首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The synthesis of arene-ruthenium(II) and C5Me5-rhodium(III) and -iridium(III) complexes of chiral arene-chromium-tricarbonyl-based PP and PN ligands is described. Three complexes were characterized in the solid state by X-ray structural analysis. The complexes were tested in the catalytic hydrogen transfer reactions as well as in the kinetic resolution of racemic alcohols, where some complexes showed good conversion, but low enantioselectivity.  相似文献   

2.
This review highlights our theoretical studies of regioselectivities of Rh-catalyzed (5 + 2) cycloadditions, Ni-catalyzed reductive alkyne-aldehyde couplings, and Rh-catalyzed hydrogenative couplings of diynes and carbonyl partners. Factors that control the regioselectivities in these reactions are analyzed; these include steric repulsions involving the substrate and ligands, electronic effects dictated by metal-substrate interactions, and directing effects of conjugated alkenyl and alkynyl groups involving coordination to the metal.  相似文献   

3.
Interactions between guanidinium cations and the sulfonate groups on the phosphine [PPh2C6H4-m-SO3] have been exploited to incorporate iridium(I) centres into hydrogen-bonded networks. The crystal structure of [C(NH2)3]2{trans-[IrCl(CO)(PPh2C6H4-m-SO3)2]} (4) contains hexagonal guanidinium sulfonate (GS) sheets in which both of the sulfonate groups from each complex anion form hydrogen bonds within the same sheet. The crystal structures of [C(NH2)2(NHMe)][PPh2C6H4-m-SO3] (5) and [C(NH2)2(NHEt)][PPh2C6H4-m-SO3] (6) reveal that the GS sheets can tolerate the loss of one hydrogen bond donor, though twisting occurs to accommodate the alkyl group. However, the crystal structure of [C(NH2)2(NMe2)][PPh2C6H4-m-SO3] (7) shows that ribbon structures are formed instead of sheets when two hydrogen bond donors are lost. The compound [C(NH2)2(NHMe)]2{trans-[IrCl(CO)(PPh2C6H4-m-SO3)2]} · 3/8H2O (8) contains hydrogen-bonded cylinders as opposed to sheets. This is a likely consequence of a mismatch between the intramolecular S?S distance present in the anion, and the closer S?S distance present in a twisted GS sheet such as that in 5. The crystal structures of [C(NH2)2(NHEt)][P(O)Ph2C6H4-m-SO3] (9) and [C(NH2)2(NMe2)][P(O)Ph2C6H4-m-SO3] · H2O (10) show that the phosphine oxide group successfully competes with the sulfonate as a hydrogen bond acceptor. The crystal structure of 9 contains hydrogen-bonded ribbons that are interlinked through the anions which act as pillars to form a layer structure. In contrast, the crystal structure of 10 contains hydrogen-bonded sheets that involve cations, sulfonate groups, phosphine oxides and the included water molecule. These sheets are linked into a three-dimensional network through the anion pillars.  相似文献   

4.
Novel as well as known C,N-palladacyclic neutral complexes were tested as precatalyst in the Heck reaction between bromobenzene and styrene under aerobic conditions. The catalytic system Pd(II) complex/K2CO3/EGME-H2O (EGME = ethylene glycol monomethyl ether) showed to be highly efficient. Best performances of the catalysts were achieved by controlling the amount of water: generally a H2O content within the 25-50% v/v range resulted in the highest conversion of the substrates into trans-stilbene. As a matter of fact, bromobenzene and styrene can be converted quantitatively using only 0.01 mol% of precatalyst with very high regioselectivity for the trans product and with a TOF of 10 000 h−1. In the absence of water, all complexes were less efficient and differences in their activity were found, while such a differentiation disappeared when water was added.  相似文献   

5.
Compounds 1-6 of the type MoO2X2L2 (X=F, Cl, Br; L=OPMePh2, OPPh3) have been prepared in order to investigate the variation in catalytic activity with changes in electronic and steric properties. All six complexes catalyze the epoxidation of cyclohexene with tert-butylhydroperoxide, and the species with X=Cl and L=OPMePh2 (2) displays the best activity with 83% conversion and 90% selectivity in one hour at ambient atmosphere. These inexpensive and easily prepared dioxo catalysts are stable to air and water. Reactions of the dioxo compounds with H2O2 and t-BuOOH have also been carried out. The structures of MoO2F2(OPMePh2)2 (1) and the product of its reaction with H2O2, MoO(O2)2(OPMePh2)2 (7) have been solved by single crystal X-ray diffraction.  相似文献   

6.
The dinuclear V(V) complexes (VOL)2O (L = SAE (1), SAMP (2), SAP (3)) have been synthesized from VO(acac)2 and the corresponding tridentate ligands LH2 in methanol under reflux conditions and subsequent air oxidation in organic solvent. They have been characterized by IR and NMR spectroscopy, by thermogravimetric analysis, and by single crystal X-ray diffraction for 1 and 2. DFT calculations were carried out for a better understanding of the vibrational pattern, principally the V-O related vibrations. Complex [VO(SAP)]2O (3) catalyzes the epoxidation of cyclooctene by TBHP in water in the absence of any added solvent with good selectivity.  相似文献   

7.
Gold catalysis is a convenient tool to oxidatively functionalize alkyne into a range of valuable compounds. In this article, we report a new access to isochroman-4-one and 2H-pyran-3(6H)-one derivatives that involves a gold-catalyzed oxidative cycloalkoxylation of an alkyne in the presence of a pyridine N-oxide. The reaction proceeds under mild conditions, is relatively efficient and exhibits a high functional group compatibility.  相似文献   

8.
Six new adducts of the form AgX:PPh3:H2C(pzx)2 (1:1:1) (H2C(pzx)2 = H2C(pz)2 = bis(pyrazolyl)methane or H2C(pzMe2)2 = bis(3,5-dimethylpyrazolyl)methane; X = ClO4, NO3, SO3CF3) have been synthesized and characterized by analytical, spectroscopic (IR, far-IR, 1H and 31P NMR) and two of them also by single crystal X-ray diffraction studies for comparison with counterpart adducts with 2,2′-bipyridyl (‘bpy’) derivatives reported in a previous paper, the bpy-derived ligands forming five-membered chelate rings, while the present H2C(pzx)2 should, potentially, form six-membered rings. Such is the case, the two adducts exhibiting quasi-planar N2AgP coordination environments, perturbed by the approach of the oxyanion, unidentate in the case of the perchlorate but, in the case of the nitrate, an interesting disordered aggregate of differing unidentate modes.  相似文献   

9.
The dimer [Ir(μ-Cl)(C8H14)2]2 reacts with the ligands (S)-(C5H4CH2CH(Ph)PPh2)Li and (R)-(C5H4CH(Cy)CH2PPh2)Li to give (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(C8H14)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(C8H14)], which upon treatment with CH3I at room temperature afford the cationic iridium(III) compounds (S,SIr)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(C8H14)][I] as a single diastereomer, and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(C8H14)][I] as a 9:1 mixture of two diastereomers. If the oxidative addition reaction is performed at reflux in methylene chloride, the starting complexes convert to the neutral compounds (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(I)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(I)] as 1.6:1 and 3.3:1 mixtures of diastereoisomers, respectively. Carbonyl iridium complexes are synthesized by reacting [IrCl(CO)(PPh3)2] with the ligands to afford (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CO)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CO)]. They give upon treatment with CH3I the cationic species (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(CO)][I] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(CO)][I] as 1.6:1 and 3:1 mixture of diastereomers, respectively. No migratory-insertion of the methyl group into the carbonyl-metal bond has been observed even after prolonged heating.  相似文献   

10.
Schiff bases of 2-hydroxybenzophenone (HBP) (C6H5)(2-HOC6H4)CN(CH2)nEAr (L1/L2: E = S, Ar = Ph, n = 2/3; L3/L4: E = Se, Ar = Ph, n = 2/3; L5/L6: E = Te, Ar = 4-MeOC6H4, n = 2/3) and their complexes [PdCl(L-H)] (L = L1L6; 1, 2, 3, 5, 7, 11), [PtCl(L3-H/L5-H)] (4/8), [PtCl2(L4/L6)2] (6/12), [(p-cymene)RuCl(L5/L6)]Cl (9/13) and [HgBr2(L5/L6)2] (10/14) have been synthesized and characterized by proton, carbon-13, selenium-77 and tellurium-125 NMR, IR and mass spectra. Single crystal structures of L1, 1, 3, 4, 5 and 7 were solved. The Pd-E bond distances (Å): 2.2563(6) (E = S), 2.3575(6)−2.392(2) (E = Se); 2.5117(5)−2.5198(5) (E = Te) are near the lower end of the bond length range known for them. The Pt-Se bond length, 2.3470(8) Å, is also closer to the short values reported so far. The Heck and Suzuki reaction were carried out using complexes 1, 3, 5 and 7 as catalysts under aerobic condition. The percentage yields for trans product in Heck reaction were found upto 85%.  相似文献   

11.
The reaction of the monoalkyl complex trans-[Pt(DMSO)2Cl(CH3)] with a large variety of heterocyclic nitrogen bases L, in chloroform solution, leads to the formation of uncharged complexes of the type [Pt(DMSO)(L)Cl(CH3)], containing four different groups coordinated to the metal center. Only two out of the three different possible isomers were detected in solution. These two trans(C,N) and cis(C,N) species can be unambiguously identified through 1H NMR spectroscopy. For the trans(C,N) isomers, average values of 2JPtH=75±4 Hz and 3JPtH=36±4 Hz have been observed for the coordinated methyl and DMSO ligands, respectively. In the case of the cis(C,N) isomers, these values increase to 2JPtH=83±2 Hz, and decrease to 3JPtH=26±3 Hz due to the mutual exchange of ligands in trans position to CH3 and DMSO. In the case of bulky asymmetric ligands, such as quinoline, 2-quinolinecarboxaldehyde, 2-methylquinoline, 5-aminoquinoline, 2-phenylpyridine and 2-chloropyridine, slow rotation of the hindered group around the Pt---N bond makes the coordinated DMSO ligand prochiral. NMR experiments have shown that the first reaction product is the trans(C,N) isomer as a consequence of the very fast removal of one DMSO ligand by the nitrogen bases from the starting complex trans-[Pt(DMSO)2Cl(CH3)]. This trans kinetic product undergoes a geometrical conversion into the more stable cis(C,N) isomer through the intermediacy of fast exchanging aqua-species. The rate of isomerization and the relative stability of the two isomers depends essentially on the rate of aquation and on the steric congestion imposed by the new L ligand on the metal.  相似文献   

12.
The iridium 1,1,1-tris(diphenylphosphinomethyl)ethane (triphos) complexes [{κ2(C1,C4)-CRCRCRCR}{CH3C(CH2PPh2)3}Ir(NCMe)]BF4 (2-NCMe, R = CO2Me) and [{κ2(C1,C4)-CRCRCRCR}{CH3C(CH2PPh2)3}Ir(CO)]BF4 (2-CO, R = CO2Me) serve as models for proposed iridium-vinylidene intermediates of relevance to the [2 + 2 + 1] cyclotrimerization of alkynes. The solid-state structures of 2-NCMe, 2-CO, and [κ2(C1,C4)-CRCRCRCR]{CH3C(CH2PPh2)3}Ir(Cl) (2-Cl), were determined by X-ray crystallography.  相似文献   

13.
A new class of chiral beta-amino disulfides was synthesized from readily available and inexpensive starting materials by a straightforward method and their abilities as ligands were examined in the enantioselective addition of diethylzinc to aldehydes. Enantiomeric excesses of up to 99% have been obtained using 0.5 mol % of the chiral catalysts.  相似文献   

14.
The catalytic activity of the zinc(II) complexes of calix[4]arenes decorated with 1,5,9-triazacyclododecane ligands at the 1,2-, 1,3-, and 1,2,3-positions of the upper rim was investigated in the basic methanolysis (pH 10.4) of aryl acetates functionalised at the meta- and para-positions with a carboxylate anchoring group. Michaelis-Menten kinetics and turnover catalysis were observed. High rate accelerations, up to more than 104-fold at 0.2 mM catalyst, were recorded in the most favourable catalyst-substrate combinations. The order of catalytic efficiency of regioisomeric bimetallic complexes is 1,2-vicinal ? 1,3-distal, resulting from a significant degree of synergism between metal ions in the former, and a complete lack in the latter. The moderately higher efficiency of the trimetallic compared with the 1,2-vicinal bimetallic catalyst provides an indication of a possible cooperation of three zinc(II) ions in the catalysis.  相似文献   

15.
A chelate cationic rhodium(I) complex with a hemilabile amino- and sulfur-containing phosphinite ligand has been synthesized and, according to the NMR data, the ligand is bound to the metal in a P,S-bidentate coordination mode without any Rh-N interaction. This complex efficiently catalyzes the hydroformylation of styrene. The chelate rhodium complex with the analogous ligand without the amino group has also been synthesized and examined as a catalyst for the same hydroformylation reaction. The reaction rate is higher using the former complex compared to the latter one without the amino group, with, however, a slightly lower regioselectivity towards the formation of the branched aldehyde.  相似文献   

16.
The first chiral bis(pyridine) N-C(H)-N pincer ligand, (5S,7S)-1,3-bis(6,6-dimethyl-5,6,7,8-tetrahydro-5,7-methanquinolin-2-yl)benzene (HL) has been synthesized and characterized by a thorough 1H NMR analysis. Reaction of HL with K2[PtCl4] in acetic acid gives [Pt(L)Cl] (1), where L acts as a tridentate N-C-N pincer ligand. The analogous palladium(II) derivatives [Pd(L)Cl] (2), and [Pd(L)(OAc)] (3), were first prepared through a transmetalation reaction between Pd(OAc)2 and the organomercury compound [Hg(L)Cl] (4). The structures of compounds 1 (Pt) and 2 (Pd), as determined by X-ray diffraction, are reported and compared. Compound 2 can also be obtained from Na2[PdCl4] and HL in refluxing acetic acid, i.e., under the same conditions used for compound 1. Apparently, this is the first palladium pincer derivative of a 1,3-bis(pyridyl)benzene ligand synthesized by direct C-H activation.The neutral complexes 1-3 are catalysts of modest activity, but devoid of enantioselectivity in the Heck reaction between iodobenzene and methyl acrylate and in the aldol condensation of benzaldehyde with methyl isocyanoacetate.  相似文献   

17.
Ligands containing the 2-organochalcogenomethylpyridine motif with substituents in the 4- or 6-position of the pyridyl ring, R4,R6-pyCH2ER1 [R4 = R6 = H, ER1 = SMe (1), SeMe (2), SPh (6), SePh (7); R4 = Me, R6 = H, ER1 = SMe (3), SPh (8), SePh (9); R4 = H, R6 = Me, ER1 = SMe (4), SPh (10), SePh (11); R4 = H, R6 = Ph, ER1 = SMe (5), SPh (12), SePh (13)] are obtained on the reaction of R4,R6-pyMe with LiBun followed by R1EER1. On reaction with PdCl2(NCMe)2, the ligands with a 6-phenyl substituent form cyclopalladated species PdCl{6-(o-C6H4)pyCH2ER1-C,N,E} (5a, 12a, 13a) with the structure of 13a (ER1 = SePh) confirmed by X-ray crystallography; other ligands form complexes of stoichiometry PdCl2(R4,R6-pyCH2ER1). Complexes with R6 = H are monomeric with N,E-bidentate configurations, confirmed by structural analysis for 3a (R4 = Me, ER1 = SMe), 7a (R4 = H, ER1 = SePh) and 9a (R4 = Me, ER1 = SePh). Two of the 6-methyl substituted complexes examined by X-ray crystallography are oligomeric with trans-PdCl2(N,E) motifs and bridging ligands, trimeric [PdCl2(μ-6-MepyCH2SPh-N,S)]3 (10a) and dimeric [PdCl2(μ-6-MepyCH2SePh-N,Se)]2 (11a). This behaviour is attributed to avoidance of the Me···Cl interaction that would occur in the cis-bidentate configuration if the pyridyl plane had the same orientation with respect to the coordination plane as observed for 3a, 7a and 9a [dihedral angles 8.0(2)-16.8(2)°]. When examined as precatalysts for the Mizoroki-Heck reaction of n-butyl acrylate with aryl halides in N,N-dimethylacetamide at 120 °C, the complexes exhibit the anticipated trends in yield (ArI > ArBr > ArCl, higher yield for electron withdrawing substituents in 4-RC6H4Br and 4-RC6H4Cl). The most active precatalysts are PdCl2(R4-pyCH2SMe-N,S) (R = H (1a), Me (3a)); complexes of the selenium containing ligands exhibit very low activity. For closely related ligands, the changes SMe to SPh, 6-H to 6-Me, and 6-H to 6-Ph lead to lower activity, consistent with involvement of both the pyridyl and chalcogen donors in reactions involving aryl bromides. The precatalyst PdCl2(pyCH2SMe-N,S) (1a) exhibits higher activity for the reaction of aryl chlorides in Bun4NCl at 120 °C as a solvent under non-aqueous ionic liquid (NAIL) conditions.  相似文献   

18.
The structure of carbon monoxide dehydrogenase/acetyl-coenzyme A synthase (CODH/ACS), a central enzyme in the anaerobic metabolism of acetyl-coenzyme A (acetyl-CoA), has been solved to a resolution of 2.2A. The active-site metal cluster responsible for catalyzing acetyl C-C bond synthesis and cleavage, designated the A center, was identified as an Fe(4)S(4) iron sulfur cluster with one of its cysteine thiolates acting as a bridge to an adjacent binuclear metal site. Nickel was found at one position in the binuclear site and the other metal was indicated to be copper - a surprising result, implying a previously unrecognized role for copper. Details of the A center provided new insight into the unusual organometallic mechanism of acetyl C-C bond formation and cleavage, with substantial conformational changes indicated for binding of the large methylcorrinoid protein substrate, and a unique intramolecular channel acting to contain carbon monoxide within the protein and transfer it to the site needed for acetyl-CoA synthesis.  相似文献   

19.
The rhodium(I) complexes TpmsRh(CO)2 (1) and TpmsRh(cod) (2) of the tripodal nitrogen ligand tris(pyrazolyl)methanesulfonate, Tpms=[(pz)3CSO3], catalyze the hydroformylation of 1-hexene. Addition of phosphine has a negative effect on the activity. The hydroformylation activity reaches a maximum at about 60 °C. At temperatures above 80 °C hydrogenation becomes an important secondary reaction. When the catalysis is performed at 60 °C in acetone with 1 or 2 as catalyst precursor all of the rhodium is recovered in the form of the rhodium(III) bis(acyl) complex TpmsRh(CO)(COC6H13)2 (9). A similar behaviour is observed with rhodium(I) complexes bearing the tripodal oxygen ligand LOMe=[(cyclopentadienyl)tris(dimethylphosphito-P) cobalt O,O,O″]. In this case all of the rhodium is transformed into LOMeRh(CO)(COC6H13)2 (10). These hitherto unknown bis(acyl) rhodium(III) complexes show the same catalytic activity as the rhodium(I) starting compounds.  相似文献   

20.
The synthesis of new platinum bipy (bipy = 2,2′-bipyridyl) complexes containing phenoxide ligands is reported, together with kinetic studies of their oxidative addition reactions with MeI to produce phenoxo platinum(IV) complexes. Complexes of the form [(bipy)Pt(OC6H4-4-X)2] (X = OCH3, CH3, H, Br, Cl) are prepared by the reaction of the chloro complex [(bipy)PtCl2] with substituted phenols and KOH in a two phase system of water and chloroform in the presence of benzyl triphenylphosphonium chloride. Platinum(IV) complexes are formed by oxidative addition of MeI to the platinum(II) complexes obtained. The complexes are characterized by elemental analysis, UV-Vis, IR, mass spectrometry and 1H and 13C NMR spectroscopy.The reaction of methyl iodide with [(bipy)Pt(OC6H4-4-OMe)2] to give [(bipy)PtMe(I)(OC6H4-4-OMe)2] follows the rate law rate = k2[(bipy)Pt(OC6H4-4-OMe)2][MeI]. The values of k2 increase with increasing polarity of the solvent, suggesting a polar transition state for the reaction.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号