a three-dimensional supramolecular array via hydrogen bonds (O4—H17…O5) between diphosphate groups, which delimits b axis tunnels where water molecules are located.
APA, Harvard, Vancouver, ISO, and other styles
21

Saad, Ahlem Ben, Ahmed Selmi, Mohamed Rzaigui, and Samah Toumi Akriche. "Bis(2,6-dimethylanilinium) diaquabis(dihydrogen diphosphato-κ2O,O′)cobaltate(II)." Acta Crystallographica Section E Structure Reports Online 70, no. 3 (February 8, 2014): m86—m87. http://dx.doi.org/10.1107/s1600536814002530.

Full text
Abstract:
In the title compound, (C8H12N)2[Co(H2P2O7)2(H2O)2], the Co2+ion lies on a crystallographic inversion centre and adopts a slightly distorted octahedral CoO6coordination geometry arising from two chelating diphosphate [H2P2O7]2−ligands and twotranswater molecules. In the crystal, the components are linked by O—H...O, N—H...O and C—H...O hydrogen bonds and weak aromatic π–π stacking [shortest centroid–centroid separation = 3.778 (2) Å] interactions. (001) layers of alternating organic cations and complex inorganic anions are apparent.
APA, Harvard, Vancouver, ISO, and other styles
22

Vendier, L., M. Grellier, and S. Sabo-Etienne. "Structural study of hydride and dihydrogen ligands ruthenium complexes: reversible hydrogen release." Acta Crystallographica Section A Foundations of Crystallography 64, a1 (August 23, 2008): C403. http://dx.doi.org/10.1107/s0108767308087084.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Mala, Deep, Balaji R. Jagirdar, Yogesh P. Patil, and Munirathinam Nethaji. "Homobimetallic hydride and dihydrogen complexes of ruthenium bearing N-heterocyclic carbene ligands." Journal of Organometallic Chemistry 830 (February 2017): 203–11. http://dx.doi.org/10.1016/j.jorganchem.2016.12.025.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Fryzuk, Michael D., Warren E. Piers, Frederick W. B. Einstein, and Terry Jones. "Coordinatively unsaturated binuclear clusters of rhodium. The reactivity of [{Pri2P(CH2)nPPri2}Rh]2(μ-H)2 (n = 2, 3, and 4) with dihydrogen, and their use in the catalytic hydrogenation of olefins." Canadian Journal of Chemistry 67, no. 5 (May 1, 1989): 883–96. http://dx.doi.org/10.1139/v89-137.

Full text
Abstract:
The coordinatively unsaturated binuclear rhodium hydrides [{Pri2P(CH2)nPPri2}Rh]2(μ-H)2 (n = 2: 1b; n = 3: 1c; n = 4: 1d) react rapidly with dihydrogen to yield the fluxional binuclear tetrahydrides [{Pri2P(CH2)nPPri2}RhH]((μ-H)3[Rh{Pri2P(CH2)nPPri2}] 2b–d(n = 2–4). The dihydride 1b was structurally characterized by single crystal X-ray diffraction. While the tetrahydrides 2b and 2c were found to be stable only in solution in the presence of excess dihydrogen, 2d was stable in the solid state. Three separate exchange processes were characterized for 2b–d via 1H and 31P{1H} NMR spectroscopy: intermolecular exchange with free dihydrogen, and two intramolecular processes exchanging the four hydride ligands. The geometry about the rhodium centres in the limiting structures involved in these processes is affected by the chelate ring size of the ancillary diphosphine ligand. A "rocking" process observed at low temperatures for each of these tetrahydrides has a ΔG≠ of <7.0(5) kcal/mol, 8.6(5) kcal/mol, and 11.5(5) kcal/mol for 2b, 2c, and 2d, respectively. The dihydrides 1b and 1c were found to catalyze the hydrogenation of 1-hexene. Turnover numbers of 850–950 h−1 and 700 h−1 were observed for 1b and 1c, respectively, along with isomerization side reactions. Catalyst concentration studies on the hydrogenation of styrene using 1c as the catalyst precursor revealed a decrease in turnover frequency with increasing total metal concentration, suggesting that the active catalyst is a mononuclear species. Chemical evidence suggests that a pathway involving binuclear intermediates is also available, but that in the 1c system at least, a pathway utilizing mononuclear species as intermediates predominates. Keywords: binuclear rhodium hydrides, homogeneous catalysis.
APA, Harvard, Vancouver, ISO, and other styles
25

Elboulali, Adel, Samah Akriche, and Mohamed Rzaigui. "Bis(2-methoxybenzylammonium) diaquabis(dihydrogen diphosphato-κ2O,O′)manganate(II) dihydrate." Acta Crystallographica Section E Structure Reports Online 69, no. 11 (October 2, 2013): m572. http://dx.doi.org/10.1107/s1600536813026366.

Full text
Abstract:
The asymmetric unit of the title compound, (C8H12NO)2[Mn(H2P2O7)2(H2O)2]·2H2O, consists of half an MnIIcomplex anion, a 2-methoxybenylammonium cation and a solvent water molecule. The MnIIcomplex anion lies across an inversion center, and has a slightly distorted octahedral coordination environment for the MnIIion, formed by two bidentate dihydrogendiphosphate ligands and two water molecules. In the crystal, the components are linked by O—H...O and N—H...O hydrogen bonds, forming layers parallel to (100). An intramolecular N—H...O hydrogen bond is also observed.
APA, Harvard, Vancouver, ISO, and other styles
26

Elboulali, Adel, Ahmed Selmi, Nicolas Ratel-Ramond, Mohamed Rzaigui, and Samah Toumi Akriche. "Bis(2-methoxybenzylammonium) diaquabis(dihydrogen diphosphato-κ2O,O′)cobaltate(II) dihydrate." Acta Crystallographica Section E Structure Reports Online 70, no. 4 (March 26, 2014): m145—m146. http://dx.doi.org/10.1107/s1600536814006102.

Full text
Abstract:
The title compound, (C8H12NO)2[Co(H2P2O7)2(H2O)2]·2H2O, crystallizes isotypically with its MnIIanalogue. It consists of alternating layers of organic cations and inorganic complex anions, extending parallel to (100). The complex cobaltate(II) anion exhibits -1 symmetry. Its Co2+atom has an octahedral coordination sphere, defined by two water molecules in apical positions and two H2P2O72−ligands in equatorial positions. The cohesion between inorganic and organic layers is accomplished by a set of O—H...O and N—H...O hydrogen bonds involving the organic cation, the inorganic anion and the remaining lattice water molecules.
APA, Harvard, Vancouver, ISO, and other styles
27

Eckert, J. "Inelastic neutron scattering studies of dihydrogen and hydride ligands in transition metal complexes." Acta Crystallographica Section A Foundations of Crystallography 52, a1 (August 8, 1996): C29. http://dx.doi.org/10.1107/s0108767396097826.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Arliguie, Therese, Bruno Chaudret, Robert H. Morris, and Andrea Sella. "Monomeric and dimeric ruthenium(II) .eta.2-dihydrogen complexes with tricyclohexylphosphine co-ligands." Inorganic Chemistry 27, no. 4 (February 1988): 598–99. http://dx.doi.org/10.1021/ic00277a006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Bakhmutov, Vladimir I. "Proton Transfer to Hydride Ligands with Formation of Dihydrogen Complexes: A Physicochemical View." European Journal of Inorganic Chemistry 2005, no. 2 (January 2005): 245–55. http://dx.doi.org/10.1002/ejic.200400697.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Klein, Hans-Friedrich, Stefan Haller, Hongjian Sun, Xiaoyan Li, Thomas Jung, Caroline Röhr, Ulrich Flörke, and Hans-Jürgen Haupt. "Halogeno(acylphenolato)cobalt(III)-Verbindungen mit Trimethylphosphan-Liganden/ Halogeno(acylphenolato)cobalt(III) Compounds Containing Trimethylphosphane Ligands." Zeitschrift für Naturforschung B 53, no. 8 (August 1, 1998): 856–64. http://dx.doi.org/10.1515/znb-1998-0814.

Full text
Abstract:
Abstract Complexes mer-CoH(CO-CR=CR′-O)(PMe3)3 react with haloalkanes RX (X = Br, I) or with acids HX (X = Cl, Br) under elimination of dihydrogen. In both reactions a change of configu­ration at the metal is brought about by directional steering through the hard/soft (acyl)enolato chelate ligands to form octahedral complexes mer-CoX(CO-CR=CR′-C))(PMe3)3 or sterically crowded ionic compounds [Co(CO-CR=CR′-O)(PMe3)4(3)]+ X-(X = ClO4) without opening of the (acyl)enolato chelate ring.
APA, Harvard, Vancouver, ISO, and other styles
31

Yang, Xiuxiu, Thomas L. Gianetti, Michael D. Wörle, Nicolaas P. van Leest, Bas de Bruin, and Hansjörg Grützmacher. "A low-valent dinuclear ruthenium diazadiene complex catalyzes the oxidation of dihydrogen and reversible hydrogenation of quinones." Chemical Science 10, no. 4 (2019): 1117–25. http://dx.doi.org/10.1039/c8sc02864h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Faulkner, Robert A., Nathan J. Patmore, Craig R. Rice, and Christopher Slater. "Dihydrogen phosphate-containing dinuclear double assemblies that demonstrate phosphate reactivity to the tetrafluoroborate anion." Chemical Communications 54, no. 66 (2018): 9159–62. http://dx.doi.org/10.1039/c8cc04900a.

Full text
Abstract:
Ligands L1 and L2 both form dinuclear assemblies with Cu(ii) and these react with dihydrogen phosphate to give [Cu2L2(H2PO4)]3+. However, in the presence of tetrafluoroborate anions the phosphate undergoes reaction with the anion forming [Cu3(L1)3(O3POBF3)]3+ and [Cu2(L2)2(O2P(OBF3)2)]+.
APA, Harvard, Vancouver, ISO, and other styles
33

Bolaño, Tamara, Miguel A. Esteruelas, Israel Fernández, Enrique Oñate, Adrián Palacios, Jui-Yi Tsai, and Chuanjun Xia. "Osmium(II)–Bis(dihydrogen) Complexes ContainingCaryl,CNHC–Chelate Ligands: Preparation, Bonding Situation, and Acidity." Organometallics 34, no. 4 (February 11, 2015): 778–89. http://dx.doi.org/10.1021/om501275c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Bakhmutov, Vladimir I., Evgenii V. Vorontsov, and Alexey B. Vymenits. "Kinetics of H2 Dissociation from Some Iridium-Dihydrogen Complexes with Structurally Different (H2) Ligands." Inorganic Chemistry 34, no. 1 (January 1995): 214–17. http://dx.doi.org/10.1021/ic00105a036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Ogata, Hideaki, Koji Nishikawa, and Wolfgang Lubitz. "Observation of a metal-hydride in [NiFe] hydrogenase." Acta Crystallographica Section A Foundations and Advances 70, a1 (August 5, 2014): C1212. http://dx.doi.org/10.1107/s2053273314087877.

Full text
Abstract:
Hydrogenases catalyze the reversible hydrogen oxidation process by cleaving dihydrogen heterolytically.(1) For this reaction, the enzyme uses the transition metals Ni and Fe, which are abundant in Nature. Standard [NiFe] hydrogenaes are mainly composed of two subunits (total ~90 kDa) The [NiFe] active site is located in the center of the molecule. The active site of [NiFe] hydrogenase is composed of the dinuclear Ni-Fe center, where the Fe ion is coordinated by non-protein ligands (1CO and 2CN¯ ). Two thiolates of cysteine residues are bridging both metals. Furthermore, the Ni is coordinated to the two thiolates of cysteine residues in a terminal fashion. A third bridging ligand is found between the Ni and Fe atom, depending on the redox state.(1) In the inactive form, a third bridging ligand (OH¯¯¯ ) is found between Ni and Fe. Once the enzyme is activated, the bridging position is supposed to be vacant or bridged by a hydride. A previous X-ray crystallographic study at 1.4 Å resolution revealed that the bridging ligand (OH) is removed upon H2 reduction.(2) Electron paramagnetic resonance (EPR) spectroscopy showed that a hydride is located in the bridge between Ni and Fe, which is lost upon illumination at cryogenic temperature.(3) Here we present a crystallographic analysis of the fully reduced (Ni-R) state of [NiFe] hydrogenase from Desulfovibrio vulgaris Miyazaki F at 0.89 Å resolution. The ultra-high resolution analysis revealed the presence of the hydride bridge at the NiFe active site in the catalytically active state. Furthermore the CO and CN ligands could be identified and a protonated thiolate sulfur ligand of the Ni is postulated based on the electron density.
APA, Harvard, Vancouver, ISO, and other styles
36

Schilter, David, Danielle L. Gray, Amy L. Fuller, and Thomas B. Rauchfuss. "Synthetic Models for Nickel–Iron Hydrogenase Featuring Redox-Active Ligands." Australian Journal of Chemistry 70, no. 5 (2017): 505. http://dx.doi.org/10.1071/ch16614.

Full text
Abstract:
The nickel–iron hydrogenase enzymes efficiently and reversibly interconvert protons, electrons, and dihydrogen. These redox proteins feature iron–sulfur clusters that relay electrons to and from their active sites. Reported here are synthetic models for nickel–iron hydrogenase featuring redox-active auxiliaries that mimic the iron–sulfur cofactors. The complexes prepared are NiII(μ-H)FeIIFeII species of formula [(diphosphine)Ni(dithiolate)(μ-H)Fe(CO)2(ferrocenylphosphine)]+ or NiIIFeIFeII complexes [(diphosphine)Ni(dithiolate)Fe(CO)2(ferrocenylphosphine)]+ (diphosphine = Ph2P(CH2)2PPh2 or Cy2P(CH2)2PCy2; dithiolate = –S(CH2)3S–; ferrocenylphosphine = diphenylphosphinoferrocene, diphenylphosphinomethyl(nonamethylferrocene) or 1,1′-bis(diphenylphosphino)ferrocene). The hydride species is a catalyst for hydrogen evolution, while the latter hydride-free complexes can exist in four redox states – a feature made possible by the incorporation of the ferrocenyl groups. Mixed-valent complexes of 1,1′-bis(diphenylphosphino)ferrocene have one of the phosphine groups unbound, with these species representing advanced structural models with both a redox-active moiety (the ferrocene group) and a potential proton relay (the free phosphine) proximal to a nickel–iron dithiolate.
APA, Harvard, Vancouver, ISO, and other styles
37

Heinekey, D. Michael, David A. Fine, T. Gregory P. Harper, and Suzanne T. Michel. "Dinuclear dihydride complexes of iridium: a study of structure and dynamics." Canadian Journal of Chemistry 73, no. 7 (July 1, 1995): 1116–25. http://dx.doi.org/10.1139/v95-138.

Full text
Abstract:
Reaction of the neutral dihydride complexes (η-C5R5)Ir(L)H2(L = P(OPh)3, CO; R = H, Me) with triflic acid (HO3SCF3) at ambient temperature or above affords hydrogen and the dimeric hydride bridged species {[(η-C5R5)Ir(L)H]2(µ-H)}O3SCF3. Deprotonation of the cationic complexes gives the neutral dimers of the form [(η-C5R5)Ir(L)H]2. Spectroscopic data are consistent with the presence of only terminal hydride ligands in these complexes. Variable temperature 1H and 31P NMR studies indicate that a rapid dynamic process exchanges the two terminal hydride ligands and that the complexes exist as unequal mixtures of the racemic and meso diastereomers. Synthesis of a lower symmetry derivative, [(η-C5Me4Et)Ir(CO)H]2, reveals that a rapid epimerization process occurs that interconverts the two diastereomers. Thermolysis of the carbonyl complex [(η-C5Me5)Ir(CO)H]2 affords [(η-C5Me5)Ir(CO)]2 and dihydrogen. Keywords: iridium, hydride, bimetallic, dynamics, epimerization.
APA, Harvard, Vancouver, ISO, and other styles
38

Jiménez, M. Victoria, Ana Ojeda-Amador, Raquel Puerta-Oteo, Joaquín Martínez-Sal, Vincenzo Passarelli, and Jesús Pérez-Torrente. "Selective Oxidation of Glycerol via Acceptorless Dehydrogenation Driven by Ir(I)-NHC Catalysts." Molecules 27, no. 22 (November 8, 2022): 7666. http://dx.doi.org/10.3390/molecules27227666.

Full text
Abstract:
Iridium(I) compounds featuring bridge-functionalized bis-NHC ligands (NHC = N-heterocyclic carbene), [Ir(cod)(bis-NHC)] and [Ir(CO)2(bis-NHC)], have been prepared from the appropriate carboxylate- or hydroxy-functionalized bis-imidazolium salts. The related complexes [Ir(cod)(NHC)2]+ and [IrCl(cod)(NHC)(cod)] have been synthesized from a 3-hydroxypropyl functionalized imidazolium salt. These complexes have been shown to be robust catalysts in the oxidative dehydrogenation of glycerol to lactate (LA) with dihydrogen release. High activity and selectivity to LA were achieved in an open system under low catalyst loadings using KOH as a base. The hydroxy-functionalized bis-NHC catalysts are much more active than both the carboxylate-functionalized ones and the unbridged bis-NHC iridium(I) catalyst with hydroxyalkyl-functionalized NHC ligands. In general, carbonyl complexes are more active than the related 1,5-cyclooctadiene ones. The catalyst [Ir(CO)2{(MeImCH2)2CHOH}]Br exhibits the highest productivity affording TONs to LA up to 15,000 at very low catalyst loadings.
APA, Harvard, Vancouver, ISO, and other styles
39

Esteruelas, Miguel A., Luis A. Oro, and Natividad Ruiz. "Synthesis of the first metal dihydrogen M(.eta.2-H2) complexes containing sulfur-donor ligands." Inorganic Chemistry 32, no. 17 (August 1993): 3793–94. http://dx.doi.org/10.1021/ic00069a043.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Picot, Alexandre, Hellen Dyer, Antoine Buchard, Audrey Auffrant, Laure Vendier, Pascal Le Floch, and Sylviane Sabo-Etienne. "Interplay between Hydrido/Dihydrogen and Amine/Amido Ligands in Ruthenium-Catalyzed Transfer Hydrogenation of Ketones." Inorganic Chemistry 49, no. 4 (February 15, 2010): 1310–12. http://dx.doi.org/10.1021/ic902339j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Mendes, Ricardo F., Nutalapati Venkatramaiah, João P. C. Tomé, and Filipe A. Almeida Paz. "Crystal structure of a compact three-dimensional metal–organic framework based on Cs+and (4,5-dicyano-1,2-phenylene)bis(phosphonic acid)." Acta Crystallographica Section E Crystallographic Communications 72, no. 12 (November 15, 2016): 1794–98. http://dx.doi.org/10.1107/s2056989016016765.

Full text
Abstract:
A new metal–organic framework compound, poly[[μ7-dihydrogen (4,5-dicyano-1,2-phenylene)diphosphonato](oxonium)caesium], [Cs(C8H4N2O6P2)(H3O)]n(I), based on Cs+and the organic linker 4,5-dicyano-1,2-phenylene)bis(phosphonic acid, (H4cpp), containing two distinct coordinating functional groups, has been prepared by a simple diffusion method and its crystal structure is reported. The coordination polymeric structure is based on a CsO8N2complex unit comprising a monodentate hydronium cation, seven O-atom donors from two phosphonium groups of the (H2cpp)2−ligand, and two N-atom donors from bridging cyano groups. The high level of connectivity from both the metal cation and the organic linker allow the formation of a compact and dense three-dimensional network without any crystallization solvent. Topologically (I) is a seven-connected uninodal network with an overall Schäfli symbol of {417.64}. Metal cations form an undulating inorganic layer, which is linked by strong and highly directional O—H...O hydrogen-bonding interactions. These metallic layers are, in turn, connected by the organic ligands along the [010] direction to form the overall three-dimensional framework structure.
APA, Harvard, Vancouver, ISO, and other styles
42

Wu, Biao, Cuirong Huo, Shaoguang Li, Yanxia Zhao, and Xiao-Juan Yang. "Anion Coordination of Bis-bisurea Ligands: Aggregation of Dihydrogen Phosphate Anion into Oligomers and Infinite Chains." Zeitschrift für anorganische und allgemeine Chemie 641, no. 10 (June 29, 2015): 1786–91. http://dx.doi.org/10.1002/zaac.201500243.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Morris, Robert H., Kelly A. Earl, Rudy L. Luck, Natalie J. Lazarowych, and Andrea Sella. "Dihydrogen vs. dihydride. Correlations between electrochemical or UV PES data and force constants for carbonyl or dinitrogen ligands in octahedral, d6 complexes and their use in explaining the behavior of the dihydrogen ligand." Inorganic Chemistry 26, no. 16 (August 1987): 2674–83. http://dx.doi.org/10.1021/ic00263a024.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Hampton, Cashman R. S. M., Ian R. Butler, William R. Cullen, Brian R. James, Jean Pierre Charland, and J. Simpson. "Molecular dihydrogen and hydrido derivatives of ruthenium(II) complexes containing chelating ferrocenyl-based tertiary phosphine amine ligands and/or monodentate tertiary phosphine ligands." Inorganic Chemistry 31, no. 26 (December 1992): 5509–20. http://dx.doi.org/10.1021/ic00052a029.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Trowbridge, Logan, Boris Averkiev, and Peter E. Sues. "Palladium complexes bearing calixpyrrole ligands with pendant hydrogen bond donors: Synthesis, structural characterization, electrochemistry and dihydrogen evolution." Polyhedron 225 (October 2022): 116046. http://dx.doi.org/10.1016/j.poly.2022.116046.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Lam, Yat Fai, Chuanqi Yin, Chi Hung Yeung, Siu Man Ng, Guochen Jia, and Chak Po Lau. "Attenuation of Intramolecular Ru−H···H−N Dihydrogen Bonding in Aminocyclopentadienyl Ruthenium Hydride Complexes Containing Phosphite Ligands." Organometallics 21, no. 9 (April 2002): 1898–902. http://dx.doi.org/10.1021/om010966z.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Albertin, Gabriele, Stefano Antoniutti, Marco Bortoluzzi, and Gianluigi Zanardo. "Synthesis and reactivity of hydride and dihydrogen complexes of ruthenium with tris(pyrazolyl)borate and phosphite ligands." Journal of Organometallic Chemistry 690, no. 7 (March 2005): 1726–38. http://dx.doi.org/10.1016/j.jorganchem.2005.01.028.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Dutta, Saikat, Balaji R. Jagirdar, and Munirathinam Nethaji. "Influence of the Electronics of the Phosphine Ligands on the H−H Bond Elongation in Dihydrogen Complexes." Inorganic Chemistry 47, no. 2 (January 2008): 548–57. http://dx.doi.org/10.1021/ic7016769.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Mathew, Nisha, and Balaji R. Jagirdar. "Observation of a Large Coupling of a Bound Dihydrogen Ligand to Phosphorus Ligands intrans-[(dppe)2Ru(η2-H2)(PF(OMe)2)][BF4]2Complex†." Inorganic Chemistry 39, no. 23 (November 2000): 5404–6. http://dx.doi.org/10.1021/ic000419q.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Baya, Miguel, Miguel A. Esteruelas, Montserrat Oliván, and Enrique Oñate. "Monocationic Trihydride and Dicationic Dihydride−Dihydrogen and Bis(dihydrogen) Osmium Complexes Containing Cyclic and Acyclic Triamine Ligands: Influence of the N−Os−N Angles on the Hydrogen−Hydrogen Interactions." Inorganic Chemistry 48, no. 6 (March 16, 2009): 2677–86. http://dx.doi.org/10.1021/ic8023259.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography