共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
3.
Rodrigo J. Lobo‐Lapidus Bruce C. Gates 《Chemistry (Weinheim an der Bergstrasse, Germany)》2010,16(37):11386-11398
Two carbonyl complexes of rhenium, [HRe(CO)5] and [CH3Re(CO)5], were used to probe surface sites of TiO2 (anatase). These complexes were adsorbed from the gas phase onto anatase powder that had been treated in flowing O2 or under vacuum to vary the density of surface OH sites. Infrared (IR) spectra demonstrate the variation in the number of sites, including Ti+3? OH and Ti+4? OH. IR and extended X‐ray absorption fine structure (EXAFS) spectra show that chemisorption of the rhenium complexes led to their decarbonylation, with formation of surface‐bound rhenium tricarbonyls, when [HRe(CO)5] was adsorbed, or rhenium tetracarbonyls, when [CH3Re(CO)5] was adsorbed. These reactions were accompanied by the formation of water and surface carbonates and removal of terminal hydroxyl groups associated with Ti+3 and Ti+4 ions on the anatase. Data characterizing the samples after adsorption of [HRe(CO)5] or [CH3Re(CO)5] determined a ranking of the reactivity of the surface OH sites, with the Ti+3? OH groups being the more reactive towards the rhenium complexes but the less likely to be dehydroxylated. The two rhenium pentacarbonyl probes provided complementary information, suggesting that the carbonate species originate from carbonyl ligands initially bonded to the rhenium and from hydroxyl groups of the titania surface, with the reaction leading to the formation of water and bridging hydroxyl groups on the titania. The results illustrate the value of using a family of organometallic complexes as probes of oxide surface sites. 相似文献
4.
Villanneau R Proust A Robert F Gouzerh P 《Chemistry (Weinheim an der Bergstrasse, Germany)》2003,9(9):1982-1990
The reactions of neutral or cationic manganese carbonyl species towards the oxo-nitrosyl complex [Na(MeOH)[Mo(5)O(13)(OCH(3))(4)(NO)]](2-) have been investigated in various conditions. This system provides an unique opportunity for probing the basic reactions involved in the preparation of solid oxide-supported heterogeneous catalysts, that is, mobility of transition-metal species at the surface and dissolution-precipitation of the support. Under nitrogen and in the dark, the reaction of in situ generated fac-[Mn(CO)(3)](+) species with (nBu(4)N)(2)[Na(MeOH)-[Mo(5)O(13)(OMe)(4)(NO)]] in MeOH yields (nBu(4)N)(2)[Mn(CO)(3)(H(2)O)[Mo(5)O(13)(OMe)(4)(NO)]] at room temperature, while (nBu(4)N)(3)[Na[Mo(5)O(13)(OMe)(4)(NO)](2)[Mn(CO)(3)](2)] is obtained under reflux. The former transforms into the latter under reflux in methanol in the presence of sodium bromide; this involves the migration of the fac-[Mn(CO)(3)](+) moiety from a basal kappa(2)O coordination site to a lateral kappa(3)O site. Oxidation and decarbonylation of manganese carbonyl species as well as degradation of the oxonitrosyl starting material and reaggregation of oxo(methoxo)molybdenum fragments occur in non-deareated MeOH, and both (nBu(4)N)(4)[Mn(H(2)O)(2)[Mo(5)O(16)(OMe)(2)](2)[Mn(CO)(3)](2)] and (nBu(4)N)(4)[Mn(H(2)O)(2)[Mo(5)O(13)(OMe)(4)(NO)](2)] as well as (nBu(4)N)(2)[MnBr[Mo(5)O(13)(OMe)(4)(NO)]] have been obtained in this way. The rhenium analogue (nBu(4)N)(2)[Re(CO)(3)(H(2)O)[Mo(5)O(13)(OMe)(4)(NO)]] has also been synthesized. The crystal structures of (nBu(4)N)(2)[Re(CO)(3)(H(2)O)[Mo(5)O(13)(OMe)(4)(NO)]], (nBu(4)N)(3)[Na[Mo(5)O(13)(OMe)(4)(NO)](2)[Mn(CO)(3)](2)], (nBu(4)N)(4)[Mn(H(2)O)(2)[Mo(5)O(16)(OMe)(2)](2)[Mn(CO)(3)](2)], (nBu(4)N)(4)[Mn(H(2)O)(2)[Mo(5)O(13)(OMe)(4)(NO)](2)] and (nBu(4)N)(2)[MnBr[Mo(5)O(13)(OMe)(4)(NO)]] have been determined. 相似文献
5.
Oxacalix[3]arenes p-methyloxacalix[3]arene (L(1)), p-isopropyloxacalix[3]arene (L2), and p-ethoxycarbonyloxacalix[3]arene (L3) are able to bind the Re(I)(CO)3 moiety with two of their three phenol-O atoms and one of their ether-O atoms. The monoanionic complexes were isolated in the salts (DBUH)[Re(CO)3(L1H-2)].L1 (1) and (NEt4)[Re(CO)3(L2H-2)].L2.0.5 MeCN (2) (DBU = 1,8-diazabicyclo[5.4.0]undec-7-ene). Over the course of its reaction with (NEt4)(2)[Re(CO)3Br(3)] and DBU, p-ethoxycarbonyloxacalix[3]arene decomposes to form [{Re(CO)3(L4H-2)}2] (3) {L4 = 1-(5-ethoxycarbonyl-2-hydroxy-3-hydroxymethyl-benzyl)-2,3,4,6,7,8,9,10-octahydro-pyrimido[1,2-a]azepin-1-ium. The expected monoanion [Re(CO)3(L3H-2)]- (4) was identified by 13C NMR and mass spectra. 相似文献
6.
Bernhardt E Finze M Willner H Lehmann CW Aubke F 《Chemistry (Weinheim an der Bergstrasse, Germany)》2006,12(32):8276-8283
The reaction of [Co(2)(CO)(8)] with (CF(3))(3)BCO in hexane leads to the Lewis acid-base adduct [Co(2)(CO)(7)CO--B(CF(3))(3)] in high yield. When the reaction is performed in anhydrous HF solution [Co(CO)(5)][(CF(3))(3)BF] is isolated. The product contains the first example of a homoleptic metal pentacarbonyl cation with 18 valence electrons and a trigonal-bipyramidal structure. Treatment of [Co(2)(CO)(8)] or [Co(CO)(3)NO] with NO(+) salts of weakly coordinating anions results in mixed crystals containing the [Co(CO)(5)](+)/[Co(CO)(2)(NO)(2)](+) ions or pure novel [Co(CO)(2)(NO)(2)](+) salts, respectively. This is a promising route to other new metal carbonyl nitrosyl cations or even homoleptic metal nitrosyl cations. All compounds were characterized by vibrational spectroscopy and by single-crystal X-ray diffraction. 相似文献
7.
Zampella G Bruschi M Fantucci P Razavet M Pickett CJ De Gioia L 《Chemistry (Weinheim an der Bergstrasse, Germany)》2005,11(2):509-520
A bridging carbonyl intermediate with key structural elements of the diiron sub-site of all-iron hydrogenase has been experimentally observed in the CN/CO substitution pathway of the {2Fe3S} carbonyl precursor, [Fe(2)(CO)(5){MeSCH(2)C(Me)(CH(2)S)(2)}]. Herein we have used density functional theory (DFT) to dissect the overall substitution pathway in terms of the energetics and the structures of transition states, intermediates and products. We show that the formation of bridging CO transitions states is explicitly involved in the intimate mechanism of dicyanation. The enhanced rate of monocyanation of {2Fe3S} over the {2Fe2S} species [Fe(2)(CO)(6){CH(2)(CH(2)S)(2)}] is found to rest with the ability of the thioether ligand to both stabilise a mu-CO transition state and act as a good leaving group. In contrast, the second cyanation step of the {2Fe3S} species is kinetically slower than for the {2Fe2S} monocyanide because the Fe2 atom is deactivated by coordination of the electron-donating thioether group. In addition, hindered rotation and the reaction coordinate of the approaching CN(-) group, are other factors which explain reactivity differences in {2Fe2S} and {2Fe3S} systems. The intermediate species formed in the second cyanation step of {2Fe3S} species is a mu-CO species, confirming the structural assignment made on the basis of FT-IR data (S. J. George, Z. Cui, M. Razavet, C. J. Pickett, Chem. Eur. J. 2002, 8, 4037-4046). In support of this we find that computed and experimental IR frequencies of structurally characterised {2Fe3S} species and those of the bridging carbonyl intermediate are in excellent agreement. In a wider context, the study may provide some insight into the reactivity of dinuclear systems in which neighbouring group on-off coordination plays a role in substitution pathways. 相似文献
8.
Reactions of the compound cis-[Rh2(DTolF)2(CH3CN)6](BF4)2, a formamidinate derivative of the class of antitumor compounds [Rh2(O2CR)4] (R=Me, Et, Pr), with 9-ethylguanine (9-EtGuaH) or the dinucleotide d(GpG) proceed by substitution of the acetonitrile groups, with the guanine bases spanning the Rh--Rh bond, in a bridging fashion, through sites N7/O6. In the case of 9-EtGuaH, both head-to-head (HH) and head-to-tail (HT) isomers are formed, whereas with the tethered bases in d(GpG), only one right-handed conformer HH1R [Rh2(DTolF)2{d(GpG)}] is present in solution. For both cis-[Rh2(DTolF)2(9-EtGuaH)2](BF4)2 and [Rh2(DTolF)2{d(GpG)}], the absence of N7 protonation at low pH and the substantial decrease of the pKa values for N1-H deprotonation, support N7/O6 binding of the bases to the dirhodium core. The N7/O6 binding of the bases is further corroborated by the downfield shift by Deltadelta approximately 4.0 ppm of the 13C NMR resonances for the C6 nuclei as compared to the corresponding resonances of the free ligands. The HH arrangement of the guanine bases in [Rh2(DTolF)2{d(GpG)}] is indicated by the intense H8/H8 ROE cross-peaks in the 2D ROESY NMR spectrum. Complete characterization of the [Rh2(DTolF)2{d(GpG)}] conformer by 2D NMR spectroscopy supports anti-orientation and N (C3'-endo) conformation for both deoxyribose residues. The N-pucker for the 5'-G base is universal in such cross-links, but it is very unusual for platinum and unprecedented for dirhodium HH cross-linked adducts to have both deoxyribose residues in the N-type conformation. The bulk, the nonlabile character, and the electron-donating ability of the formamidinate bridging groups spanning the dirhodium core affect the nature of the preferred dirhodium DNA adducts. Molecular modeling studies performed on [Rh2(DTolF)2{d(GpG)}] corroborate the structural features obtained by NMR spectroscopy. 相似文献
9.
10.
11.
12.
13.
14.
Capon JF Ezzaher S Gloaguen F Pétillon FY Schollhammer P Talarmin J 《Chemistry (Weinheim an der Bergstrasse, Germany)》2008,14(6):1954-1964
Electrochemical investigations on a structural analogue of the [2Fe](H) subsite of [FeFe]H(2)ases, namely, [Fe(2)(CO)(6){micro-SCH(2)N(CH(2)CH(2)- OCH(3))CH(2)S}] (1), were conducted in MeCN/NBu(4)PF(6) in the presence of HBF(4)/Et(2)O or HOTs. Two different catalytic proton reduction processes operate, depending on the strength and the concentration of the acid used. The first process, which takes place around -1.2 V for both HBF(4)/Et(2)O and HOTs, is limited by the slow release of H(2) from the product of the {2 H(+)/2 e} pathway, 1-2H. The second catalytic process, which occurs at higher acid concentrations, takes place at different potentials depending on the acid present. We propose that this second mechanism is initiated by protonation of 1-2H when HBF(4)/Et(2)O is used, whereas the reduction of 1-2H is the initial step in the presence of the weaker acid HOTs. The potential of the second process, which occurs around -1.4 V (reduction potential of 1-3H(+)) or around -1.6 V (the reduction potential of 1-2H) is thus dependent on the strength of the available proton source. 相似文献
15.
Synthesis of Carboxylate Substituted Rhenium Gold Metallatetrahedranes Re2(AuPPh3)2(μ-PCy2)(CO)7(η1-OC(R)O) (R = H, Me, CF3, Ph, 3,4-(OMe)2C6H3) The reaction of the in situ prepared salt Li[Re2(μ-H)(μ-PCy2)(CO)7(ax-C(Ph)O)] ( 2 ) with 1,5 equivalents of monocarboxylic acid RCOOH (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e ) in tetrahydrofruan (THF) solution at 60 °C gives within 4 h under release of benzaldehyde (PhCHO) the η1-carboxylate substituted dirhenium salt Li[Re2(μ-H)(μ-PCy2)(CO)7(η1-OC(R)O)] (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e )) in almost quantitative yield. The lower the pKa value of the respective carboxylic acid the faster the reaction proceeds. It was only in the case of CF3COOH possible to prove the formation of the hydroxycarbene complex Re2(μ-H)(μ-PCy2)(CO)7(=C(Ph)OH) ( 5 ) prior to elimination of PhCHO. The new compounds 4 a–4 e were only characterized by 31P NMR and ν(CO) IR spectroscopy as they are only stable in solution. They are converted with two equivalents of BF4AuPPh3 at 0 °C in a so-called cluster expansion reaction into the heterometallic metallatetrahedrane complexes Re2(AuPPh3)2(μ-PCy2)(CO)7(η1-OC(R)O) (R = H ( 7 a ), Me ( 7 b ), CF3 ( 7 c ), Ph ( 7 d ), 3,4-(OMe)2C6H3 ( 7 e )) (yield 47–71% ). The expected precursor complexes of 7 a–7 e Li[Re2(AuPPh3)(μ-PCy2)(CO)7(η1-OC(R)O] ( 8 ) were not detected by NMR and IR spectroscopy in the course of the reaction. Their existence was retrosynthetically proved by the reaction of 7 b with an excess of the chelating base TBD (1,5,7-Triazabicyclo[4.4.0]dec-5-en) forming [(TBD)xAuPPh3][Re2(AuPPh3)(μ-PCy2)(CO)7(η1-OC(Me)O] ( 8 b ) in solution. The η1-bound carboxylate ligand in 7 a–7 e can photochemically be converted into a μ-bound ligand in Re2(AuPPh3)2(μ-PCy2)(μ-OC(R)O)(CO)6 (R = H ( 9 a ), Me ( 9 b ), CF3 ( 9 c ), Ph ( 9 d ), 3.4-(MeO)2C6H3 ( 9 e )) under release of one equivalent CO. All isolated cluster complexes were characterized and identified by the following analytical methods: elementary analysis, NMR (1H, 31P) spectroscopy, ν(CO) IR spectroscopy and in the case of 7 d and 9 b by X-ray structure analysis. 相似文献
16.
《Mendeleev Communications》2021,31(6):810-812
The reaction of 1,2,4,5-tetrakis(diphenylphosphino)pyridine (L) with Re(CO)5Br (in a molar ratio of 1:2) leads to the bis-chelated complex [Re2(CO)6(L)Br2] in 95% yield. At ambient temperature, the solid complex exhibits green phosphorescence (λmax = 535 nm) with a quantum yield of 12% and a lifetime of 90 μs. 相似文献
17.
18.
Bourrez M Molton F Chardon-Noblat S Deronzier A 《Angewandte Chemie (International ed. in English)》2011,50(42):9903-9906
Manganese at work: Carbonyl bipyridyl complexes based on manganese, a non-noble abundant and inexpensive metal, have been proved to be excellent molecular catalysts for the selective electrochemical reduction of CO(2) to CO under mild conditions. Another advantage of manganese complexes over rhenium complexes is that these catalysts operate at markedly less overpotential (0.40?V gain). 相似文献
19.