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 31 毫秒
1.
The intermediacy of metallocarbenes in decomposition reactions of iodonium ylides with [Rh2(OAc)4] was established by comparison with reactions of the corresponding diazo compounds. The sensitivity of the RhII-catalyzed intermolecular cyclopropane formation from substituted styrenes and bis(methoxycarbonyl)(phenyliodono)methanide ( 1a ) or dimethyl diazomalonate ( 1b ) is identical. The Hammett plot (with σ+) has a slope of ?0.47. Iodonium ylides and diazo compounds afford the same products in [Rh2(OAc)4]-catalyzed cyclopropane formations, cycloadditions, and intramolecular CH insertions, and exhibit the identical selectivity in intramolecular competitions for cyclopropane formation and insertion. The intramolecular CH insertion of the ylide 20c , when carried out in the presence of a chiral catalyst ([Rh2{(?)-(S)-ptpa}4]), results in formation of 21a having an ee of 67%, identical to the ee obtained with the diazo compound 20b .  相似文献   

2.
The Rh11-catalyzed carbenoid addition of diazoacetates to olefins was investigated with [Rh2{(4S)-phox}4] ( 1 ;phox = tetrakis[(4S)-tetrahydro-4-phenyloxazol-2-one]), [Rh2{(2S)-mepy}4] ( 2 ; mepy = tetrakis[methyl (2S)-tetrahydro-5-oxopyrrole-2-carboxylate]), and [Rh2(OAc)4] ( 3 ). While catalysis with 2 and 3 afford preferentially trans-cyclopropanecarboxylates, the cis-isomers are the major products with 1 . In general, the enantioselectivities achieved with 1 and 2 are comparable. Additions catalyzed by 1 are strongly sensitive to steric effects. Highly substituted olefins afford cyclopropanes in only poor yield. The preferential cis-selectivity observed in reactions catalyzed by 1 is attributed to dominant interactions between the ligand of the catalyst and the substituents of both olefin and diazoacetate, which overrule the steric interactions between olefin and diazoacetate in the transition state for carbene transfer.  相似文献   

3.
A series of Zn(II) and Cu(II) complexes were synthesized using unsymmetrical N,N′‐ diarylformamidine ligands, i.e. N‐(2‐methoxyphenyl)‐N′‐2,6‐dichorophenyl)‐formamidine ( L1 ), N‐(2‐methoxyphenyl)‐N′‐phenyl)‐formamidine ( L2 ), N‐(2‐methoxyphenyl)‐N′‐(2,6‐dimethylphenyl)‐formamidine ( L3 ) and N‐(2‐methoxyphenyl)‐N′‐(2,6‐diisopropylphenyl)‐formamidine ( L4 ). The complexes, [Zn2( L1 )2(OAc)4] ( 1) , [Zn2( L2 )2(OAc)4] ( 2 ), [Zn2( L3 )2(OAc)4] ( 3 ), [Zn2( L4 )2(OAc)4] ( 4 ), [Cu2( L1 )2(OAc)4] ( 5 ), [Cu2( L2 )2(OAc)4] ( 6 ), [Cu2( L3 )2(OAc)4] ( 7 ) and [Cu2( L4 )2(OAc)4] ( 8 ), were prepared via a mechanochemical method with excellent yields between 95 ‐ 98% by reacting the metal acetates and corresponding ligands. Structural studies showed that both complexes are dimeric with a paddlewheel core structure in which the separation between the two metal centres are 2.9898 (8) and 2.6653 (7) Å in complexes 3 and 7 , respectively. Complexes 1 – 8 were used in ring‐opening polymerization of ε‐caprolactone (ε‐CL) and rac‐lactide (rac‐LA). Zn(II) complexes were more active than Cu(II) complexes, with complex 1 bearing electron withdrawing chloro groups being the most active (kapp = 0.0803 h‐1). Low molecular weight poly‐(ε‐CL) and poly‐(rac‐LA) ranging from 1720 to 6042 g mol‐1, with broad molecular weight distribution (PDIs, 1.78 – 1.87) were obtained. Complex 2 gave reaction orders of 0.56 and 1.52 with respect to ε‐CL and rac‐LA, respectively.  相似文献   

4.
Radiation reduction of binuclear [Rh2(OAc)2(phen)2(H2O)2](OAc)2, [Rh2(OAc)(tpy)2Cl2]Cl·2H2O and [Rh2Cl2(HCOO)2(bpy)2]·4H2O complexes in aqueous-methanol solution have been studied. The reduction yields as equal to ca. 6 equiv/100eV and the rate constants of reactions: complex+e solv as equal to 2.9·1010, 3.2·1010 and 3.7·1010 M−1·s−1, respectively, have been determined. On the basis of electronic spectra it has been shown that Rh(II) compounds were reduced giving several Rh(I) complexes being in equilibrium. The mechanism of the processes has been discussed.  相似文献   

5.
Abstract

The reaction of antitumor active dirhodium(II) tetraacetate, [Rh2(AcO)4], with S-methyl-L-cysteine (HSMC) was studied at the pH of mixing (=4.8) in aqueous media at various temperatures under aerobic conditions. The results from UV–vis spectroscopy and electrospray ionization mass spectrometry (ESI–MS) showed that HSMC initially coordinates via its sulfur atom to the axial positions of the paddlewheel framework of the dirhodium(II) complex, and was confirmed by the crystal structure of [Rh2(AcO)4(HSMC)2]. After some time (48?h at 25?°C), or at elevated temperature (40?°C), Rh-SMC chelate formation causes breakdown of the paddlewheel structure, generating the mononuclear Rh(III) complexes [Rh(SMC)2]+, [Rh(AcO)(SMC)2] and [Rh(SMC)3], as indicated by ESI–MS. These aerobic reaction products of [Rh2(AcO)4] with HSMC have been compared with those of the two proteinogenic sulfur-containing amino acids methionine and cysteine. Comparison shows that the (S,N)-chelate ring size influences the stability of the [Rh2(AcO)4] paddlewheel cage structure and its RhII–RhII bond, when an amino acid with a thioether group coordinates to dirhodium(II) tetraacetate.  相似文献   

6.
The synthesis of the reactive PN(CH) ligand 2‐di(tert‐butylphosphanomethyl)‐6‐phenylpyridine ( 1H ) and its versatile coordination to a RhI center is described. Facile C?H activation occurs in the presence of a (internal) base, thus resulting in the new cyclometalated complex [RhI(CO)(κ3P,N,C‐ 1 )] ( 3 ), which has been structurally characterized. The resulting tridentate ligand framework was experimentally and computationally shown to display dual‐site proton‐responsive reactivity, including reversible cyclometalation. This feature was probed by selective H/D exchange with [D1]formic acid. The addition of HBF4 to 3 leads to rapid net protonolysis of the Rh?C bond to produce [RhI(CO)(κ3P,N,(C?H)‐ 1 )] ( 4 ). This species features a rare aryl C?H agostic interaction in the solid state, as shown by X‐ray diffraction studies. The nature of this interaction was also studied computationally. Reaction of 3 with methyl iodide results in rapid and selective ortho‐methylation of the phenyl ring, thus generating [RhI(CO)(κ2P,N‐ 1Me )] ( 5 ). Variable‐temperature NMR spectroscopy indicates the involvement of a RhIII intermediate through formal oxidative addition to give trans‐[RhIII(CH3)(CO)(I)(κ3P,N,C‐ 1 )] prior to C?C reductive elimination. The RhIIItrans‐diiodide complex [RhI(CO)(I)23P,N,C‐ 1 )] ( 6 ) has been structurally characterized as a model compound for this elusive intermediate.  相似文献   

7.
Transformation of N‐alkylated anilines to N‐aryloxamates was studied using ethyl 2‐diazoacetoacetate as an alkylating agent and dirhodium tetraacetate (Rh2(OAc)4) as the catalyst. The general applicability of the reaction as a synthetic method for N‐aryloxamates was studied with a number of substituted N‐alkylated anilines. The results revealed that the oxamate was formed by a radical reaction with molecular O2 and Rh2(OAc)4 as initiator.  相似文献   

8.
Reduction of [Rh2(-OAc)2(bpy)2(H2O)2](OAc)2 and [Rh2Cl2(-OAc)2(bpy)2] · 3H2O complexes with ethanol and [Cr2(OAc)4(H2O)2] has been investigated using e.p.r. and u.v.–vis. spectra. The results indicate that stable complexes containing the [Rh2 3+] entity are not formed. The X-ray structure of [Rh2Cl2(-OAc)2(bpy)2] · 3H2O has been determined. Coordination around the Rh atom is in the form of a distorted octahedron. The complex shows an almost ideal eclipsed conformation. The equatorial coordination sites are occupied by bridging carboxylato ligands and 2,2-bipyridine and axial positions by the Cl ligand and the rhodium atom. The Rh–Rh distance is 2.574 Å.  相似文献   

9.
The Cu‐catalyzed intramolecular CH insertion of phenyliodonium ylide 1b was investigated at 0° in the presence of several chiral ligands. Enantioselectivities varied in the range 38–72%, and were higher than those resulting from reaction of the diazo compound 1c at 65°. The intramolecular insertion of the enantiomerically pure methyl diazoacetate (R)‐ 20 and of the corresponding phenyliodonium ylide (R)‐ 21 proceeded to (R)‐ 23 with retention of configuration with [Cu(hfa)2] (hfa=hexafluoroacetylacetone=1,1,1,5,5,5‐hexafluoropentane‐2,4‐dione) and [Rh2(OAc)4]. These results are consistent with a carbenoid mechanism for the Cu‐catalyzed insertion with phenyliodonium ylides. However, the insertion of the perfluorosulfonated phenyliodonium ylide (R)‐ 29 afforded with [Cu(hfa)2] as well as with [Rh2(OAc)4] the cyclopentanone derivative 30 as a cis/trans mixture with only 56–67% enantiomeric excess.  相似文献   

10.
In diaqua­tetra‐μ‐acetamidato‐κ4N:O4O:N‐di­rhodium(II,III) hexa­fluoro­phosphate, [Rh2(C2H4NO)4(H2O)2]PF6, and diaqua­tetra‐μ‐acetamidato‐κ4N:O4O:N‐di­rho­dium(II,III)hexa­fluoro­phosphate dihydrate, [Rh2(C2H4NO)4(H2O)2]PF6·2H2O, the cations and anions lie on inversion centers. Diaqua­tetra‐μ‐propionamidato‐κ4N:O4O:N‐dirhodium(II,III) hexa­fluoro­phosphate dihydrate, [Rh2(C3H6NO)4(H2O)2]PF6·2H2O, and diaqua­tetra‐μ‐butyramidato‐κ4N:O4O:N‐dirhodium(II,III) hexa­fluoro­phosphate, [Rh2(C4H8NO)4(H2O)2]PF6, crystallize with two crystallographically independent complexes that lie on inversion centers. In all of the structures, the dirhodium units are hydrogen bonded to one another. The hydrogen‐bonded networks vary with the alkyl substituents.  相似文献   

11.
Realizing a need for labeled dirhodium tetraacetate [Rh2/OAc/4] for binding studies by the equilibrium dialysis procedure, we have established a method for the preparation and purification of [3H]Rh2/OAc/4. This is obtained with a tritium label of a sufficiently high specific activity in the methyl group of acetate residues. The method consists of replacement of the existing acetate bridge by [3H]OAc. A typical preparation involves refluxing Rh2/OAc/4 with [3H]NaOAc in dry ethanol at 80°C for 6 h. The reaction time is critical for obtaining the product with a high specific activity. A simple purification of reaction products yields the compound of interest with a high specific activity, i. e., 315±10 Ci per mol of Rh2/OAc/4 and in satisfactory yield /87% of the starting material., This labeled material can be synthesized with greatly increased specific activity /of tritium in acetate groups/ by employing [3H]NaOAc without prior dilution. The labeled Rh2/OAc/4 is stable and yields a characteristic product with adenylic acid.  相似文献   

12.
Summary Reactions of glyoxal bis(morpholineN-thiohydrazone), H2gbmth, with NiCl2·6H2O, Ni(OAc)2·4H2O, Ni(acac)2· H2O, CuCl2·2H2O, Cu(OAc)2·H2O, Cu(acac)2, CoCl2· 6H2O, Co(OAc)2·4H2O and Co(acac)2·2H2O yield complexes of the type [M(gbmth)], [M=NiII, CuII or CoII]. Diacetyl reacts with morpholineN-thiohydrazide in the presence of nickel salts to yield [NiII(dbmth)], [NiII(dmth)(OAc)]H2O and [NiII(Hdmth)(NH3)Cl2] involving N2S2 and NSO donor ligands. Copper and cobalt complexes of N2S2 and NSO donor ligands with compositions [CuII(dbmth)], [CoII(dbmth)]·4H2O and [CoII(H2dbmth)]Cl2, have been isolated. The compounds have been characterised by elemental analyses, magnetic moments, molar conductance values and spectroscopic (electronic and infrared) data.  相似文献   

13.
N-Carboethoxy-4-chlorobenzene thioamide (Hcct or HL) and N-carboethoxy-4-bromobenzene thioamide (Hcbt or HL) react with bivalent (Ni, Co, Cu, Ru, Pd and Pt), trivalent (Ru and Rh) and tetravalent (Pt) transition metal ions to give [MII(L)2], [RuIII(L)3], [RhIII(L)(HL)Cl2] and [Pt(L)2Cl2] complexes, respectively. In the presence of pyridine, CoII and NiII salts react with the ligands (HL) to give [MII(L)2Py] (M = Co and Ni) complexes. Soft metal ions abstract sulphur from the ligands to yield the corresponding sulphide, together with oxygenated forms of the ligands. All the metal complexes have been characterised by chemical analyses, conductivity, spectroscopic and magnetic measurements.  相似文献   

14.
Synthesis and Dynamic Behaviour of [Rh2(μ-H)3H2(PiPr3)4]+. Contributions to the Reactivity of the Tetrahydridodirhodium Complex [Rh2H4(PiPr3)4] An improved synthesis of [Rh2H4(PiPr3)4] ( 2 ) from [Rh(η3-C3H5)(PiPr3)2] ( 1 ) or [Rh(η3-CH2C6H5)(PiPr3)2] ( 3 ) and H2 is described. Compound 2 reacts with CO or CH3OH to give trans-[RhH(CO)(PiPr3)2] ( 4 ) and with ethene/acetone to yield a mixture of 4 and trans-[RhCH3(CO)(PiPr3)2] ( 5 ). The carbonyl(methyl) complex 5 has also been prepared from trans-[RhCl(CO)(PiPr3)2] ( 6 ) and CH3MgI. Whereas the reaction of 2 with two parts of CF3CO2H leads to [RhH22-O2CCF3) · (PiPr3)2] ( 8 ), treatment of 2 with one equivalent of CF3CO2H in presence of NH4PF6 gives the dinuclear compound [Rh2H5(PiPr3)4]PF6 ( 9a ). The reactions of 2 with HBF4 and [NO]BF4 afford the complexes [Rh2H5(PiPr3)4]BF4 ( 9b ) and trans-[RhF(NO)(PiPr3)2]BF4 ( 11 ), respectively. In solution, the cation [Rh2(μ-H)3H2(PiPr3)4]+ of the compounds 9a and 9b undergoes an intramolecular rearrangement in which the bridging hydrido and the phosphane ligands are involved.  相似文献   

15.
The synthesis of [Ir2Rh2(CO)12] ( 1 ) by the literature method gives a mixture 1 /[IrRh3(CO)12] which cannot be separated using chromatography. The reaction of [Ir(CO)4]? with 1 mol-equiv. of [Rh(CO)2(THF)2]+ in THF gives pure 1 in 61% yield. Crystals of 1 are highly disordered, unlike those of its derivative [Ir2Rh2(CO)52-CO)3(norbornadiene)2] which were analysed using X-ray diffraction. The ground-state geometry of 1 in solution has three edge-bridging CO's on the basal IrRh2 face of the metal tetrahedron. Time averaging of CO's takes place above 230 K. The CO site exchange of lowest activation energy is due to one synchronous change of basal face, as shown by 2D- and VT-13C-NMR. Substitution of CO by X? in 1 takes place at a Rh-atom giving [Ir2Rh2(CO)82-CO)3X]? (X = Br, I). Substitution by bidentate ligands gives [Ir2Rh2(CO)72-CO)34-L)] (L = norbornadiene, cycloocta-1,5-diene) where the ligand L is chelating a Rh-atom of the basal IrRh2 face. Carbonyl substitution by tridentate ligands gives [Ir2Rh2(CO)62-CO)33-L)] (L = 1,3,5-trithiane, tripod) with L capping the triangular basal face of the metal tetrahedron. Carbonyl scrambling is also observed in these substituted derivatives of 1 and is mainly due to the rotation of three terminal CO's about a local C3 axis on the apical Ir-atom.  相似文献   

16.
The synthesis, reduction, optical and e.p.r. spectral properties of a series of new binuclear copper(II) complexes, containing bridging moieties (OH, MeCO2 , NO2 , and N3 ), with new proline-based binuclear pentadentate Mannich base ligands is described. The ligands are: 2,6-bis[(prolin-1-yl)methyl]4-bromophenol [H3L1], 2,6-bis[(prolin-1-yl)methyl]4-t-butylphenol [H3L2] and 2,6-bis[(prolin-1-yl)methyl]4-methoxyphenol [H3L3]. The exogenous bridging complexes thus prepared were hydroxo: [Cu2L1(OH)(H2O)2] · H2O (1a), [Cu2L2(OH)(H2O)2] · H2O (1b), [Cu2L3(OH)(H2O)2] · H2O (1c), acetato [Cu2L1(OAc)] · H2O (2a), [Cu2L2(OAc)] · H2O (2b), [Cu2L3(OAc)] · H2O (2c), nitrito [Cu2L1(NO2)(H2O)2] · H2O (3a), [Cu2L2(NO2)(H2O)2] · H2O (3b), [Cu2L3(NO2)(H2O)2] · H2O (3c) and azido [Cu2L1(N3)(H2O)2] · H2O (4a), [Cu2L2(N3)(H2O)2] · H2O (4b) and [Cu2L3(N3)(H2O)2] · H2O (4c). The complexes were characterized by elemental analysis and by spectroscopy. They exhibit resolved copper hyperfine e.p.r. spectra at room temperature, indicating the presence of weak antiferromagnetic coupling between the copper atoms. The strength of the antiferromagnetic coupling lies in the order: NO2 N3 OH OAc. Cyclic voltammetry revealed the presence of two redox couples CuIICuII CuIICuI CuICuI. The conproportionality constant K con for the mixed valent CuIICuI species for all the complexes have been determined electrochemically.  相似文献   

17.
6‐(Diazomethyl)‐1,3‐bis(methoxymethyl)uracil ( 5 ) was prepared from the known aldehyde 3 by hydrazone formation and oxidation. Thermolysis of 5 and deprotection gave the pyrazolo[4,3‐d]pyrimidine‐5,7‐diones 7a and 7b . Rh2(OAc)4 catalyzed the transformation of 5 into to a 2 : 1 (Z)/(E) mixture of 1,2‐diuracilylethenes 9 (67%). Heating (Z)‐ 9 in 12n HCl at 95° led to electrocyclisation, oxidation, and deprotection to afford 73% of the pyrimido[5,4‐f]quinazolinetetraone 12 . The Rh2(OAc)4‐catalyzed reaction of 5 with 3,4‐dihydro‐2H‐pyran and 2,3‐dihydrofuran gave endo/exo‐mixtures of the 2‐oxabicyclo[4.1.0]heptane 13 (78%) and the 2‐oxabicyclo[3.1.0]hexane 15 (86%), Their treatment with AlCl3 or Me2AlCl promoted a vinylcyclopropane–cyclopentene rearrangement, leading to the pyrano‐ and furanocyclopenta[1,2‐d]pyrimidinediones 14 (88%) and 16 (51%), respectively. Similarly, the addition product of 5 to 2‐methoxypropene was transformed into the 5‐methylcyclopenta‐pyrimidinedione 18 (55%). The Rh2(OAc)4‐catalyzed reaction of 5 with thiophene gave the exo‐configured 2‐thiabicyclo[3.1.0]hexane 19 (69%). The analoguous reaction with furan led to 8‐oxabicyclo[3.2.1]oct‐2‐ene 20 (73%), and the reaction with (E)‐2‐styrylfuran yielded a diastereoisomeric mixture of hepta‐1,4,6‐trien‐3‐ones 21 (75%) that was transformed into the (1E,4E,6E)‐configured hepta‐1,4,6‐trien‐3‐one 21 (60%) at ambient temperature.  相似文献   

18.
A highly regio- and stereoselective method has been developed for the synthesis of spiro[furo[3,4-c]chromene-1,3′-indoline]-2′,4(9bH)-dione derivatives via a three component reaction of cyclic diazoamide and aldehyde with methyl 2-oxo-2H-chromene-3-carboxylate, 3-acetyl-2H-chromen-2-one and 3-benzoyl-2H-chromen-2-one using 3 mol % of Rh2(OAc)4. Similarly, acyclic diazoesters also undergo smooth coupling with carbonyl compounds and 3-substituted coumarin in the presence of 1 mol % of Rh2(OAc)4 to afford a novel series of tetrahydro-1H-furo[3,4-c]chromene-1-carboxylates in 78–88% yield with high regio- and stereo-selectivity for the first time.  相似文献   

19.
The reaction of the [Ni6(CO)12]2− dianion with [Rh(COD)Cl]2 (COD = cyclooctadiene) in acetone affords a mixture of bimetallic Ni–Rh clusters, mainly consisting of the new [Ni7Rh3(CO)18]3− and [Ni8Rh(CO)18]3− trianions. A study of the reactivity of [Ni7Rh3(CO)18]3− led to isolation of the new [Ni3Rh3(CO)13]3− and [NiRh8(CO)19]2− anions. All these new bimetallic Ni–Rh carbonyl clusters have been isolated in the solid state as tetrasubstituted ammonium salts and have been characterised by elemental analysis, X-ray diffraction studies, ESI-MS and electrochemistry. The unit cell of the [NEt4]3[Ni7Rh3(CO)18] salt contains two orientationally-disordered ν2-tetrahedral [Ni7Rh3(CO)18]3− trianions with occupancy factors of 0.75 and 0.25. Besides, their inner Ni3Rh3 octahedral moieties show two cis sites purely occupied by Rh atoms, two trans sites purely occupied by Ni atoms and the remaining two cis sites are disordered Ni and Rh sites with respective occupancy fraction of 0.5. At difference from the parent [Ni7Rh3(CO)18]3−, the octahedral [Ni3Rh3(CO)13]3− displays an ordered distribution of Ni and Rh atoms in two staggered triangles. The [NiRh8(CO)19]2− dianion adopts an isomeric metal frame with respect to that of the [PtRh8(CO)19]2− congener. As a fallout of this work, new high-yield synthesis of the known [Ni6Rh3(CO)17]3− and [Ni6Rh5(CO)21]3−, as well as other currently-investigated bimetallic Ni–Rh clusters have been obtained.  相似文献   

20.
The reaction of PP(NO2) with M4(CO)12 (M = Co, Rh) gives the nitrido clusters [M6N(CO)15]? in 13 and 21% yields, respectively. A high yield synthesis (77%) of [Rh6N(CO)15)]? directly from Rh6(CO)16 and PPN(NO2) is also presented. PPN(NO2) reacts with Ir4(CO)12 to give the new isocyanato cluster, [Ir4(NCO)(CO)11]? in 34% yield, while the direct synthesis of this isocyanate product occurs in 77% yield from PPN(N3) and Ir4(CO)12. Modifications of published procedures for the preparation of [N(C2H5)4]2 [Ir6(CO)15] and Ir6(CO)16 are reported that allow shorter reaction times and give higher yields. The reaction of Ir6(CO)16 with one equivalent of PPN(NO2) generates a new cluster, PPN[Ir6(CO)15(NO)], in 57% yield which is proposed to contain a bent nitrosyl ligand. An additional equivalent of PPN(NO2) gives (PPN)2[Ir6(CO)15] in 84% yield with the evolution of N2O as well as CO2.  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