Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 13 de 13
Filtrar
Mais filtros

Base de dados
Tipo de documento
Intervalo de ano de publicação
1.
ACS Appl Nano Mater ; 7(9): 10120-10129, 2024 May 10.
Artigo em Inglês | MEDLINE | ID: mdl-38752019

RESUMO

Nanostructured niobium-titanium carbonitrides, (Nb,Ti)C1-xNx, with the cubic-rock salt structure are prepared without the use of reactive gases via thermal treatment (700-1200 °C) under nitrogen of mixtures of guanidine carbonate and ammonium niobate (V) oxalate hydrate, with addition of ammonium titanyl oxalate monohydrate as a titanium source. The bulk structure and chemical composition of the materials are characterized using powder X-ray diffraction (XRD) and powder neutron diffraction, elemental homogeneity is studied using energy dispersive spectroscopy (EDS) mapping using transmission electron microscopy (TEM), and surface chemical analysis is examined using X-ray photoelectron spectroscopy (XPS). Nanoscale crystallites of between 10 and 50 nm are observed by TEM, where EDS reveals the homogeneity of metal distribution for the mixed-metal materials. Titanium carbonitrides are found to be air sensitive, reacting with air under ambient conditions, while titanium-niobium carbonitrides are found to degrade in aqueous sulfuric acid. The niobium carbonitrides, however, show some stability toward acidic solutions. Materials are produced with composition NbC1-xNx with x between 0.35 and 0.45, and more carbon-rich materials (x ≈ 0.35) are found as the synthesis temperature is increased, as proven by Rietveld refinement of crystal structure against powder neutron diffraction data. Despite phase purity seen by diffraction and negligible bulk carbon content, XPS shows a complex surface chemistry for the NbC1-xNx materials, with evidence for Nb2O5-like oxide species in a carbon-rich environment. The NbC1-xNx prepared at 900 °C has a surface area around 50 m2 g-1, making it suitable as a catalyst support. Loading with iridium provides a material active for the oxygen evolution reaction in 0.1 M sulfuric acid, with minimal leaching of either Nb or Ir after 1000 cycles.

2.
Phys Chem Chem Phys ; 15(40): 17195-207, 2013 Oct 28.
Artigo em Inglês | MEDLINE | ID: mdl-24013445

RESUMO

This study demonstrates the utility of the novel Field Sweep Fourier Transform (FSFT) method for acquiring wideline (195)Pt NMR data from various sized Pt nanoparticles, Pt-Sn intermetallics/bimetallics used to catalyse oxidative processes in fuel cell applications, and various other related Pt3X alloys (X = Al, Sc, Nb, Ti, Hf and Zr) which can facilitate oxygen reduction catalysis. The (195)Pt and (119)Sn NMR lineshapes measured from the PtSn intermetallic and Pt3Sn bimetallic systems suggest that these are more ordered than other closely related bimetallic alloys; this observation is supported by other characterisation techniques such as XRD. From these reconstructed spectra the mean number of atoms in a Pt nanoparticle can be accurately determined, along with detailed information regarding the number of atoms present effectively in each layer from the surface. This can be compared with theoretical predictions of the number of Pt atoms in these various layers for cubo-octahedral nanoparticles, thereby providing an estimate of the particle size. A comparison of the common NMR techniques used to acquire wideline data from the I = 1/2 (195)Pt nucleus illustrates the advantages of the automated FSFT technique over the Spin Echo Height Spectroscopy (SEHS) (or Spin Echo Integration Spectroscopy (SEIS)) approach that dominates the literature in this area of study. This work also presents the first (195)Pt NMR characterisation of novel small Pt13 nanoclusters which are diamagnetic and thus devoid of metallic character. This unique system provides a direct measure of an isotropic chemical shift for these Pt nanoparticles and affords a better basis for determining the actual Knight shift when compared to referencing against the primary IUPAC shift standard (1.2 M Na2PtCl6(aq)) which has a very different local chemical environment.

3.
Dalton Trans ; 49(8): 2661-2670, 2020 Feb 25.
Artigo em Inglês | MEDLINE | ID: mdl-32048696

RESUMO

A synthesis method for the preparation of mixed manganese-ruthenium oxides is presented along with a detailed characterisation of the solids produced. The use of 1 M aqueous sulfuric acid mediates the redox reaction between KRuO4, KMnO4 and Mn2+ to form ternary oxides. At reaction temperature of 100 °C the products are mixtures of α-MnO2 (hollandite-type) and ß-MnO2 (rutile-type), with some evidence of Ru incorporation in each from their expanded unit cell volumes. At reaction temperature of 200 °C solid-solutions ß-Mn1-xRuxO2 are formed and materials with x ≤ 0.6 have been studied. The amount of Ru included in the oxide is greater than expected from the ratio of metals used in the synthesis, as determined by elemental analysis, implying that some Mn remains unreacted in solution. Powder X-ray diffraction (XRD) shows that while the unit cell volume expands in a linear manner, following Vegard's law, the tetragonal lattice parameters, and the a/c ratio, do not follow the extrapolated trends: this anisotropic behaviour is consistent with the different local coordination of the metals in the end members. Powder XRD patterns show increased peak broadening with increasing ruthenium content, which is corroborated by electron microscopy that shows nanocrystalline material. X-ray absorption near-edge spectra show that the average oxidation state of Mn in the solid solutions is reduced below +4 while that of Ru is increased above +4, suggesting some redistribution of charge. Analysis of the extended X-ray absorption fine structure provides complementary local structural information, confirming the formation of a solid solution, while X-ray photoelectron spectroscopy shows that the surface oxidation states of both Ru and Mn are on average lower than +4, suggesting a disordered surface layer may be present in the materials.

4.
Inorg Chem ; 47(24): 11745-54, 2008 Dec 15.
Artigo em Inglês | MEDLINE | ID: mdl-19006386

RESUMO

The reaction of 1,2-diketone bis(2,6-diisopropylphenyl)glyoxal (L(Ox)) with ZnCl(2) or CoCl(2) (ratio 2:1) in dry diethyl ether with 2 equiv of sodium (per transition-metal ion) afforded the neutral complexes [Zn(II)(L(*))(2)(Et(2)O)] (1) and [Co(II)(L(*))(2)(Et(2)O)] (2), which were characterized by X-ray crystallography, magnetochemistry, IR, electron paramagnetic resonance, and UV-vis spectroscopy. When 4 equiv of sodium were added, complex Na(2)(Et(2)O)(4)[Co(II)(L(Red))(2)] (4) was isolated, which included some crystals of a minor (<2%) product Na(Et(2)O)(2)[Co(III)(L(Red))(2)] (3). (L(*))(-) represents the pi-radical monoanion of the 1,2-diketone, and (L(Red))(2-) is its enediolate(2-) analogue. The electronic structures of 1, 2, and 4 have been elucidated by spectroscopy, and results are corroborated by broken-symmetry density functional theory calculations using the B3LYP functional. 1 possesses an S = 0 ground state with an excited triplet state that is 130 cm(-1) higher in energy; 2 and 4 have an S = 1/2 ground state. These complexes corroborate the notion that acyclic 1,2-diketones are redox noninnocent ligands.


Assuntos
Glioxal/análogos & derivados , Cetonas/química , Cloretos/química , Cobalto/química , Cristalografia por Raios X/métodos , Glioxal/química , Iminas/química , Ligantes , Modelos Moleculares , Conformação Molecular , Estrutura Molecular , Fotometria/métodos , Espectrofotometria/métodos , Compostos de Zinco/química
5.
Inorg Chem ; 47(23): 10935-44, 2008 Dec 01.
Artigo em Inglês | MEDLINE | ID: mdl-18959367

RESUMO

The electronic structures of chromium and vanadium centers coordinated by three reduced 1,2-diketones have been elucidated by using density functional theory (DFT) calculations and a host of physical methods: X-ray crystallography; cyclic voltammetry; ultraviolet-visible (UV-vis), nuclear magnetic resonance (NMR), and electron paramagnetic resonance (EPR) spectroscopy; and magnetic susceptibility measurements. The metal center in octahedral [CrIII(L*)3]0 (1), a CrIII (d3) ion is coupled antiferromagnetically to three monoanionic ligand pi-radicals affording an S ) 0 ground state. In contrast, Na2(Et2O)2[VIV(LRed)3] (2) (S ) 1/2), possesses a central VIV (d1) ion O,OE-coordinated to three closed-shell, doubly reduced ligands which in turn are coordinated by two Na cations enforcing a trigonal prismatic geometry at the vanadium center. 2 can be oxidized electrochemically by one and two electrons generating a monoanion, [V(L)3]1-, and a neutral species, [V(L)3]0, respectively. DFT calculations atthe B3LYP level show that the one-electron oxidized product contains an octahedral VIV ion coupled antiferromagnetically to one monoanionic ligand pi-radical [VIV(L*)(LRed)2]1- (S ) 0). In contrast, the two-electron oxidized product contains a VIII ion coupled antiferromagnetically to three ligand pi-radicals in an octahedral field[VIII(L*)3]0 (S ) 1/2).


Assuntos
Cromo/química , Elétrons , Cetonas/química , Compostos Organometálicos/química , Vanádio/química , Cristalografia por Raios X , Espectroscopia de Ressonância de Spin Eletrônica , Ligantes , Modelos Moleculares , Conformação Molecular , Oxirredução , Teoria Quântica , Temperatura
6.
Chem Commun (Camb) ; (1): 85-7, 2007 Jan 07.
Artigo em Inglês | MEDLINE | ID: mdl-17279269

RESUMO

The reaction of the "digermyne" Ar'GeGeAr' (Ar' = C6H3-2,6(C6H3-2,6-Pr(i)2)2; Ge-Ge = 2.2850(6) A) with mesityl isocyanide affords the bis adduct [Ar'GeGeAr'(CNMes)2] which results in the conversion of a Ge-Ge multiple bond to a long Ge-Ge single bond (= 2.6626(8) A).

7.
Chem Commun (Camb) ; (42): 4339-41, 2007 Nov 14.
Artigo em Inglês | MEDLINE | ID: mdl-17957279

RESUMO

The reduction of the bulky 1,2-diketone bis(2,6-diisopropylphenyl)glyoxal () and FeBr(2) with 1.5 equivalents of Na results in a Class 2 mixed valent H.S. Fe(II) L.S. Fe(III) complex (2) with two five-coordinate Fe centres which are antiferromagnetically exchange-coupled to give a total spin S(t) = 3/2 ground state and an S(t) = 5/2 excited state that are separated by about 25 cm(-1) (for Delta(J) approximately 5J).


Assuntos
Compostos Férricos/química , Compostos Ferrosos/química , Glioxal/análogos & derivados , Cristalografia por Raios X , Glioxal/química , Indicadores e Reagentes , Espectroscopia de Ressonância Magnética , Magnetismo , Modelos Moleculares , Oxirredução , Espectrofotometria Ultravioleta , Espectroscopia de Mossbauer
8.
Chem Commun (Camb) ; (48): 6041-3, 2005 Dec 28.
Artigo em Inglês | MEDLINE | ID: mdl-16333521

RESUMO

The reactions of the digermanium and ditin alkyne analogues Ar'MMAr' (M = Ge or Sn) with R2NO, (R2NO = Me2C(CH2)3CMe2NO or N2O), result in complete MM bond cleavage to afford the germylene :Ge(Ar')ONR2 or the germanium(II) or tin(II) hydroxides {M(Ar')(micro-OH)}2.

10.
Inorg Chem ; 46(8): 2971-8, 2007 Apr 16.
Artigo em Inglês | MEDLINE | ID: mdl-17338516

RESUMO

Reaction of the hindered phosphino- and arsinoboranes, Ar*Pn(H)-B(Br)Tmp (Ar* = -C6H3-2,6-(C6H2-2,4,6-iPr3)2; Tmp = 2,2,6,6-tetramethylpiperidino; Pn = P and As, 1 and 3, respectively) with 4-dimethylaminopyridine, DMAP, afforded the boranylidenephosphane and arsane, Ar*Pn=B(DMAP)Tmp (Pn = P and As, 2 and 4) as deep red-purple solids. The analogous aminoboranes Ar'N(H)-B(X)Tmp (Ar' = -C6H3-2,6-(C6H2-2,4,6-Me3)2; X = Cl and Br; 5 and 6) did not display any reactivity with DMAP, but in the presence of the amide base, Na[N(SiMe3)2], the clean formation of the uncomplexed iminoborane Ar'NBTmp (7) was observed. Attempts to generate an Sb=B bond were unsuccessful, as the required stibinoborane precursor, Ar*Sb(H)-B(Br)Tmp, could not be prepared; in place of clean Sb-B bond formation, the reduced product Ar*Sb=SbAr* was obtained. All compounds were characterized spectroscopically, and the X-ray crystal structures of 1, 2, 4, 6, and 7 were determined.

11.
Inorg Chem ; 45(22): 9132-6, 2006 Oct 30.
Artigo em Inglês | MEDLINE | ID: mdl-17054374

RESUMO

The "distannynes" Ar'SnSnAr' (Ar' = C6H3-2,6(C6H3-2,6-Pr(i)2)2) and ArSnSnAr (Ar = C6H3-2,6(C6H2-2,4,6-Pr(i)3)2) were examined by solid-state (119)Sn NMR and Mössbauer spectroscopy. The two compounds display substantially different spectroscopic parameters, while differing only in the absence (Ar'SnSnAr') or presence (ArSnSnAr) of a para-Pr(i) group in the flanking aryl rings of their terphenyl substituents. The spectroscopic differences can be interpreted in terms of a more trans-bent geometry and a longer Sn-Sn bond for ArSnSnAr in comparison to the wider Sn-Sn-C angle (125.24(7) degrees ) and shorter Sn-Sn bond length (2.6675(4)A) determined from the crystal structure of Ar'SnSnAr'. The differences are consistent with previously published calculations by Nagase and Takagi for ArSnSnAr.

12.
J Am Chem Soc ; 127(35): 12232-3, 2005 Sep 07.
Artigo em Inglês | MEDLINE | ID: mdl-16131195

RESUMO

The germanium alkyne analogue Ar'GeGeAr' (1, Ar' = C6H3-2,6(C6H3-2,6-Pri2)2) reacts with 1, 2, or 3 equiv of dihydrogen at room temperature, and at 1 atm pressure, to afford a mixture of the products Ar'HGeGeHAr' (2), Ar'H2GeGeH2Ar' (3), or Ar'GeH3 (4). The relative amounts of each product are governed by the number of equivalents of hydrogen used. A mechanism for the initial step in the reaction is proposed. The appearance of 4 among the reaction products was accounted for in terms of either its dissociation to monomers or isomerization to the bridged Ar'Ge(mu-H)2GeAr'. The reactions were monitored by 1H NMR spectroscopy. The products 2, 3, and 4 were characterized by X-ray crystallography, and 4 was synthesized independently by the reduction of Ar'Ge(OMe)3. These reactions represent the first direct addition of hydrogen to a closed shell unsaturated main group compound under ambient conditions.

13.
J Am Chem Soc ; 127(49): 17530-41, 2005 Dec 14.
Artigo em Inglês | MEDLINE | ID: mdl-16332105

RESUMO

Reactions of the alkyne analogues Ar'EEAr' (Ar' = C6H3-2,6(C6H3-2,6-Pr(i)2)2; E = Ge (1); Sn (2)) with unsaturated molecules are described. Reaction of 1 and 2 with azobenzene afforded the new hydrazine derivatives Ar'E{(Ph)NN(Ph)}EAr' (E = Ge (3); Sn (4)). Treatment of 1 with Me3SiN3 gave the cyclic singlet diradicaloid Ar'Ge{mu2-(NSiMe3)}2GeAr' (5), whereas 2 afforded the monoimide bridged Ar'Sn{mu2-N(SiMe3)}SnAr' (6). Reaction of 1 with t-BuNC or PhCN yielded the adduct Ar'GeGe(CNBu(t))Ar' (7) or the ring compound (8). In contrast, the tin compound 2 did not react with either t-BuNC or PhCN. Treatment of 1 with N2CH(SiMe3) generated Ar'Ge{mu2-CH(SiMe3)}{mu2:eta2-N2CH(SiMe3)}{mu2-N2CH(SiMe3)}GeAr' (9) which contains ligands in three different bridging modes and no Ge-Ge bonding. Reaction of 1 with an excess of N(2)O gave a germanium peroxo species Ar'(HO)Ge(mu2-O)(mu2:eta2-O2)Ge(OH)Ar' (10) which features a ring. Oxidation of 1 by tetracyanoethylene (TCNE) led to cleavage of the Ge-Ge bond and formation of a large multiring system of formula Ar'Ge3+{(TCNE)2-}3{(GeAr')+}3. The digermyne 1 also reacted with 1 equiv of PhCPh to give the 1,2-digermacyclobutadiene 12, which has a ring, and with Me(3)SiCCH or PhCC-CCPh to activate a flanking C6H3-2,6-Pr(i)2 ring and give the tricyclic products 13 and 14. The "distannyne" 2 did not react with these acetylenes. Overall, the experiments showed that 1 is highly reactive toward unsaturated molecules, whereas the corresponding tin congener 2 is much less reactive. A possible explanation of the reactivity differences in terms of the extent of the singlet diradical character of the Ge-Ge and Sn-Sn bonds is discussed.

SELEÇÃO DE REFERÊNCIAS
Detalhe da pesquisa