To see the other types of publications on this topic, follow the link: Rh(I) chemistry].

Journal articles on the topic 'Rh(I) chemistry]'

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 'Rh(I) chemistry].'

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

Dickson, Ron S. "Alkyne rotation on a RhRh bond: An entry to diverse chemistry." Polyhedron 10, no. 17 (January 1991): 1995–2023. http://dx.doi.org/10.1016/s0277-5387(00)86026-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Yusenko, Kirill V., Aleksandr S. Sukhikh, Werner Kraus, and Sergey A. Gromilov. "Synthesis and Crystal Chemistry of Octahedral Rhodium(III) Chloroamines." Molecules 25, no. 4 (February 11, 2020): 768. http://dx.doi.org/10.3390/molecules25040768.

Full text
Abstract:
Rhodium(III) octahedral complexes with amine and chloride ligands are the most common starting compounds for preparing catalytically active rhodium(I) and rhodium(III) species. Despite intensive study during the last 100 years, synthesis and crystal structures of rhodium(III) complexes were described only briefly. Some [RhClx(NH3)6-x] compounds are still unknown. In this study, available information about synthetic protocols and the crystal structures of possible [RhClx(NH3)6−x] octahedral species are summarized and critically analyzed. Unknown crystal structures of (NH4)2[Rh(NH3)Cl5], trans–[Rh(NH3)4Cl2]Cl⋅H2O, and cis–[Rh(NH3)4Cl2]Cl are reported based on high quality single crystal X-ray diffraction data. The crystal structure of [Rh(NH3)5Cl]Cl2 was redetermined. All available crystal structures with octahedral complexes [RhClx(NH3)6-x] were analyzed in terms of their packings and pseudo-translational sublattices. Pseudo-translation lattices suggest face-centered cubic and hexagonal closed-packed sub-cells, where Rh atoms occupy nearly ideal lattices.
APA, Harvard, Vancouver, ISO, and other styles
3

Gellrich, Urs, and Thorsten Koslowski. "Rh chemistry through the eyes of theory." Wiley Interdisciplinary Reviews: Computational Molecular Science 6, no. 3 (February 25, 2016): 311–20. http://dx.doi.org/10.1002/wcms.1250.

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

Patchett, Ruth, and Adrian B. Chaplin. "Coordination chemistry of a calix[4]arene-based NHC ligand: dinuclear complexes and comparison to IiPr2Me2." Dalton Transactions 45, no. 21 (2016): 8945–55. http://dx.doi.org/10.1039/c6dt01001f.

Full text
Abstract:
The preparation and coordination chemistry of 5,17-bis(3-methyl-1-imidazol-2-ylidene)-25,26,27,28-tetrapropoxycalix[4]arene with rhodium(i) dimers [Rh(COD)Cl]2 and [Rh(CO)2Cl]2 has been explored.
APA, Harvard, Vancouver, ISO, and other styles
5

Gyton, Matthew R., Thomas M. Hood, and Adrian B. Chaplin. "A convenient method for the generation of {Rh(PNP)}+ and {Rh(PONOP)}+ fragments: reversible formation of vinylidene derivatives." Dalton Transactions 48, no. 9 (2019): 2877–80. http://dx.doi.org/10.1039/c8dt05049j.

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

DICKSON, R. S. "ChemInform Abstract: Alkyne Rotation on a Rh-Rh Bond: An Entry to Diverse Chemistry." ChemInform 23, no. 6 (August 22, 2010): no. http://dx.doi.org/10.1002/chin.199206247.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Ojima, Iwao, Robert J. Donovan, Patrizia Ingallina, N�ria Clos, William R. Shay, Masakatsu Eguchi, Qingping Zeng, and Anna Korda. "Organometallic chemistry and homogeneous catalysis of Rh and Rh-Co mixed metal carbonyl clusters." Journal of Cluster Science 3, no. 4 (December 1992): 423–38. http://dx.doi.org/10.1007/bf00702749.

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

Atencio, Reinaldo, Gustavo Chacón, Lisbeth Mendoza, Teresa González, Julia Bruno-Colmenarez, Merlin Rosales, Briceño Alexander, and Edgar Ocando-Mavárez. "Chemistry of transition-metal complexes containing functionalized phosphines: synthesis and structural analysis of rhodium(I) complexes containing allyl and cyanoalkylphosphines." Acta Crystallographica Section C Structural Chemistry 76, no. 9 (August 30, 2020): 932–46. http://dx.doi.org/10.1107/s2053229620011420.

Full text
Abstract:
A series of related acetylacetonate–carbonyl–rhodium compounds substituted by functionalized phosphines has been prepared in good to excellent yields by the reaction of [Rh(acac)(CO)2] (acac is acetylacetonate) with the corresponding allyl-, cyanomethyl- or cyanoethyl-substituted phosphines. All compounds were fully characterized by 31P, 1H, 13C NMR and IR spectroscopy. The X-ray structures of (acetylacetonato-κ2 O,O′)(tert-butylphosphanedicarbonitrile-κP)carbonylrhodium(I), [Rh(C5H7O2)(CO)(C8H13N2)] or [Rh(acac)(CO)(tBuP(CH2CN)2}] (2b), (acetylacetonato-κ2 O,O′)carbonyl[3-(diphenylphosphanyl)propanenitrile-κP]rhodium(I), [Rh(C5H7O2)(C15H14N)(CO)] or [Rh(acac)(CO){Ph2P(CH2CH2CN)}] (2h), and (acetylacetonato-κ2 O,O′)carbonyl[3-(di-tert-butylphosphanyl)propanenitrile-κP]rhodium(I), [Rh(C5H7O2)(C11H22N)(CO)] or [Rh(acac)(CO){tBu2P(CH2CH2CN)}] (2i), showed a square-planar geometry around the Rh atom with a significant trans influence over the acetylacetonate moiety, evidenced by long Rh—O bond lengths as expected for poor π-acceptor phosphines. The Rh—P distances displayed an inverse linear dependence with the coupling constants J P-Rh and the IR ν(C[triple-bond]O) bands, which accounts for the Rh—P electronic bonding feature (poor π-acceptors) of these complexes. A combined study from density functional theory (DFT) calculations and an evaluation of the intramolecular H...Rh contacts from X-ray diffraction data allowed a comparison of the conformational preferences of these complexes in the solid state versus the isolated compounds in the gas phase. For 2b, 2h and 2i, an energy-framework study evidenced that the crystal structures are mainly governed by dispersive energy. In fact, strong pairwise molecular dispersive interactions are responsible for the columnar arrangement observed in these complexes. A Hirshfeld surface analysis employing three-dimensional molecular surface contours and two-dimensional fingerprint plots indicated that the structures are stabilized by H...H, C...H, H...O, H...N and H...Rh intermolecular interactions.
APA, Harvard, Vancouver, ISO, and other styles
9

Pandey, Krishna K. "Coordination chemistry of NSO−, NSO2− and S3N− ligands: comparison of electronic structure of RhNSO, RhSH and RhCl complexes." Inorganica Chimica Acta 182, no. 2 (April 1991): 163–71. http://dx.doi.org/10.1016/s0020-1693(00)90151-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Loreto, Domenico, Giarita Ferraro, and Antonello Merlino. "Unusual Structural Features in the Adduct of Dirhodium Tetraacetate with Lysozyme." International Journal of Molecular Sciences 22, no. 3 (February 2, 2021): 1496. http://dx.doi.org/10.3390/ijms22031496.

Full text
Abstract:
The structures of the adducts formed upon reaction of the cytotoxic paddlewheel dirhodium complex [Rh2(μ-O2CCH3)4] with the model protein hen egg white lysozyme (HEWL) under different experimental conditions are reported. Results indicate that [Rh2(μ-O2CCH3)4] extensively reacts with HEWL:it in part breaks down, at variance with what happens in reactions with other proteins. A Rh center coordinates the side chains of Arg14 and His15. Dimeric Rh–Rh units with Rh–Rh distances between 2.3 and 2.5 Å are bound to the side chains of Asp18, Asp101, Asn93, and Lys96, while a dirhodium unit with a Rh–Rh distance of 3.2–3.4 Å binds the C-terminal carboxylate and the side chain of Lys13 at the interface between two symmetry-related molecules. An additional monometallic fragment binds the side chain of Lys33. These data, which are supported by replicated structural determinations, shed light on the reactivity of dirhodium tetracarboxylates with proteins, providing useful information for the design of new Rh-containing biomaterials with an array of potential applications in the field of catalysis or of medicinal chemistry and valuable insight into the mechanism of action of these potential anticancer agents.
APA, Harvard, Vancouver, ISO, and other styles
11

Medici, Serenella, Massimiliano Peana, Alessio Pelucelli, and Maria Antonietta Zoroddu. "Rh(I) Complexes in Catalysis: A Five-Year Trend." Molecules 26, no. 9 (April 27, 2021): 2553. http://dx.doi.org/10.3390/molecules26092553.

Full text
Abstract:
Rhodium is one of the most used metals in catalysis both in laboratory reactions and industrial processes. Despite the extensive exploration on “classical” ligands carried out during the past decades in the field of rhodium-catalyzed reactions, such as phosphines, and other common types of ligands including N-heterocyclic carbenes, ferrocenes, cyclopentadienyl anion and pentamethylcyclopentadienyl derivatives, etc., there is still lively research activity on this topic, with considerable efforts being made toward the synthesis of new preformed rhodium catalysts that can be both efficient and selective. Although the “golden age” of homogeneous catalysis might seem over, there is still plenty of room for improvement, especially from the point of view of a more sustainable chemistry. In this review, temporally restricted to the analysis of literature during the past five years (2015–2020), the latest findings and trends in the synthesis and applications of Rh(I) complexes to catalysis will be presented. From the analysis of the most recent literature, it seems clear that rhodium-catalyzed processes still represent a stimulating challenge for the metalloorganic chemist that is far from being over.
APA, Harvard, Vancouver, ISO, and other styles
12

Bickley, Jamie, Simon J. Higgins, and Annemarie A. La Pensée. "Homoleptic Rh(III)-Diphosphine and Rh(III)-Diarsine Complexes." Collection of Czechoslovak Chemical Communications 68, no. 8 (2003): 1461–66. http://dx.doi.org/10.1135/cccc20031461.

Full text
Abstract:
The first examples of homoleptic diarsine and diphosphine coordination to Rh(III), viz. [Rh(1,2-C6H4{AsMe2}2)3]3+ and [Rh(Me2PCH2PMe2)3]3+, are reported, together with details of their characterisation by multinuclear NMR spectroscopy and mass spectrometry.
APA, Harvard, Vancouver, ISO, and other styles
13

Roy, Dipak Kumar, K. Yuvaraj, R. Jagan, and Sundargopal Ghosh. "Chemistry of Rh-N,S heterocyclic carbene complexes." Journal of Organometallic Chemistry 811 (June 2016): 8–13. http://dx.doi.org/10.1016/j.jorganchem.2016.03.012.

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

Jacob, K. T., Preeti Gupta, M. Vinay, and Y. Waseda. "Phase chemistry of the system Nb–Rh–O." Journal of Solid State Chemistry 202 (June 2013): 234–40. http://dx.doi.org/10.1016/j.jssc.2013.03.056.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Imler, Gregory H., Michael J. Zdilla, and Bradford B. Wayland. "Evaluation of the Rh(II)–Rh(II)Bond Dissociation Enthalpy for [(TMTAA)Rh]2by1H NMR T2Measurements: Application in Determining the Rh–C(O)– BDE in [(TMTAA)Rh]2C═O." Inorganic Chemistry 52, no. 19 (October 7, 2013): 11509–13. http://dx.doi.org/10.1021/ic401778c.

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

Groenewald, Ferdi G., and Andreas Roodt. "Theoretical insights into the Rh···Rh interactions." Acta Crystallographica Section A Foundations and Advances 72, a1 (August 28, 2016): s331. http://dx.doi.org/10.1107/s2053273316095061.

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

Erasmus, Johannes, and Jeanet Conradie. "Oxidative addition of methyl iodide to [Rh(PhCOCHCOPh)(CO)(P(OCH2)3CCH3)]: an experimental and computational study." Open Chemistry 10, no. 1 (February 1, 2012): 256–66. http://dx.doi.org/10.2478/s11532-011-0137-0.

Full text
Abstract:
AbstractThe reaction rate of the oxidative addition and CO insertion steps of methyl iodide with [Rh(PhCOCHCOPh)(CO)(P(OCH2)3CCH3)] are presented. Large negative experimental values for the activation entropy and results from a density functional theory computational chemistry study indicated trans addition of the CH3I to [Rh(PhCOCHCOPh)(CO)(P(OCH2)3CCH3)]. A study of the molecular orbitals gives insight into the flow of electrons during the oxidative addition reaction. CO insertion leads to a square pyramidal [Rh(PhCOCHCOPh)(P(OCH2)3CCH3)(COCH3)(I)] acyl product with the COCH3 moiety in the apical position. The strong electron donation of the P(OCH2)3CCH3 ligand accelerates the oxidation addition step of methyl iodide to [Rh(PhCOCHCOPh)(CO)(P(OCH2)3CCH3)] by ca. 265 times faster (at 35°C) than that of the Monsanto catalyst, but inhibits the CO insertion step.
APA, Harvard, Vancouver, ISO, and other styles
18

Park, Kwang-Min, and Il-Wun Shim. "Preparation of Rh-containing cellulose acetate films and the chemistry of Rh in cellulose acetate." Journal of Applied Polymer Science 42, no. 5 (March 5, 1991): 1361–69. http://dx.doi.org/10.1002/app.1991.070420519.

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

Gloaguen, Yann, Christophe Rebreyend, Martin Lutz, Pravin Kumar, Martina Huber, Jarl Ivar van der Vlugt, Sven Schneider, and Bas de Bruin. "An Isolated Nitridyl Radical-Bridged {Rh(N.)Rh} Complex." Angewandte Chemie International Edition 53, no. 26 (May 19, 2014): 6814–18. http://dx.doi.org/10.1002/anie.201403445.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Caglar, B., M. Olus Ozbek, J. W. (Hans) Niemantsverdriet, and C. J. (Kees-Jan) Weststrate. "The effect of C–OH functionality on the surface chemistry of biomass-derived molecules: ethanol chemistry on Rh(100)." Physical Chemistry Chemical Physics 18, no. 43 (2016): 30117–27. http://dx.doi.org/10.1039/c6cp06069b.

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

Cao, Shiwei, Yang Wang, Zhi Qin, Fangli Fan, Hiromitsu Haba, Yukiko Komori, Xiaolei Wu, Cunmin Tan, and Xin Zhang. "Gas-phase chemistry of ruthenium and rhodium carbonyl complexes." Physical Chemistry Chemical Physics 18, no. 1 (2016): 119–25. http://dx.doi.org/10.1039/c5cp05670e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Yamazaki, Shin-ichi, Yusuke Yamada, and Kazuaki Yasuda. "Reversible Electrochemical Conversion between Rh(II) and Rh(III) States in Rh Porphyrin Adsorbed on Carbon Black." Inorganic Chemistry 43, no. 23 (November 2004): 7263–65. http://dx.doi.org/10.1021/ic0490472.

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

HÜCKSTÄDT, HEINER, CLEMENS BRUHN, and HEINER HOMBORG. "Dinuclear RhII-Phthalocyaninates with a Rh-Rh Single Bond: X-ray Crystal Structure of Di(pyridinephthalocyaninato(2-)rhodium(II))." Journal of Porphyrins and Phthalocyanines 01, no. 04 (October 1997): 367–78. http://dx.doi.org/10.1002/(sici)1099-1409(199710)1:4<367::aid-jpp31>3.0.co;2-q.

Full text
Abstract:
Red, diamagnetic di(phthalocyaninato(2-)rhodium(II)), [( Rhpc 2−)2] is prepared by thermal decomposition of acidophthalocyaninatorhodium(III) or di(acido)phthalocyaninatorhodates(III) in inert, high boiling solvents or under reduced pressure at T < 350° C . It is insoluble in most common solvents except pyridine ( py ), yielding red, diamagnetic di(pyridinephthalocyaninato(2-)rhodium(II)), [{ Rh ( py ) pc 2−}2]. This crystallizes as benzene solvate in the cubic space group Ia-3 (no. 206) (a = 35.960(7) Å) with Z = 24. The Rh - Rh distance (d( Rh - Rh ) = 2.741(2) Å) indicates a strong unsupported Rh - Rh single bond. Due to the axial coordination of py the Rh atoms are virtually bonded in the centre ( ct ) of the pc 2− core (d( Rh - ct ) = 0.08(1) Å). The pc 2− ligands are staggered with the skew angle φ( N iso - Rh - Rh '- N iso ') = 42(1)°. Asymmetrical doming of the pc 2− ligands is caused by a short interplanar distance (d( ct - ct ) = 2.89(1) Å) and axial coordination of pyridine. Due to the labilizing trans influence of the Rh - Rh bond the d( Rh - N py ) = 2.309(8) Å is rather long. Two quasi-reversible anodic and four cathodic electron transfer processes are observed in the differential pulse voltammogram of [{ Rh ( py ) pc 2−}2] at 0.81, 0.55, −0.33, −0.50, −0.75 and −1.33 V. The UV/vis NIR spectrum of [{ Rh ( py ) pc 2−}2] shows the typical π-π* transitions of the pc 2− ligand. The B region is split into two bands at 15 440 ( B −) and 16 890 cm−1( B +) of equal intensity due to strong excitonic coupling. The totally symmetric Rh - Rh stretching vibration is selectively enhanced in the Fourier Transform–Raman spectrum at 176 cm−1.
APA, Harvard, Vancouver, ISO, and other styles
24

Moore, William, Wade Henke, Davide Lionetti, Victor Day, and James Blakemore. "Single-Electron Redox Chemistry on the [Cp*Rh] Platform Enabled by a Nitrated Bipyridyl Ligand." Molecules 23, no. 11 (November 2, 2018): 2857. http://dx.doi.org/10.3390/molecules23112857.

Full text
Abstract:
[Cp*Rh] complexes (Cp* = pentamethylcyclopentadienyl) are attracting renewed interest in coordination chemistry and catalysis, but these useful compounds often undergo net two-electron redox cycling that precludes observation of individual one-electron reduction events. Here, we show that a [Cp*Rh] complex bearing the 4,4′-dinitro-2,2′-bipyridyl ligand (dnbpy) (3) can access a distinctive manifold of five oxidation states in organic electrolytes, contrasting with prior work that found no accessible reductions in aqueous electrolyte. These states are readily generated from a newly isolated and fully characterized rhodium(III) precursor complex 3, formulated as [Cp*Rh(dnbpy)Cl]PF6. Single-crystal X-ray diffraction (XRD) data, previously unavailable for the dnbpy ligand bound to the [Cp*Rh] platform, confirm the presence of both [η5-Cp*] and [κ2-dnbpy]. Four individual one-electron reductions of 3 are observed, contrasting sharply with the single two-electron reductions of other [Cp*Rh] complexes. Chemical preparation and the study of the singly reduced species with electronic absorption and electron paramagnetic resonance spectroscopies indicate that the first reduction is predominantly centered on the dnbpy ligand. Comparative cyclic voltammetry studies with [NBu4][PF6] and [NBu4][Cl] as supporting electrolytes indicate that the chloride ligand can be lost from 3 by ligand exchange upon reduction. Spectroelectrochemical studies with ultraviolet (UV)-visible detection reveal isosbestic behavior, confirming the clean interconversion of the reduced forms of 3 inferred from the voltammetry with [NBu4][PF6] as supporting electrolyte. Electrochemical reduction in the presence of triethylammonium results in an irreversible response, but does not give rise to catalytic H2 evolution, contrasting with the reactivity patterns observed in [Cp*Rh] complexes bearing bipyridyl ligands with less electron-withdrawing substituents.
APA, Harvard, Vancouver, ISO, and other styles
25

Toth, Robert T., and Douglas W. Stephan. "Towards supported catalyst models: the synthesis, characterization, redox chemistry, and structures of the complexes Ti(OAr′)4 (Ar′ = C6H4(2-t-Bu), C6H(2,3,5,6-Me)4)." Canadian Journal of Chemistry 69, no. 1 (January 1, 1991): 172–78. http://dx.doi.org/10.1139/v91-027.

Full text
Abstract:
Reaction of substituted phenoxides with TiCl4 affords the species Ti(OAr′)4 (Ar′ = C6H4(2-t-Bu), 1; Ar′ = C6H(2,3,5,6-Me)4, 2). The compound Ti(OC6H4(2-t-Bu))4, 1, crystallizes in the tetragonal space group [Formula: see text], with a = 15.203(4) Å, c = 8.026(3) Å, Z = 2, and V = 1855(2) Å3. The compound Ti(OC6H(2,3,5,6-Me)4)4, 2, crystallizes in the orthorhombic space group Pbcn, with a = 16.539(7) Å, b = 16.136(6) Å, c = 27.716(12) Å, Z = 8, and V = 7397(9) Å3. The geometry of the Ti coordination sphere in these complexes is best described as pseudo-tetrahedral. In the case of 1 strict crystallographic [Formula: see text] symmetry is imposed. The complex 2 exhibits reversible cyclic voltammetric behaviour consistent with a one electron reduction to the Ti(III) analogue. Chemical reduction of 2 employing sodium amalgam affords the quantitative formation of (C6H(2,3,5,6-Me)4O)2Ti(μ-OC6H(2,3,5,6-Me)4)2Na(THF)2, 3. The reaction of 3 with [(COD)Rh(μ-Cl)]2 does not afford the Ti(III)/Rh(I) early–late heterobimetallic (ELHB) complex (C6H(2,3,5,6-Me)4O)2Ti(μ-OC6H(2,3,5,6-Me)4)2Rh(COD). The nature of all products is not known; however, redox chemistry, in which electron transfer from Ti(III) to Rh(I) occurs is evidenced by the generation of 2 and Rh(0). In addition, ligand transfer reactions giving uncharacterized Rh-alkoxides are suggested by the spectral data. The implications and ramifications for the synthesis of alkoxide bridged ELHB models of bimetallic heterogeneous catalyst systems are discussed. Key words: titanium phenoxides, redox chemistry, structures.
APA, Harvard, Vancouver, ISO, and other styles
26

Dick, David G., and Douglas W. Stephan. "Synthesis and structural studies of rhodium complexes of phosphorus—sulfur ligands." Canadian Journal of Chemistry 64, no. 9 (September 1, 1986): 1870–75. http://dx.doi.org/10.1139/v86-308.

Full text
Abstract:
Rhodium complexes of the phosphorus—sulfur ligands, 2-diphenylphosphinoethyl methyl sulfide (MeSP), 1, and 2-diphenylphosphinothiophene (PTH), 2, have been prepared and studied by single crystal X-ray diffraction methods. [Rh(MeSP)2]BF4•H2O, 3, crystallizes in the space group P21/n with a = 16.939(6) Å, b = 17.152(5) Å, c = 12.049(9) Å, β = 106.50(4)°, and Z = 4. The MeSP ligands chelate to Rh yielding a distorted square-planar geometry. The disposition of the methyl groups on the cis sulfur atoms is transoid. Average Rh—P and Rh—S bond distances were found to be 2.225(3) and 2.347(3) Å, respectively. [Rh(PTH)2(COD)]BF4, 4, crystallizes in the space group Cc with a = 15.862(2) Å, b = 15.112(3) Å, c = 16.029(3) Å, β = 103.32(1)°, and Z = 4. The Rh atom in 4 also has essentially a square-planar coordination geometry. 2 does not chelate but rather is monohapto through phosphorus. Rh—P distances of 2.319(3) and 2.378(3) Å and Rh—C distances of 2.17(1), 2.22(1), 2.24(1), and 2.27(1) Å were found. The small variations in the Rh—P and Rh—C bonds distances appear to be a result of steric interactions between 2 and the COD ligand.
APA, Harvard, Vancouver, ISO, and other styles
27

Crozier, P. A., and P. Claus. "Nano-characterization of Rh-Sn bimetallic catalysts." Proceedings, annual meeting, Electron Microscopy Society of America 54 (August 11, 1996): 224–25. http://dx.doi.org/10.1017/s0424820100163587.

Full text
Abstract:
Bimetallic catalysts are of considerable importance because they possess both high activity and selectivity. Recently, there has been considerable interest in bimetallic catalysts produced by surface organometallic chemistry on metals. A detailed description of the structure and composition at the nanometer level is critical to fully understand the behavior of such catalysts. We have undertaken a study of the microstructure of Rh-Sn/SiO2 bimetallic hydrogenation catalysts.A Rh-Sn precursor was prepared by reacting tetrabutyltin with a silica supported Rh parent catalyst (1 %Rh/SiO2). The bimetallic catalysts were produced by thermal decomposition of the precursor in flowing hydrogen at 623 K. Five different catalysts were produced with a range of different Sn loadings from Rh(1%)/SiO2 to Rh(1%)-Sn(1.85%)/SiO2. The resulting bimetallic catalysts were able to selectively hydrogenate isolated and conjugated C=O functional groups. In situ XPS showed that the Sn and Rh were in the fully reduced state. Mossbauer spectroscopy studies confirmed that Sn was present in the zerovalent state indicating that no oxidized Sn was present. Preliminary IR data suggests that most of the Rh atoms are isolated from their neighbors (presumably by Sn).
APA, Harvard, Vancouver, ISO, and other styles
28

Cooper, David A., Steven J. Rettig, and Alan Storr. "Synthesis, characterization, and X-ray structures of Rh(I) monocarbonyl complexes containing unsymmetric tridentate pyrazolylgallate ligands." Canadian Journal of Chemistry 64, no. 3 (March 1, 1986): 566–74. http://dx.doi.org/10.1139/v86-091.

Full text
Abstract:
The reaction of the [Rh(CO)2Cl]2 dimer with the Na+[Me2Ga(N2C3H3)(OCH2(C5H4N))]− ligand has yielded the planar four-coordinate species [Me2Ga(N2C3H3)(OCH2(C5H4N))]Rh(CO), (LRh(CO)), displaying the tridentate gallate ligand in a meridional coordination mode. In addition, a second product, of similar geometry but with one of the Me groups on the Ga replaced by a Cl atom, viz, [(Cl)MeGa(N2C3H3)(OCH2(C5H4N))]Rh(CO), has also been isolated and characterized. The former complex undergoes a facile oxidative addition reaction with MeI, the transient six-coordinate Rh(III) species produced being rapidly converted, in a methyl migration step, to the five-coordinate Rh(III) acetyl complex, LRh(COMe)I. Crystals of [Me2Ga(N2C3H3)(OCH2(C5H4N))]Rh(CO) are monoclinic, a = 13.139(2), b = 13.324(2), c = 17.352(2) Å, β = 103.251(7)°, Z = 8, space group I2/a, and those of [(Cl)MeGa(N2C3H3)(OCH2(C5H4N))]Rh(CO) are triclinic, a = 8.846(2), b = 12.714(3), c = 7.631(2) Å, α = 93.82(1), β = 113.94(1), γ = 107.99(1)°, Z = 2, space group [Formula: see text] Both structures were solved by conventional heavy-atom methods and were refined by full-matrix least-squares procedures to final R values of 0.029 and 0.048 for 1890 and 1939 reflections with I ≥ 3σ(I), respectively. Both molecules display irregular square planar coordination geometry about Rh with Rh—O = 2.038(3) and 2.048(3), Rh—N(pyrazolyl) = 2.022(4) and 2.025(7), Rh—N(pyridyl) = 2.038(3) and 2.020(6), Rh—CO = 1.778(5) and 1.808(9) Å, respectively, for the two compounds. Molecules of [Me2Ga(N2C3H3)(OCH2(C5H4N))]Rh(CO) form weakly associated, centrosymmetric dimers via an intermolecular [Formula: see text] interaction of 3.5445(7) Å.
APA, Harvard, Vancouver, ISO, and other styles
29

Ngoune, Bernard Bobby, Hamida Hallil, Bérengère Lebental, Guillaume Perrin, Shekhar Shinde, Eric Cloutet, Julien George, Stéphane Bila, Dominique Baillargeat, and Corinne Dejous. "Selective Outdoor Humidity Monitoring Using Epoxybutane Polyethyleneimine in a Flexible Microwave Sensor." Chemosensors 11, no. 1 (December 23, 2022): 16. http://dx.doi.org/10.3390/chemosensors11010016.

Full text
Abstract:
The rise of gas-sensing applications and markets has led to microwave sensors associated to polymer-based sensitive materials gaining a lot of attention, as they offer the possibility to target a large variety of gases (as polymers can be easily functionalised) at ultra-low power and wirelessly (which is a major concern in the Internet of Things). A two-channel microstrip sensor with one resonator coated with 1,2 epoxybutane-functionalised poly(ethyleneimine) (EB-PEI) and the other left bare was designed and fabricated for humidity sensing. The sensor, characterised under controlled laboratory conditions, showed exponential response to RH between 0 and 100%, which is approximated to −1.88 MHz/RH% (−0.03 dB/RH%) and −8.24 MHz/RH% (−0.171 dB/RH%) in the RH ranges of 30–80% and 80–100%, respectively. This is the first reported use of EB-PEI for humidity sensing, and performances, especially at high humidity level (RH > 80%), as compared with transducer working frequencies, are better than the state of the art. When further tested in real outdoor conditions, the sensor shows satisfying performances, with 4.2 %RH mean absolute error. Most importantly, we demonstrate that the sensor is selective to relative humidity alone, irrespective of the other environmental variables acquired during the campaign (O3, NO, NO2, CO, CO2, and Temperature). The sensitivities obtained outdoors in the ranges of 50–70% and 70–100% RH (−0.61 MHz/%RH and −3.68 MHz/%RH, respectively) were close to lab results (−0.95 MHz/%RH and −3.51 MHz/%RH, respectively).
APA, Harvard, Vancouver, ISO, and other styles
30

Tang, Mingjin, James Keeble, Paul J. Telford, Francis D. Pope, Peter Braesicke, Paul T. Griffiths, N. Luke Abraham, et al. "Heterogeneous reaction of ClONO<sub>2</sub> with TiO<sub>2</sub> and SiO<sub>2</sub> aerosol particles: implications for stratospheric particle injection for climate engineering." Atmospheric Chemistry and Physics 16, no. 23 (December 12, 2016): 15397–412. http://dx.doi.org/10.5194/acp-16-15397-2016.

Full text
Abstract:
Abstract. Deliberate injection of aerosol particles into the stratosphere is a potential climate engineering scheme. Particles injected into the stratosphere would scatter solar radiation back to space, thereby reducing the temperature at the Earth's surface and hence the impacts of global warming. Minerals such as TiO2 or SiO2 are among the potentially suitable aerosol materials for stratospheric particle injection due to their greater light-scattering ability than stratospheric sulfuric acid particles. However, the heterogeneous reactivity of mineral particles towards trace gases important for stratospheric chemistry largely remains unknown, precluding reliable assessment of their impacts on stratospheric ozone, which is of key environmental significance. In this work we have investigated for the first time the heterogeneous hydrolysis of ClONO2 on TiO2 and SiO2 aerosol particles at room temperature and at different relative humidities (RHs), using an aerosol flow tube. The uptake coefficient, γ(ClONO2), on TiO2 was ∼ 1.2 × 10−3 at 7 % RH and remained unchanged at 33 % RH, and increased for SiO2 from ∼ 2 × 10−4 at 7 % RH to ∼ 5 × 10−4 at 35 % RH, reaching a value of ∼ 6 × 10−4 at 59 % RH. We have also examined the impacts of a hypothetical TiO2 injection on stratospheric chemistry using the UKCA (United Kingdom Chemistry and Aerosol) chemistry–climate model, in which heterogeneous hydrolysis of N2O5 and ClONO2 on TiO2 particles is considered. A TiO2 injection scenario with a solar-radiation scattering effect very similar to the eruption of Mt Pinatubo was constructed. It is found that, compared to the eruption of Mt Pinatubo, TiO2 injection causes less ClOx activation and less ozone destruction in the lowermost stratosphere, while reduced depletion of N2O5 and NOx in the middle stratosphere results in decreased ozone levels. Overall, no significant difference in the vertically integrated ozone abundances is found between TiO2 injection and the eruption of Mt Pinatubo. Future work required to further assess the impacts of TiO2 injection on stratospheric chemistry is also discussed.
APA, Harvard, Vancouver, ISO, and other styles
31

Brown, Nicole F., and Mark A. Barteau. "Epoxides as probes of oxametallacycle chemistry on Rh(111)." Surface Science 298, no. 1 (December 1993): 6–17. http://dx.doi.org/10.1016/0039-6028(93)90075-u.

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

Mague, Joel. "Gmelin Handbook of Inorganic Chemistry, 8th Edition. Rh. Rhodium." Organometallics 4, no. 7 (July 1985): 1320. http://dx.doi.org/10.1021/om00126a900.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Houtman, C. J., N. F. Brown, and M. A. Barteau. "The Chemistry of Acetates on the Rh(111) Surface." Journal of Catalysis 145, no. 1 (January 1994): 37–53. http://dx.doi.org/10.1006/jcat.1994.1005.

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

Chen, J., C. S. Zhao, N. Ma, and P. Yan. "Aerosol hygroscopicity parameter derived from the light scattering enhancement factor measurements in the North China Plain." Atmospheric Chemistry and Physics 14, no. 15 (August 13, 2014): 8105–18. http://dx.doi.org/10.5194/acp-14-8105-2014.

Full text
Abstract:
Abstract. The relative humidity (RH) dependence of aerosol light scattering is an essential parameter for accurate estimation of the direct radiative forcing induced by aerosol particles. Because of insufficient information on aerosol hygroscopicity in climate models, a more detailed parameterization of hygroscopic growth factors and resulting optical properties with respect to location, time, sources, aerosol chemistry and meteorology are urgently required. In this paper, a retrieval method to calculate the aerosol hygroscopicity parameter, κ, is proposed based on the in situ measured aerosol light scattering enhancement factor, namely f(RH), and particle number size distribution (PNSD) obtained from the HaChi (Haze in China) campaign. Measurements show that f(RH) increases sharply with increasing RH, and that the time variance of f(RH) is much greater at higher RH. A sensitivity analysis reveals that the f(RH) is more sensitive to the aerosol hygroscopicity than PNSD. f(RH) for polluted cases is distinctly higher than that for clean periods at a specific RH. The derived equivalent κ, combined with the PNSD measurements, is applied in the prediction of the cloud condensation nuclei (CCN) number concentration. The predicted CCN number concentration with the derived equivalent κ agrees well with the measured ones, especially at high supersaturations. The proposed calculation algorithm of κ with the f(RH) measurements is demonstrated to be reasonable and can be widely applied.
APA, Harvard, Vancouver, ISO, and other styles
35

Caglar, B., J. W. (Hans) Niemantsverdriet, and C. J. (Kees-Jan) Weststrate. "Correction: Modeling the surface chemistry of biomass model compounds on oxygen-covered Rh(100)." Physical Chemistry Chemical Physics 19, no. 1 (2017): 893. http://dx.doi.org/10.1039/c6cp90289h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Mészáros, János P., Veronika F. S. Pape, Gergely Szakács, Gábor Németi, Márk Dénes, Tamás Holczbauer, Nóra V. May, and Éva A. Enyedy. "Half-sandwich organometallic Ru and Rh complexes of (N,N) donor compounds: effect of ligand methylation on solution speciation and anticancer activity." Dalton Transactions 50, no. 23 (2021): 8218–31. http://dx.doi.org/10.1039/d1dt00808k.

Full text
Abstract:
Synthesis of organometallic half-sandwich polypyridyl Ru and Rh complexes. Anticancer activity against resistant cancer cell lines and effects of ligand methylation on aqueous chemistry and structure.
APA, Harvard, Vancouver, ISO, and other styles
37

Cheng, Guanghao, Gurong Shen, Jun Wang, Yunhao Wang, Weibo Zhang, Jianqiang Wang, and Meiqing Shen. "The Hydrothermal Stability and the Properties of Non- and Strongly-Interacting Rh Species over Rh/γ, θ-Al2O3 Catalysts." Catalysts 11, no. 1 (January 13, 2021): 99. http://dx.doi.org/10.3390/catal11010099.

Full text
Abstract:
The present work reports the effects of γ-, θ-phase of alumina on the hydrothermal stability and the properties of non- and strongly-interacting Rh species of the Rh/Al2O3 catalysts. Comparing to γ-Al2O3, θ-Al2O3 can not only reduce the amount of occluded Rh but also better stabilize Rh during hydrothermal aging treatment. When the aging time was prolonged to 70 h, all the non-interacting Rh was transformed into strongly-interacting Rh and occluded Rh. The XPS results indicated that non- and strongly-interacting Rh might exist in the form of Rh/Rh3+ and Rh4+, respectively. CO-NO reaction was chosen as a probe reaction to research more information about non- and strongly-interacting Rh. The two Rh species had similar apparent activation energy (Eapp) of 170 kJ/mol, which indicated that non- and strongly-interacting Rh follow the same reaction path. The non-interacting Rh was removed from aged samples by the acid-treated method, and obtained results showed that only 2.5% and 4.0% non-interacting Rh was maintained in aged Rh/γ-Al2O3 and Rh/θ-Al2O3.
APA, Harvard, Vancouver, ISO, and other styles
38

Schumann, Herbert, and Susanne Stenz. "Carbonyl(diphenylcyclopentadienyl)-und Carbonyl(tetraphenylcyclopentadienyl)rhodium(I) Komplexe/Carbonyl(diphenylcyclopentadienyl)-and Carbonyl(tetraphenylcyclopentadienyl)rhodium(I) Complexes." Zeitschrift für Naturforschung B 56, no. 12 (December 1, 2001): 1293–96. http://dx.doi.org/10.1515/znb-2001-1207.

Full text
Abstract:
Abstract [Rh(CO)2Cl]2 reacts with 1,3-diphenylcyclopentadienyl sodium or tetraphenylcyclopentadienyl sodium with formation of (1,3 -Ph2C5H3)Rh(CO)2 (1) or (C5HPh4)Rh(CO)2 (2), respectively. The reaction of 1 or 2 with PPh3 yields the corresponding carbonyl-(cyclopentadienyl)(triphenylphosphane)rhodium complexes (1,3 -Ph2C5H3)Rh(CO)(PPh3) (3) or (C5HPh4)Rh(CO)(PPh3) (4). The 1H and 1C reported and discussed.
APA, Harvard, Vancouver, ISO, and other styles
39

Kuang, Ye, Chunsheng Zhao, Jiangchuan Tao, Yuxuan Bian, Nan Ma, and Gang Zhao. "A novel method for deriving the aerosol hygroscopicity parameter based only on measurements from a humidified nephelometer system." Atmospheric Chemistry and Physics 17, no. 11 (June 7, 2017): 6651–62. http://dx.doi.org/10.5194/acp-17-6651-2017.

Full text
Abstract:
Abstract. Aerosol hygroscopicity is crucial for understanding roles of aerosol particles in atmospheric chemistry and aerosol climate effects. Light-scattering enhancement factor f(RH, λ) is one of the parameters describing aerosol hygroscopicity, which is defined as f(RH, λ) = σsp(RH, λ)∕σsp(dry, λ), where σsp(RH, λ) or σsp(dry, λ) represents σsp at wavelength λ under certain relative humidity (RH) or dry conditions. Traditionally, an overall hygroscopicity parameter κ can be retrieved from measured f(RH, λ), hereinafter referred to as κf(RH), by combining concurrently measured particle number size distribution (PNSD) and mass concentration of black carbon. In this paper, a new method is proposed to directly derive κf(RH) based only on measurements from a three-wavelength humidified nephelometer system. The advantage of this newly proposed approach is that κf(RH) can be estimated without any additional information about PNSD and black carbon. This method is verified with measurements from two different field campaigns. Values of κf(RH) estimated from this new method agree very well with those retrieved by using the traditional method: all points lie near the 1 : 1 line and the square of correlation coefficient between them is 0.99. The verification results demonstrate that this newly proposed method of deriving κf(RH) is applicable at different sites and in seasons of the North China Plain and might also be applicable in other regions around the world.
APA, Harvard, Vancouver, ISO, and other styles
40

Thackray, David C., Sara Ariel, Tak W. Leung, Kusum Menon, Brian R. James, and James Trotter. "The synthesis, X-ray structure, and substitution lability of chloro(2,3,7,8,12,13,17,18-octaethylporphinato)(triphenylphosphine)rhodium(III)." Canadian Journal of Chemistry 64, no. 12 (December 1, 1986): 2440–46. http://dx.doi.org/10.1139/v86-404.

Full text
Abstract:
The rhodium(III) octaethylporphyrin complex Rh(OEP)(PPh3)Cl (1) has been synthesized via Rh(III) or Rh(I) precursors, and fully characterized both by spectroscopy and single crystal data. The crystals, available as a bis(chloroform) solvate are triclinic, P1, a = 13.478(5), b = 14.300(5), c = 15.346(4) Å, α = 102.33(2), β = 102.89(2), γ = 90.56(3)°, Z = 2, Dx = 1.384 g cm−3. The structure was determined from Mo diffractometer data and refined by least-squares methods to R = 0.095, Rw = 0.068 for 5189 reflections. The octahedrally coordinated rhodium atom is displaced by 0.077 Å from the mean plane of the four N atoms, towards the triphenylphosphine group. The average Rh – ring nitrogen distance is 2.024 Å, Rh—P is 2.306(3) Å and Rh—Cl, 2.442(2) Å. Solution equilibria studies on 1 describe formation of Rh(OEP)L2+ (L = PPh3, PnBu3) via thermal reactions (including thermodynamic data for the PPh3 system), and formation of Rh(OEP)Cl(L′) species (L′ = CO, THF, MeCN) via photochemical processes.
APA, Harvard, Vancouver, ISO, and other styles
41

Zhang, Jia-Sheng, and Guo-Xin Jin. "Synthesis and characterization of hetero-binuclear Co–Rh complexes [Cp∗CoS2C2(B9H10)][Rh(COD)] and [Cp∗CoSe2C2(B10H10)][Rh(COD)]." Journal of Organometallic Chemistry 692, no. 18 (August 2007): 3944–48. http://dx.doi.org/10.1016/j.jorganchem.2007.06.002.

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

Hu, Zhiwei, and Linjuan Zhang. "Catalytic activity of bimetallic Rh/Rh-M nanosheets governed by CO spillover." Chem Catalysis 2, no. 7 (July 2022): 1512–14. http://dx.doi.org/10.1016/j.checat.2022.05.008.

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

Watanabe, Kohei, Atsushi Ueno, Xin Tao, Karel Škoch, Xiaoming Jie, Sergei Vagin, Bernhard Rieger, et al. "Reactions of an anionic chelate phosphane/borata-alkene ligand with [Rh(nbd)Cl]2, [Rh(CO)2Cl]2 and [Ir(cod)Cl]2." Chemical Science 11, no. 28 (2020): 7349–55. http://dx.doi.org/10.1039/d0sc02223c.

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

Sun, Xiaohui, Harrie Jansma, Toshihito Miyama, Rasika Dasanayake Sanjeewa Aluthge, Kenichi Shinmei, Noritoshi Yagihashi, Haruka Nishiyama, Dmitrii Osadchii, Bart van der Linden, and Michiel Makkee. "Unveiling the Structure Sensitivity for Direct Conversion of Syngas to C2-Oxygenates with a Multicomponent-Promoted Rh Catalyst." Catalysis Letters 150, no. 2 (November 6, 2019): 482–92. http://dx.doi.org/10.1007/s10562-019-03016-x.

Full text
Abstract:
Abstract Mn and Li promoted Rh catalysts supported on SiO2 with a thin TiO2 layer were synthesized by stepwise incipient wetness impregnation approach. The thin TiO2 layer on the surface of SiO2 was proved to stabilize those small Rh nanoparticles and hinder their agglomeration. The reducibility of Rh on these catalysts depends on Rh particle size as well as the position of manganese oxide, and large Rh nanoparticles with MnO on Rh nanoparticles can be only reduced at an elevated temperature. Catalyst with large Rh particles exhibits a higher CO conversion and higher products selectivity towards long chain hydrocarbons and C2-oxygenates at the expense of decreasing methane formation than a similar catalyst with smaller Rh particles. This was attributed to the synergistic effect of Mn and Li promotion and molar ratio between Rh0 and Rhδ+ sites on the surface of Rh nanoparticles. Moreover, Rh nanoparticles on MnO are proved to be more efficient in promoting hydrogenation of acetaldehyde to ethanol than its counterpart with MnO on Rh nanoparticles. Finally, in order to target high C2-oxygenates selectivity, low reaction temperature together with a low H2/CO ratio in the feed is recommended. Graphic Abstract
APA, Harvard, Vancouver, ISO, and other styles
45

Köck, Eva-Maria, Michaela Kogler, Chen Zhuo, Lukas Schlicker, Maged F. Bekheet, Andrew Doran, Aleksander Gurlo, and Simon Penner. "Surface chemistry and stability of metastable corundum-type In2O3." Physical Chemistry Chemical Physics 19, no. 29 (2017): 19407–19. http://dx.doi.org/10.1039/c7cp03632a.

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

Gao, Hanrong, and Robert J. Angelici. "Rh2Cl2(CO)4 adsorbed and tethered on gold powder: IR spectroscopic characterization and olefin hydrogenation activity." Canadian Journal of Chemistry 79, no. 5-6 (May 1, 2001): 578–86. http://dx.doi.org/10.1139/v00-190.

Full text
Abstract:
Catalysts were prepared by adsorbing Rh2Cl2(CO)4 directly on gold powder or on gold that contained the tethered ligands 2-(diphenylphosphino)ethane-1-thiol (DPET) or methyl 2-mercaptonicotinate (MMNT). Infrared (IR) studies (diffuse reflectance infrared Fourier transform (DRIFT)) of the catalyst Rh–Au prepared by adsorbing Rh2Cl2(CO)4 directly on Au indicate that a RhI(CO)2 species is present. IR studies of Rh–DPET-Au suggest that tethered cis-Rh(DPET)(CO)2Cl is the major species at relatively high Rh2Cl2(CO)4 loadings, but trans-Rh(DPET)2(CO)Cl is observable at low Rh2Cl2(CO)4 loadings. Spectral investigations of the catalyst Rh–MMNT-Au prepared by adsorbing Rh2Cl2(CO)4 on MMNT-Au suggest that tethered [cis-Rh(MMNT)2(CO)2]+Cl– and (or) Rh(MMNT)(CO)2Cl are the major species at low Rh2Cl2(CO)4 loadings, while a new unidentified species predominates at high Rh2Cl2(CO)4 loadings. All three catalysts are active 1-hexene hydrogenation catalysts under the mild conditions of 40°C and 1 atm of H2; they are much more active than Au powder or Rh2Cl2(CO)4 in solution. Of the three catalysts, Rh–Au is the most active with a maximum turnover frequency (TOF) of 800 mol H2 per mol Rh per min while its turnover (TO) is 29 600 mol H2 per mol Rh during a 2-hour run. Under the conditions of 1-hexene hydrogenation, the catalysts lose their CO ligands. Thus, it appears that a form of Rh metal on Au is the catalytically active species.Key words: catalysis, olefin hydrogenation, gold powder, tethered rhodium complexes, infrared studies, adsorption, rhodium complexes.
APA, Harvard, Vancouver, ISO, and other styles
47

Halcovitch, Nathan R., Christopher M. Vogels, Andreas Decken, and Stephen A. Westcott. "Synthesis, characterization, and reactivity of a novel thallium arylspiroboronate ester." Canadian Journal of Chemistry 87, no. 1 (January 1, 2009): 139–45. http://dx.doi.org/10.1139/v08-107.

Full text
Abstract:
Addition of 3,5-di-tert-butylcatechol (butcat) to solutions of H3B·SMe2 gave the novel diboron species B2(butcat)3 (2) in moderate to high yields. Compound 2 reacts with Tl(acac) to give butcatB(acac) (4) and Tl(Bbutcat2) (5). Attempts to abstract the chlorides from [(dppb)Rh(µ-Cl)]2 (where dppb = 1,4-bis(diphenylphosphino)buthane) using 5 led to the unusual dimer [(dppb)Rh(µ-Cl)2(µ-Tl)Rh(dppb)][Bbutcat2] (6), which contains an unsymmetrical Rh–Tl–Rh bridge.Key words: arylspiroboronate ester, non-coordinating anion, rhodium, thallium.
APA, Harvard, Vancouver, ISO, and other styles
48

Zikanová, Zuzana, Věra Vaisarová, and Jiří Hetflejš. "Hydrogenation of methyl ester of Z-α-acetamidocinnamic acid catalysed by rhodium-phosphine carboxylates." Collection of Czechoslovak Chemical Communications 51, no. 6 (1986): 1287–92. http://dx.doi.org/10.1135/cccc19861287.

Full text
Abstract:
The title reaction has been studied using the in situ catalysts prepared from [Rh(COD)Cl]2, [Rh(COD)OOCCH3]2 or Rh(COD) (acac) and a series of mono- and diphosphines. The use of (+)-N-acetylphenylalanine as a chiral carboxylato ligand gave catalytic systems with low asymmetric efficiency ( optical yields from 5 to 10%). Kinetics of the hydrogenation catalysed by [Rh(COD)OOCCH3]2 + (C6H5)2P(CH2)4P(C6H5)2 or P(n-C4H9)3 (Rh : P mol.ratio = 1 : 2) is reported.
APA, Harvard, Vancouver, ISO, and other styles
49

Simanko, Walter, Kurt Mereiter, Roland Schmid, Karl Kirchner, Anna M. Trzeciak, and Jozef J. Ziołkowski. "Rh(acac)(CO)(PR 3 ) and Rh(oxinate)(CO)(PR 3 ) complexes—substitution chemistry and structural aspects." Journal of Organometallic Chemistry 602, no. 1-2 (May 2000): 59–64. http://dx.doi.org/10.1016/s0022-328x(00)00118-2.

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

Suopanki, Aslak, Raija Polvinen, Mika Valden, and Matti Härkönen. "Rh oxide reducibility and catalytic activity of model Pt–Rh catalysts." Catalysis Today 100, no. 3-4 (February 2005): 327–30. http://dx.doi.org/10.1016/j.cattod.2004.10.020.

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