To see the other types of publications on this topic, follow the link: Metal hydride bond.

Journal articles on the topic 'Metal hydride bond'

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 'Metal hydride bond.'

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

Ziegler, Tom, and Jian Li. "Bond energies for cationic bare metal hydrides of the first transition series: a challenge to density functional theory." Canadian Journal of Chemistry 72, no. 3 (1994): 783–89. http://dx.doi.org/10.1139/v94-104.

Full text
Abstract:
Density functional methods based on the Local Density Approximation (LDA) and its nonlocal extensions (LDA/NL) are used to calculate the bond energies, as well as the bond lengths and vibrational frequencies of the high spin, open shell first-row transition metal hydride cations MH+. The D298(M+—H) LDA/NL bond energies are in good agreement with experiment for the early transition metals with errors within 5 kcal/mol. However, the error increases to 6–l3 kcal/mol for the late metal hydrides. An analysis based on atomic properties such as 4s ionization potentials and 4s to 3d promotion energies revealed that the large error in bonding energies among the late transition metals can be attributed to an overestimation of the exchange energy in the DFT schemes. It is shown that a simple remedy, based on a thermodynamic cycle, can improve the agreement between experimental and theoretical bond energies. However, simple cationic bare metal complexes such as MH+ remains a challenge to approximate DFT.
APA, Harvard, Vancouver, ISO, and other styles
2

Jacobsen, Heiko. "Localized-orbital locator (LOL) profiles of transition-metal hydride and dihydrogen complexes,." Canadian Journal of Chemistry 87, no. 7 (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
3

Verma, Kanupriya, and K. S. Viswanathan. "The borazine dimer: the case of a dihydrogen bond competing with a classical hydrogen bond." Physical Chemistry Chemical Physics 19, no. 29 (2017): 19067–74. http://dx.doi.org/10.1039/c7cp04056c.

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

Shalimov, Yuri N., Igor K. Shuklin, Vladimir I. Parfenyuk, Vladimir I. Korolkov, Alexander V. Russu, and Vladlen I. Kudryash. "INVESTIGATION OF EFFECTS OF HEAT RELEASE IN ELECTROCHEMICAL SYSTEMS AND THEIR USE IN TECHNOLOGIES FOR PRODUCTION OF ENERGY-INTENSIVE SOURCES FOR AIRCRAFT." IZVESTIYA VYSSHIKH UCHEBNYKH ZAVEDENIY KHIMIYA KHIMICHESKAYA TEKHNOLOGIYA 62, no. 1 (2019): 46–53. http://dx.doi.org/10.6060/ivkkt.20196201.5798.

Full text
Abstract:
The search for new, more energy-intensive types of fuel for the operation of the power plants of aircraft is the most important task in aviation. The unique fuel that has no analogues is hydrogen. The paper attempts to substantiate the technology of metal hydride hydrogen storage in electrochemical systems based on aluminum and its alloys as the most affordable materials from fossil metals, since the traditional methods based on the use of cylinders and cryostats are not effective in transport systems. It is shown that the volumetric storage of hydrogen in the porous structure of metals with the formation of hydrides on atomic bond defects is maximally suitable for the implementation of the system, eliminating the excessive pressure and the low temperatures. The porous structure of the material provides both a high degree of availability of the electrolyte solution to the electrode for the accumulation of hydrides in the entire volume of the metal, and not only on its surface, but also the conditions for the realization of the reduction effect that excludes the explosive nature of hydrogen extraction. The problem of increasing the temperature in the reaction zone, which sometimes causes a slowdown in the rate of certain stages of the electrochemical process, is considered. Using the example of galvanic chrome plating, it has been established that an increase in the temperature inhibits the process of the reducing of the metallic chromium. Therefore, the detailed account of the thermal effects in the electrochemical system allows us to determine the mechanism of the processes. The work revealed that the thermal effects arising at the cathode determine the kinetics of the hydrogen reduction processes during the formation of a hydride. And the thermal effects at the anode determine the kinetics of the formation of a porous structure in the metal. The authors proposed to use the principle of action associated with the transition to the technologies of the volumetric storage of hydrogen in a solid-phase system based on a metal hydride compound for the formation of a new class of aircraft - diaplan.
APA, Harvard, Vancouver, ISO, and other styles
5

Bullock, R. Morris. "Metal-Hydrogen Bond Cleavage Reactions of Transition Metal Hydrides: Hydrogen Atom, Hydride, and Proton Transfer Reactions." Comments on Inorganic Chemistry 12, no. 1 (1991): 1–33. http://dx.doi.org/10.1080/02603599108018617.

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

Merola, Joseph, та Trang Le Husebo. "μ-Oxido-bis[hydridotris(trimethylphosphane-κP)iridium(III)](Ir—Ir) bis(tetrafluoridoborate) dihydrate". Acta Crystallographica Section E Structure Reports Online 70, № 4 (2014): m122—m123. http://dx.doi.org/10.1107/s160053681400453x.

Full text
Abstract:
The title compound, [Ir2H2O(C3H9P)6](BF4)2·2H2O, was isolated from the reaction between [Ir(COD)(PMe3)3]BF4and H2in water (COD is cycloocta-1,5-diene). The asymmetric unit consists of one IrIIIatom bonded to three PMe3groups, one hydride ligand and half an oxide ligand, in addition to a BF4−counter-ion and one water molecule of hydration. The single oxide ligand bridging two IrIIIatoms is disordered across an inversion center with each O atom having a 50% site occupancy. Each IrIIIatom has three PMe3groups occupying facial positions, with the half-occupancy O atoms, a hydride ligand and an Ir—Ir bond completing the coordination sphere. The Ir—Ir distance is 2.8614 (12) Å, comparable to other iridium(III) metal–metal bonds. Two water molecules hydrogen bond to two BF4−anions in the unit cell.
APA, Harvard, Vancouver, ISO, and other styles
7

Edelbach, Brian L., A. K. Fazlur Rahman, Rene J. Lachicotte, and William D. Jones. "Carbon−Fluorine Bond Cleavage by Zirconium Metal Hydride Complexes." Organometallics 18, no. 16 (1999): 3170–77. http://dx.doi.org/10.1021/om9902481.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Elkind, J. L., and P. B. Armentrout. "Transition-metal hydride bond energies: first and second row." Inorganic Chemistry 25, no. 8 (1986): 1078–80. http://dx.doi.org/10.1021/ic00228a004.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Ding, Wen, та Qiuling Song. "Chemoselective catalytic reduction of conjugated α,β-unsaturated ketones to saturated ketones via a hydroboration/protodeboronation strategy". Organic Chemistry Frontiers 3, № 1 (2016): 14–18. http://dx.doi.org/10.1039/c5qo00289c.

Full text
Abstract:
A novel copper-catalyzed chemoselective reduction of a carbon–carbon double or triple bond to a carbon–carbon single bond on α,β-unsaturated ketones is developed, this reaction proceeds under hydrogen gas or stoichiometric metal hydride free conditions.
APA, Harvard, Vancouver, ISO, and other styles
10

Oh, Changjin, Joëlle Siewe, Thao T. Nguyen та ін. "The electronic structure of a β-diketiminate manganese hydride dimer". Dalton Transactions 49, № 41 (2020): 14463–74. http://dx.doi.org/10.1039/d0dt02842h.

Full text
Abstract:
The absence of a metal–metal multiple bond in a dimeric manganese hydride catalyst supported by β-diketiminate ligands, [(<sup>2,6-iPr2Ph</sup>BDI) Mn(μ-H)]<sub>2</sub>, was investigated with density functional theory in conjunction with experimental evidence.
APA, Harvard, Vancouver, ISO, and other styles
11

Alvarez, M. Angeles, M. Esther García, Daniel García-Vivó, Miguel A. Ruiz та Adrián Toyos. "The doubly-bonded ditungsten anion [W2Cp2(μ-PPh2)(NO)2]−: an entry to the chemistry of unsaturated nitrosyl complexes". Dalton Transactions 45, № 34 (2016): 13300–13303. http://dx.doi.org/10.1039/c6dt02319c.

Full text
Abstract:
The title complex, the first anionic nitrosyl complex featuring a metal-metal double bond, displays substantial nucleophilicity, thus providing synthetic access to different unsaturated hydride and heterometallic derivatives, as well as a variety of electron-precise molecules.
APA, Harvard, Vancouver, ISO, and other styles
12

Labinger, Jay A., and John E. Bercaw. "Metal-hydride and metal-alkyl bond strengths: the influence of electronegativity differences." Organometallics 7, no. 4 (1988): 926–28. http://dx.doi.org/10.1021/om00094a022.

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

BULLOCK, R. M. "ChemInform Abstract: Metal-Hydrogen Bond Cleavage Reactions of Transition Metal Hydrides: Hydrogen Atom, Hydride, and Proton Transfer Reactions." ChemInform 22, no. 46 (2010): no. http://dx.doi.org/10.1002/chin.199146331.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Goodfellow, Alister S., and Michael Bühl. "Hydricity of 3d Transition Metal Complexes from Density Functional Theory: A Benchmarking Study." Molecules 26, no. 13 (2021): 4072. http://dx.doi.org/10.3390/molecules26134072.

Full text
Abstract:
A range of modern density functional theory (DFT) functionals have been benchmarked against experimentally determined metal hydride bond strengths for three first-row TM hydride complexes. Geometries were found to be produced sufficiently accurately with RI-BP86-D3(PCM)/def2-SVP and further single-point calculations with PBE0-D3(PCM)/def2-TZVP were found to reproduce the experimental hydricity accurately, with a mean absolute deviation of 1.4 kcal/mol for the complexes studied.
APA, Harvard, Vancouver, ISO, and other styles
15

Ohanessian, Gilles, and William A. Goddard. "Valence-bond concepts in transition metals: metal hydride diatomic cations." Accounts of Chemical Research 23, no. 11 (1990): 386–92. http://dx.doi.org/10.1021/ar00179a007.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Guyon, Carole, Marie-Christine Duclos, Marc Sutter, Estelle Métay, and Marc Lemaire. "Reductive alkylation of active methylene compounds with carbonyl derivatives, calcium hydride and a heterogeneous catalyst." Organic & Biomolecular Chemistry 13, no. 25 (2015): 7067–75. http://dx.doi.org/10.1039/c5ob00849b.

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

Zeng, Xu, Guo-Dong Yin, Yang-Yuan Zhou, and Jian-Fu Zhao. "New Insight into CO2 Reduction to Formate by In Situ Hydrogen Produced from Hydrothermal Reactions with Iron." Molecules 27, no. 21 (2022): 7371. http://dx.doi.org/10.3390/molecules27217371.

Full text
Abstract:
To reveal the nature of CO2 reduction to formate with high efficiency by in situ hydrogen produced from hydrothermal reactions with iron, DFT calculations were used. A reaction pathway was proposed in which the formate was produced through the key intermediate species, namely iron hydride, produced in situ in the process of hydrogen gas production. In the in situ hydrogenation of CO2, the charge of H in the iron hydride was −0.135, and the Fe–H bond distance was approximately 1.537 Å. A C-H bond was formed as a transition state during the attack of Hδ− on Cδ+. Finally, a HCOO species was formed. The distance of the C-H bond was 1.107 Å. The calculated free energy barrier was 16.43 kcal/mol. This study may provide new insight into CO2 reduction to formate in hydrothermal reactions with metal.
APA, Harvard, Vancouver, ISO, and other styles
18

Ouis, Sakina, Djamil Azzedine Rouag, Lamia Bendjeddou, and Corinne Bailly. "Crystal structure of homodinuclear platinum complex containing a metal–metal bond bridged by hydride and phosphide ligands." Acta Crystallographica Section E Crystallographic Communications 74, no. 7 (2018): 977–80. http://dx.doi.org/10.1107/s205698901800868x.

Full text
Abstract:
In the title compound, μ-diphenylphosphido-μ-hydrido-bis[bromido(triphenylphosphane-κP)platinum(II)] diethyl ether monosolvate, [Pt2Br2(C12H10P)H(C18H15P)2]·C4H10O or [Pt2(μ-H)(μ-PPh2)Br2(PPh3)2]·(C2H5)2O, the PtII atoms are coordinated in a distorted square-planar arrangement, with one hydrido and one phosphido ligand bridging in a trans position. In the lattice, C—H...·O and C—H...π interactions are present. This complex has a total number of 32 electrons, 16 electrons for each PtII atom. One of the Br atoms is disordered over two positions in a 0.92:0.08 ratio.
APA, Harvard, Vancouver, ISO, and other styles
19

Fryzuk, Michael D., Warren E. Piers, and Steven J. Rettig. "Reactions of nitriles with binuclear rhodium hydrides. The stepwise reduction of a carbon–nitrogen triple bond at two metal centres." Canadian Journal of Chemistry 70, no. 9 (1992): 2381–89. http://dx.doi.org/10.1139/v92-301.

Full text
Abstract:
The reaction of simple nitriles, R′CN (R′ = CH3, Ph, o-tol) with the electron-rich binuclear rhodium hydride derivatives [(R2PCH2CH2PR2)Rh]2(μ-H)2 (R = Pri: [(dippe) Rh]2(μ-H)2; R = OPri: [(dipope)Rh]2(μ-H)2) results in the formation of alkylideneimido derivatives [(R2PCH2CH2PR2)Rh]2(μ-H)(μ-N=CHR′), apparently by insertion of the nitrile moiety into a bridging hydride bond; this was confirmed by the reaction of nitriles with the dideuteride [(dippe)Rh]2(μ-D)2, which resulted in the formation of [(dippe)Rh]2(μ-D)(μ-N=CHR′). Further reduction can take place by addition of H2 to generate the corresponding amide hydride derivatives [(dippe)Rh]2(μ-H)(μ-NHCH2R′); this represents an overall stoichiometric reduction of a nitrile to a coordinated amide at a binuclear centre. These same amido-hydride complexes can be accessed by addition of amine to the starting binuclear rhodium hydride derivatives. The X-ray structure of [(Pri2PCH2CH2PPri2)Rh]2(μ-H)(μ-N=CHCH3) was undertaken to confirm the structure of these particular intermediates. Crystals of this material are monoclinic, a = 19.036(2), b = 15.139(1), c = 13.604(1) Å, β = 104.119(7)°, Z = 4, space group P21/c. The structure was solved by heavy atom methods and was refined by full-matrix least-squares procedures to R = 0.038 and Rw = 0.041 for 5814 reflections with I ≥ 3σ(I).
APA, Harvard, Vancouver, ISO, and other styles
20

Belt, Simon T., J. C. Scaiano, and Michael K. Whittlesey. "Determination of metal-hydride and metal-ligand (L = CO, N2) bond energies using photoacoustic calorimetry." Journal of the American Chemical Society 115, no. 5 (1993): 1921–25. http://dx.doi.org/10.1021/ja00058a043.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Kuriakose, Nishamol, and Kumar Vanka. "Can main group systems act as superior catalysts for dihydrogen generation reactions? A computational investigation." Dalton Transactions 45, no. 14 (2016): 5968–77. http://dx.doi.org/10.1039/c5dt01058f.

Full text
Abstract:
The density functional theory (DFT) calculations reveal the potential of newly proposed main group germanium hydride systems to effect important chemical transformations, such as the catalytic cleavage of the O–H bond in water and alcohols, with significantly greater efficiency than the existing, state-of-the-art post-transition metal based systems.
APA, Harvard, Vancouver, ISO, and other styles
22

Silva, Thiago S., and Fernando Coelho. "Methodologies for the synthesis of quaternary carbon centers via hydroalkylation of unactivated olefins: twenty years of advances." Beilstein Journal of Organic Chemistry 17 (July 7, 2021): 1565–90. http://dx.doi.org/10.3762/bjoc.17.112.

Full text
Abstract:
Olefin double-bond functionalization has been established as an excellent strategy for the construction of elaborate molecules. In particular, the hydroalkylation of olefins represents a straightforward strategy for the synthesis of new C(sp3)–C(sp3) bonds, with concomitant formation of challenging quaternary carbon centers. In the last 20 years, numerous hydroalkylation methodologies have emerged that have explored the diverse reactivity patterns of the olefin double bond. This review presents examples of olefins acting as electrophilic partners when coordinated with electrophilic transition-metal complexes or, in more recent approaches, when used as precursors of nucleophilic radical species in metal hydride hydrogen atom transfer reactions. This unique reactivity, combined with the wide availability of olefins as starting materials and the success reported in the construction of all-carbon C(sp3) quaternary centers, makes hydroalkylation reactions an ideal platform for the synthesis of molecules with increased molecular complexity.
APA, Harvard, Vancouver, ISO, and other styles
23

Wang, Xiaoping, Tianbiao Liu, R. Morris Bullock, and Christina Hoffmann. "Heterolytic Cleavage of H2 Revealed by Neutron Single Crystal Diffraction." Acta Crystallographica Section A Foundations and Advances 70, a1 (2014): C922. http://dx.doi.org/10.1107/s2053273314090779.

Full text
Abstract:
Synthetic biologically inspired complexes exhibiting reactivity similar to hydrogenase enzymes have provided evidence of hydride transfer to the metal and proton transfer to an amine, but key structural information about the intermediate is not readily discernible with X-rays. The greater sensitivity of neutron to hydrogen makes it ideal for studying the structure and dynamics of catalytic materials. The newly commissioned TOPAZ neutron single crystal diffractometer at the SNS is capable of continuous 3D diffraction space mapping from a small stationary crystal, permitting detailed structural study at atomic resolution. The structure measured on TOPAZ for an Fe-based mononuclear electrocatalyst confirms that reaction of [CpFeN-L)](BARF) (1) with H2 under mild conditions leads to heterolytic cleavage of the H-H bond into a proton and hydride[1]. The precise location of H atoms in [Fe-H···H-N]+ reveals an unconventional H-bonding interaction, where the ferrous hydridic site {Fe(II)-H-} acts as the H-bond acceptor and the nitrogen of the protic pendant amine {L-N-H+} as the H-bond donor. The neutron structure provides clear evidence of a crucial intermediate involving an Fe-H···H-N interaction in the oxidation of H2. The result clarifies the key role of the pendant amine in the iron complex and provides insights into the design of synthetic electrocatalysts sought as cost-effective alternatives to platinum in fuel cells. The reaction is also a critical step in homogeneous catalysts for hydrogenation of C=O and C=N bonds. A preliminary result from TOPAZ measurement shows that 1 undergoes further single-crystal to single-crystal chemical reaction with moisture in the air, leading to a Fe(H2O)+ complex. Abbreviations: Cp = pentafluoropyridylcyclopentadienide; N-L= 1, 5-di(tert-butyl)-3,7-di(benzyl)-1,5-diaza-3,7-diphospha-cyclooctane; BARF = [B[3,5-(CF3)2C6H3]4]–
APA, Harvard, Vancouver, ISO, and other styles
24

Wang, Shu-Jian, Ying Li, Di Wu, and Zhi-Ru Li. "Metal hydrides as sodium bond acceptors: hydride-sodium bond in the XH ··· NaH (X = HBe, LiBe, NaBe, HMg, LiMg, and NaMg) complexes." Molecular Physics 110, no. 24 (2012): 3053–60. http://dx.doi.org/10.1080/00268976.2012.695807.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Xu, Dongdong, Chunhui Shan, Yingzi Li та ін. "Bond dissociation energy controlled σ-bond metathesis in alkaline-earth-metal hydride catalyzed dehydrocoupling of amines and boranes: a theoretical study". Inorganic Chemistry Frontiers 4, № 11 (2017): 1813–20. http://dx.doi.org/10.1039/c7qi00459a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Wang, Jianqiang, Qi Wang, Xinghua Jiang, Zhongneng Liu, Weimin Yang, and Anatoly I. Frenkel. "Determination of Nanoparticle Size by Measuring the Metal–Metal Bond Length: The Case of Palladium Hydride." Journal of Physical Chemistry C 119, no. 1 (2014): 854–61. http://dx.doi.org/10.1021/jp510730a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

PIKE, R. D. "ChemInform Abstract: Metal-Alkyl and Metal-Hydride Bond Formation and Fission; Oxidative Addition and Reductive Elimination." ChemInform 23, no. 22 (2010): no. http://dx.doi.org/10.1002/chin.199222281.

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

PIKE, R. D. "ChemInform Abstract: Metal-Alkyl and Metal-Hydride Bond Formation and Fission; Oxidative Addition and Reductive Elimination." ChemInform 25, no. 50 (2010): no. http://dx.doi.org/10.1002/chin.199450273.

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

Joska, Jiří, Jan Fajkoš та Jaroslav Zajíček. "Synthesis of 3β,4β-cyclopropano-19-nor-A-homosteroids". Collection of Czechoslovak Chemical Communications 50, № 7 (1985): 1611–17. http://dx.doi.org/10.1135/cccc19851611.

Full text
Abstract:
Acetolysis of the tosylate XIII, afforded the olefin XIV as the sole product. Epoxidation of the double bond gave the epoxide XV which on metal hydride reduction yielded the alcohol XVI. The structures of these products were established by spectral and chemical means and conformation of the A-homo ring in the epoxide XV is discussed on the basis of the 1H NMR spectra.
APA, Harvard, Vancouver, ISO, and other styles
30

Ellis, William W., Alex Miedaner, Calvin J. Curtis, Dorothy H. Gibson, and Daniel L. DuBois. "Hydride Donor Abilities and Bond Dissociation Free Energies of Transition Metal Formyl Complexes." Journal of the American Chemical Society 124, no. 9 (2002): 1926–32. http://dx.doi.org/10.1021/ja0116831.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Schlangen, Maria, and Helmut Schwarz. "Insertion of Molecular Oxygen in Transition-Metal Hydride Bonds, Oxygen-Bond Activation, and Unimolecular Dissociation of Metal Hydroperoxide Intermediates. Short Communication." Helvetica Chimica Acta 91, no. 3 (2008): 379–86. http://dx.doi.org/10.1002/hlca.200890043.

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

Wiles, Jason A., Steven H. Bergens, and Victor G. Young, Jr. "Stereochemistry at carbon upon protonolysis of a late transition metal-alkyl bond: a reaction of relevance to catalytic enantioselective hydrogenation of olefins." Canadian Journal of Chemistry 79, no. 5-6 (2001): 1019–25. http://dx.doi.org/10.1139/v01-051.

Full text
Abstract:
Reaction of [Ru((R)-BINAP)(H)(MeCN)n(acetone)3–n](BF4) (where n = 0–3) (2) with 1 equiv of the olefin substrate methyl α-acetamidoacrylate (MAA) in acetone at room temperature immediately generated a mixture (72:28) of two diastereomers of the complex [Ru((R)-BINAP)(MeCN)(MAA(H))](BF4) (3). The olefin–hydride insertion reaction between 2 and MAA to generate 3 was regioselective, with transfer of the hydride to the β-olefinic carbon and transfer of ruthenium to the α-carbon in both diastereomers of 3. The two diastereomers of 3 differ by the absolute configuration at the α-carbon of MAA(H) ((SCα)-3 and (RCα)-3). The absolute configuration of the major ((SCα)-3) diastereomer was determined by X-ray diffraction in conjunction with NMR spectroscopic data. Protonolysis of the ruthenium–carbon bond in 3 and in the methyl α-acetamidocinnamate (MAC) analog ([Ru((R)-BINAP)(MeCN)((S)- MAC(H))](BF4) ((SCα)-4)) by addition of 2 equiv HBF4·Et2O in CH2Cl2 at room temperature was not stereospecific and did not occur with β-hydride elimination from the methyl or benzyl groups.Key words: ruthenium, BINAP, enantioselective, hydrogenation, catalysis.
APA, Harvard, Vancouver, ISO, and other styles
33

Uhl, Werner, Philipp Wegener, Marcus Layh, Alexander Hepp та Ernst-Ulrich Würthwein. "Reaction of an Al/P-based frustrated Lewis pair with benzophenone: formation of adducts and aluminium alcoholates via β-hydride elimination". Zeitschrift für Naturforschung B 71, № 10 (2016): 1043–50. http://dx.doi.org/10.1515/znb-2016-0118.

Full text
Abstract:
AbstractTreatment of the Al/P-based frustrated Lewis pair (FLP) Mes2P–C(AltBu2)=C(H)-Ph (1) with benzophenone afforded the simple 1:1 adduct (4) with a O=CPh2 molecule coordinated to the aluminum atom by an Al←O donor-acceptor bond. Steric repulsion may prevent an interaction between the electrophilic carbonyl carbon atom and the Lewis-basic phosphorus atom. 4 is unstable in solution at room temperature, the coordination to aluminium increases the polarisation of the carbonyl group and favours its reduction. As suggested by quantum chemical calculations, a C–H bond of a tBu group approaches the electrophilic center and facilitates β-hydride elimination with the release of isobutene and the formation of an Al–OCHPh2 ligation. An intact O=CPh2 molecule completes the coordination sphere of the metal atom (5). The second tBu group at aluminium reacts similarly by the selective formation of an Al(OCHPh2)2 moiety (6). The thermodynamics of adduct formation and the mechanism of the hydride shift have been evaluated by quantum chemical DFT calculations.
APA, Harvard, Vancouver, ISO, and other styles
34

Raubenheimer, Helgard G., Gert J. Kruger, Fred Scott, and Ronald Otte. "Formal nitrogen hydride insertion into the metal-carbene bond of Fischer-type carbene complexes." Organometallics 6, no. 8 (1987): 1789–95. http://dx.doi.org/10.1021/om00151a028.

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

IYAKUTTI, K., Y. KAWAZOE, M. RAJARAJESWARI, and V. J. SURYA. "STUDY OF HYDROGEN STORAGE IN ALUMINUM HYDRIDE COATED SINGLE-WALLED CARBON NANOTUBE." International Journal of Nanoscience 08, no. 01n02 (2009): 43–47. http://dx.doi.org/10.1142/s0219581x09005633.

Full text
Abstract:
In this study, we report the hydrogen storage in aluminum hydride coated single-walled carbon nanotube. All the H 2 adsorption is molecular with H – H bond length of 0.756 Å. The hydrogen storage capacities with half and full coverages are 6.01 (8.3) wt% and 7.2 (10.3) wt%, respectively, without (with) H 3 of AlH 3. At high coverage of AlH 3 ( C 10 AlH 3) interesting clustering/ dimerization of AlH 3 is observed. These systems are quite stable and the H 2 can be extracted from the system without disturbing the C – Al bonding or detaching the AlH 3 from the carbon nanotube. This present study on a full molecular adsorption of hydrogen via light metal-hydride AlH 3 is new and it leads to a practically viable hydrogen storage process.
APA, Harvard, Vancouver, ISO, and other styles
36

Ding, Shengda, Pokhraj Ghosh, Marcetta Y. Darensbourg, and Michael B. Hall. "Interplay of hemilability and redox activity in models of hydrogenase active sites." Proceedings of the National Academy of Sciences 114, no. 46 (2017): E9775—E9782. http://dx.doi.org/10.1073/pnas.1710475114.

Full text
Abstract:
The hydrogen evolution reaction, as catalyzed by two electrocatalysts [M(N2S2)·Fe(NO)2]+, [Fe-Fe]+ (M = Fe(NO)) and [Ni-Fe]+ (M = Ni) was investigated by computational chemistry. As nominal models of hydrogenase active sites, these bimetallics feature two kinds of actor ligands: Hemilabile, MN2S2 ligands and redox-active, nitrosyl ligands, whose interplay guides the H2 production mechanism. The requisite base and metal open site are masked in the resting state but revealed within the catalytic cycle by cleavage of the MS–Fe(NO)2 bond from the hemilabile metallodithiolate ligand. Introducing two electrons and two protons to [Ni-Fe]+ produces H2 from coupling a hydride temporarily stored on Fe(NO)2 (Lewis acid) and a proton accommodated on the exposed sulfur of the MN2S2 thiolate (Lewis base). This Lewis acid–base pair is initiated and preserved by disrupting the dative donation through protonation on the thiolate or reduction on the thiolate-bound metal. Either manipulation modulates the electron density of the pair to prevent it from reestablishing the dative bond. The electron-buffering nitrosyl’s role is subtler as a bifunctional electron reservoir. With more nitrosyls as in [Fe-Fe]+, accumulated electronic space in the nitrosyls’ π*-orbitals makes reductions easier, but redirects the protonation and reduction to sites that postpone the actuation of the hemilability. Additionally, two electrons donated from two nitrosyl-buffered irons, along with two external electrons, reduce two protons into two hydrides, from which reductive elimination generates H2.
APA, Harvard, Vancouver, ISO, and other styles
37

Ni, Shao-Fei, and Li Dang. "Insight into the electronic effect of phosphine ligand on Rh catalyzed CO2 hydrogenation by investigating the reaction mechanism." Physical Chemistry Chemical Physics 18, no. 6 (2016): 4860–70. http://dx.doi.org/10.1039/c5cp07256e.

Full text
Abstract:
The effect of the outer coordination sphere of the diphosphine ligand on the catalytic efficiency of [Rh(PCH<sub>2</sub>X<sup>R</sup>CH<sub>2</sub>P)<sub>2</sub>]<sup>+</sup> (X<sup>R</sup> = CH<sub>2</sub>, N–CH<sub>3</sub>, CF<sub>2</sub>) catalyzed CO<sub>2</sub> hydrogenation was studied. It was found that the hydricity of the metal hydride bond determined the activation energy of the rate determining step of the reaction.
APA, Harvard, Vancouver, ISO, and other styles
38

Venâncio, Ana I. F., Maxim L. Kuznetsov, M. Fátima C. Guedes da Silva, Luísa M. D. R. S. Martins, João J. R. Fraústo da Silva, and Armando J. L. Pombeiro. "Metal−Hydride Bond Activation and Metal−Metal Interaction in Dinuclear Iron Complexes with Linking Dinitriles: A Synthetic, Electrochemical, and Theoretical Study." Inorganic Chemistry 41, no. 24 (2002): 6456–67. http://dx.doi.org/10.1021/ic025835k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Clegg, W., M. Capdevila, P. González-Duarte, and J. Sola. "Structural investigation of homonuclear Pt2 and heteronuclear PdPt complexes containing a metal–metal bond bridged by hydrido and sulfido ligands." Acta Crystallographica Section B Structural Science 52, no. 2 (1996): 270–76. http://dx.doi.org/10.1107/s0108768195010640.

Full text
Abstract:
The complex [Pt2(μ-H)(μ-S)(dppe)2](PF6) undergoes a displacive order–disorder transformation at ca 230 K. The low-temperature structure is ordered with one cation–anion pair as the asymmetric unit in space group P21/n. At room temperature the b axis is halved and the space group is P2/n, imposing crystallographic twofold rotation symmetry on both ions; the anion shows major disorder and there is probably minor disorder in the cation, but its internal geometry remains essentially unchanged. The heteronuclear complex [PdPt(μ-H)(μ-S)(dppe)2](PF6) is isostructural with the Pt2 complex at room temperature. All three structures have been determined crystallographically and both complexes have been extensively characterized by NMR spectroscopy, unambiguously confirming the genuine heteronuclear nature of the mixed-metal complex and the presence of the bridging hydride ligand.
APA, Harvard, Vancouver, ISO, and other styles
40

Ciancanelli, Rebecca, Bruce C. Noll, Daniel L. DuBois, and M. Rakowski DuBois. "Comprehensive Thermodynamic Characterization of the Metal−Hydrogen Bond in a Series of Cobalt-Hydride Complexes." Journal of the American Chemical Society 124, no. 12 (2002): 2984–92. http://dx.doi.org/10.1021/ja0122804.

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

Srivastava, R. S. "Carbon-oxygen bond cleavage in allylic esters promoted by low-valent transition-metal hydride complexes." Applied Organometallic Chemistry 7, no. 8 (1993): 607–11. http://dx.doi.org/10.1002/aoc.590070803.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Al-Ibadi et al., Muhsen. "A Theoretical Investigation on Chemical Bonding of the Bridged Hydride Triruthenium Cluster: [Ru3 (μ-H)( μ3-κ2-Hamphox-N,N)(CO)9]". Baghdad Science Journal 17, № 2 (2020): 0488. http://dx.doi.org/10.21123/bsj.2020.17.2.0488.

Full text
Abstract:
Ruthenium-Ruthenium and Ruthenium–ligand interactions in the triruthenium "[Ru3(μ-H)(μ3-κ2-Hamphox-N,N)(CO)9]" cluster are studied at DFT level of theory. The topological indices are evaluated in term of QTAIM (quantum theory of atoms in molecule). The computed topological parameters are in agreement with related transition metal complexes documented in the research papers. The QTAIM analysis of the bridged core part, i.e., Ru3H, analysis shows that there is no bond path and bond critical point (chemical bonding) between Ru(2) and Ru(3). Nevertheless, a non-negligible delocalization index for this non-bonding interaction is calculated. The interaction in the core Ru3H can be described as a (4centre–4electron) type. For Ru-N (oxazoline ring) bond, the calculated topological data propose a pure σ-bond. The computed topological parameters of oxazoline ligand reveal the presence of slightly some double bond characters within ligand ring.
APA, Harvard, Vancouver, ISO, and other styles
43

Song, Zhenjun, Xiji Shao, Qiang Wang, Chunxin Ma, Kedong Wang, and Deman Han. "Generation of molybdenum hydride species via addition of molecular hydrogen across metal-oxygen bond at monolayer oxide/metal composite interface." International Journal of Hydrogen Energy 45, no. 4 (2020): 2975–88. http://dx.doi.org/10.1016/j.ijhydene.2019.11.135.

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

Macgregor, Stuart A., та Bruce Sweeney. "An unexpected electronic preference for transfer of a β-hydrogen trans to a metal–hydride bond". New Journal of Chemistry 24, № 11 (2000): 855–58. http://dx.doi.org/10.1039/b006546n.

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

Storch, Christine, Reinald Fischer, Dirk Walther та Rhett Kempe. "Der erste stabile Metallkomplex mit Hydrid- und Carbonyl-Brücke ohne weitere Brücken- und π - Akzeptorliganden [(TMEDA)Ni(μ-H,μ-CO)Ni(TMEDA]X (TMEDA: N,N,N′,N′ -Tetramethylethylendiamin, X = Monoanion der cis-Cyclohex-4-en oder Cyclohexandicarbonsäure) / The First Stable Metal Complex with Hydrid- and Carbonyl Bridges without Supporting Bridging and π-Acceptor Ligands [(TMEDA)Ni(μ-H,μ-CO)Ni(TMEDA)]X (TMEDA: N,N,N′,N′-Tetramethylethylenediamine, X = Mono-anions of cis-Cyclohex-4-ene or Cyclohexanedicarboxylic Acid)". Zeitschrift für Naturforschung B 50, № 5 (1995): 816–20. http://dx.doi.org/10.1515/znb-1995-0521.

Full text
Abstract:
Abstract The complex [(TMEDA)Ni(μ-H,μ-CO)Ni (TMEDA)]X 1a (TMEDA: N,N,N',N'-Tetramethylethylenediamine, X: Monoanions of cis-cyclohex-4-ene- or cyclohexane-dicarboxylic acid) could be isolated as the product of a stepwise reaction in THF solution of Ni(COD)2, TMEDA and cis-cyclohex-4-ene-dicarboxylic-anhydride in the presence of a small amount of water. Similar results are obtained with cyclohexane-dicarboxylic-anhydride. NMR- and IR-spectroscopic measurements and the single crystal X-Ray diffraction analysis of 1a have established the structure: Both a hydride- and a carbonyl bridge are linking two (TMEDA)Ni fragments in the cationic part. The Ni(I) centers are connected by a Ni-Ni-bond . The compounds are the first (TMEDA)Ni(I) complexes and the first binuclear hydrido-carbonylcomplexes which are stable without further bridging and π-acceptor ligands. They contain hydrido and protic hydrogen in the same compound.
APA, Harvard, Vancouver, ISO, and other styles
46

Kiss, Gabor, Steven P. Nolan, and Carl D. Hoff. "Direct solution calorimetric measurements of enthalpies of proton and electron transfer reactions for transition metal complexes. Thermochemical study of metal-hydride and metal-metal bond energies." Inorganica Chimica Acta 227, no. 2 (1994): 285–92. http://dx.doi.org/10.1016/0020-1693(94)04218-7.

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

Schneider, Jörg J., Richard Goddard та Carl Krüger. "Oxidative Abbaureaktionen des homonuklearen Cobalthydridclusters [(η5-Cp*)Co]3H4 / Oxidative Degradation Reactions of the Homonuclear Cobalt Hydride Cluster [(η5-Cp*)Co]3H4". Zeitschrift für Naturforschung B 50, № 4 (1995): 448–59. http://dx.doi.org/10.1515/znb-1995-0401.

Full text
Abstract:
The homonuclear cobalt cluster [(η5-Cp*)Co]3H4 2 reacts with various organic and inorganic substrates with complete degradation of the cluster core to afford mainly mononuclear organometallic cobalt compounds. Thus, whereas reaction of the cluster with CO, NO or alkynes results in the retention of the Co3 ring, the cluster reacts with Br2, I2, I2/CO , I2/ P(C 2H5)3, CCl4, HCI3, (CH2)2Br2, Hacac, CH2Br2, (C6H5CO)OOC(CH3)3, HBF4, and BrCN to give mononuclear complexes or bridged dinuclear complexes without a metal-metal bond. In all cases formal oxidation of the metal center is observed. The crystal structures of six organocobalt complexes have been determined by X-ray crystallography.
APA, Harvard, Vancouver, ISO, and other styles
48

Li, Jun, and Kazunari Yoshizawa. "Catalytic Hydrogenation of Carbon Dioxide with a Highly Active Hydride on Ir(III)–Pincer Complex: Mechanism for CO2Insertion and Nature of Metal–Hydride Bond." Bulletin of the Chemical Society of Japan 84, no. 10 (2011): 1039–48. http://dx.doi.org/10.1246/bcsj.20110128.

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

Robertson, GB, and PA Tucker. "The Crystal and Molecular Structure of trans-Dichloro-mer-tris(dimethylphenyl-phosphine)hydridoiridium(III) by X-Ray and Neutron Diffraction." Australian Journal of Chemistry 40, no. 6 (1987): 1043. http://dx.doi.org/10.1071/ch9871043.

Full text
Abstract:
The crystal and molecular structure of mer -(PMe2Ph)3H- trans-Cl2IrIII (1) has been determined by single-crystal X-ray and neutron diffraction analyses. Crystals are monoclinic, space group P1, with a 8.91 9(4), b 9.895(4), c 12.269(6) �, α 70.33(4), β 81.75(4), γ 88.26(4)�, Z 2. The structure was solved by conventional heavy atom methods and refined by full-matrix least-squares analyses to final R values of 0.067 (X-ray data, R = ∑// Fo/-/Fc2/∑Fo/, 7516 independent reflections) and 0.086 (neutron data, R = ∑/Fo2- Fc2/∑ Fo2 2290 independent reflections). Weighted mean values for metal-ligand distances are Ir -P(trans to H) 2.365(2) �, Ir -P(trans to PMe2Ph) 2.329(2) � and Ir -Cl(trans to Cl) 2.402(2) and 2.364(2) �. The iridium-hydrogen bond [1.616(7) �] is trans-lengthened, vis-a-vis Ir -H trans to Cl, by c. 0.05 �. The difference in the mutually trans Ir -Cl bond lengths is ascribed to the inequivalence of intramolecular non-bonding interactions. The Ir -P(trans to PMe2Ph) bond length in (1) is shorter than that in mer -(PMe2Ph)Cl3IrIII. The smaller steric requirement of the hydride ligand in (1) allows the mutually trans phosphines to move away from their cis neighbour. In turn, the phosphines approach the metal more closely to compensate for the decreased phosphine substituent-phosphine substituent non-bonding interactions.
APA, Harvard, Vancouver, ISO, and other styles
50

Al-hamidi, Jehan, Abdulhamid Alsaygh, and Ibrahim Al-Najjar. "Hydridothiazole Rhodium Complexes as a Result of C-H Bond Activation in Iminothiazoles Chelating Ligands." Open Chemistry Journal 1, no. 1 (2014): 27–32. http://dx.doi.org/10.2174/1874842201401010027.

Full text
Abstract:
A series of 20 Schiff base ligands derived from 2-aminothiazole and its derivatives and aryl aldehydes with either [RhCl(PPh3)3] or [Rh(µ-Cl)(COD)]2 in the presence of 4 equivalents of PPh3 lead to an Rh(III) cyclometallated complex and the imine ligand (C-H) bond has been added to the metal (C-M-H). The complexes were investigated by using I.R., 1H, 13C and 31P NMR Spectroscopic techniques. The signal of the (C-H) ligand was observed as trans to the nitrogen atom in the complex which is a donor ligand. Graphical Abstract: Total synthesis of hydridothiazole rhodium complexes possessing rhodium hydride signal at δ (-14.60 to-15.04) ppm, trans to N-donor ligand and iminoyl carbon (7C=N) signal in Rh (III) observed at δ (220.1-237.6)ppm, lower field and suggestive of carbine like properties.
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