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 78 毫秒
1.
From Li+ well-solvating solvents or complex ligands such as THF, [12]crown-4, amines etc., lithium cuprates R2CuLi(*LiX) crystallise in a solvent-separated ion pair (SSIP) structural type (e.g. 10). In contrast, solvents with little donor qualities for Li+ such as diethyl ether or dimethyl sulfide lead to solid-state structures of the contact ion pair (CIP) type (e.g. 11). 1H,6Li HOESY NMR investigations in solutions of R2CuLi(*LiX) (15, 16) are in agreement with these findings: in THF the SSIP 18 is strongly favoured in the equilibrium with the CIP 17, and in diethyl ether one observes essentially only the CIP 17. Salts LiX (X=CN, Cl, Br, I, SPh) have only a minor effect on the ion pair equilibrium. These structural investigations correspond perfectly with Bertz's logarithmic reactivity profiles (LRPs) of reactions of R2CuLi with enones in diethyl ether and THF: the faster reaction in diethyl ether is due to the predominance of the CIP 17 in this solvent, which is the reacting species; in THF only little CIP 17 is present in a fast equilibrium with the SSIP 18. A kinetic analysis of the LRPs quantifies these findings. Recent quantum-chemical studies are also in agreement with the CIP 17 being the reacting species. Thus a uniform picture of structure and reactivity of lithium cuprates emerges.  相似文献   

2.
The comprehension of factors influencing the reactivity of organocuprates is still far from enabling a rational control of their reactions. Especially the degree of aggregation and structures of organocuprates are the focus of discussion about the factors affecting their reactivity. Therefore, this study combines kinetic measurements and NMR investigations to elucidate the influence of disaggregation via addition of tetrahydrofuran (THF) on the reactivity and aggregate structure of Gilman cuprates. As model systems, Me(2)CuLi.LiI (1.LiI) and Me(2)CuLi.LiCN (1.LiCN) in diethyl ether (DEE) were chosen; as model reaction, the 1,4-addition to 4,4-dimethylcyclohex-2-enone. The kinetic data show for 1.LiI a pronounced acceleration effect upon addition of distinct amounts of THF, whereas the reactivity of 1.LiCN continuously decreases with the addition of THF. Series of NMR diffusion measurements as well as (1)H-(7)Li heteronuclear Overhauser effect (HOE), and (1)H-(1)H nuclear Overhauser effect (NOE) spectra show different structural influences of THF on 1.LiI and 1.LiCN. For 1.LiI, small salt units are separated from the cuprate aggregate by THF. In contrast to this, THF disaggregates the oligomeric structures of 1.LiCN, while the core structures remain intact with salt attached. Thus, the reactivity of 1.LiI seems to be fine-tuned through distinct amounts of salt or THF, whereas the decreasing reactivity of 1.LiCN correlates with the disaggregation of oligomers via THF. Thus, for synthetic chemists with reactivity problems in specific reactions of iododialkylcuprates, the addition of small amounts of THF might be useful to enhance the reactivity. In addition to these structure-reactivity studies, the CN(-) group is shown to be directly attached to the cuprate moiety via a combination of (1)H-(13)C HOE- and (1)H-(1)H NOEs. This represents the first direct experimental evidence in solution for the position of the CN(-) group relative to the cuprate moiety in cyano-Gilman cuprates.  相似文献   

3.
The molecular sizes of higher aggregates of dimethylcuprates (Me(2)CuLi (1), 1.LiI, and 1.LiCN) and bis[(trimethylsilyl)methyl]cuprates ((Me(3)SiCH(2))(2)CuLi (2), 2.LiI, and 2.LiCN) in diethyl ether (Et(2)O) were determined by pulsed field gradient (PFG) NMR diffusion measurements. The obtained diffusion coefficients show molecular sizes larger than those of dimers for all systems. In these higher aggregates, steric hindrance and dilution reduce aggregation, whereas LiCN increases it. The molecular sizes were first determined by a spherical model-free approach and then refined by structure models of higher aggregates. These models were built by a combination of diffusion results, known NMR studies, and crystal structures. Thus, polymeric chains with homodimeric cores connected by solvent (salt-free case) or solvent and salt (salt-containing case) were proposed. These models were confirmed by a solvation analysis, whereby the number of solvent molecules attached to the aggregates was determined by a weighted average study. On the basis of these structure models, the number of repetition units (length index) was determined to be between 1.3 and 5.2, with the general trends in aggregation independent of the structure model used. A combined analysis of the determined length indices and known relative reactivities led for the first time to a correlation between higher aggregation and reactivity of dimethylcuprates in the addition reaction with enones: aggregates higher than dimers reduce the reactivity. Consequently, despite their consistent homodimeric core structures, for the first time the remaining reactivity differences between iodo- and cyanodimethylcuprates in Et(2)O are explained by the difference in their aggregation.  相似文献   

4.
H-Li distances and (1)H-(1)H dipolar interactions in Me(2)CuLiLiCN and Me(2)CuLi in diethyl ether (Et(2)O), obtained by NMR spectroscopy, were used to gain structural information about the contact ion pair of the salt-containing organocuprate Me(2)CuLiLiCN in this solvent. The H-Li distances of Me(2)CuLiLiCN and Me(2)CuLi in Et(2)O, resulting from the initial buildup rates in conjunction with the motional correlation times, are almost identical, indicating a similar homodimeric core structure [Me(2)CuLi](2) for both samples. However, the H-Li distances obtained for Me(2)CuLiLiCN do not rigorously exclude a heterodimeric structure [Me(2)CuLiLiCN] as proposed by ab initio calculations. Therefore, (1)H-(1)H dipolar interactions were investigated by SYM-BREAK-NOE/ROE-HSQC experiments, which allow for the observation of NOEs between equivalent protons. Since these experiments showed similar (1)H-(1)H dipolar interactions of Me(2)CuLiLiCN and Me(2)CuLi, we propose that for Me(2)CuLiLiCN a homodimeric core structure [Me(2)CuLi](2) indeed is predominant in Et(2)O.  相似文献   

5.
Rapid-injection is a very useful technique for the preparation of temperature-sensitive and air-sensitive compounds in the cold, nitrogen-filled probe of an NMR spectrometer. We have used this method to prepare solutions of pi-complexes from 2-cyclohexenone and prototypical cuprates Me2CuLi.LiI and Me2CuLi.LiCN, and we have assigned structures on the basis of 1H and 13C NMR. In each case two pi-complexes were observed, and in the former, their rates of formation were measured by rapid-injection 1H NMR and EXSY spectroscopy. These results provide insights into the normal and anomalous conjugate addition reactions of organocuprates.  相似文献   

6.
Reaction of in situ generated lithium phosphides with 0.5 eq. Cu(I) is employed as a means of targeting lithium phosphidocuprates of either Gilman- or Lipshutz-type formulation--e.g., (R(2)P)(2)CuLi·n(LiX) (n = 0, 1). For R = Ph, X = CN in toluene followed by thf or R = Ph, X = I in thf/toluene an unexpected product results. [(Ph(2)P)(6)Cu(4)][Li·4thf](2)1 reveals an ion-separated structure in the solid state, with solvated lithium cations countering the charge on an adamantyl dianion [(Ph(2)P)(6)Cu(4)](2-). Deployment of R = Ph, X = CN in thf affords a novel network based on the dimer of Ph(2)PCu(CN)Li·2thf 2 with trianions based on 6-membered (PCu)(3) rings acting as nodes in the supramolecular array and solvated alkali metal counter-ions completing the linkers. Cy(2)PLi (Cy = cyclohexyl) has been reacted with CuCN in thf/toluene to yield Gilman-type lithium bis(phosphido)cuprate (Cy(2)P)(2)CuLi·2thf 3 by the exclusion of in situ generated LiCN. A polymer is noted in the solid state.  相似文献   

7.
Exposure of easily reduced aromatic bis(enones) 1a-1e to the methyl Gilman reagent Me(2)CuLi.LiI at 0 degrees C in tetrahydrofuran solvent provides the products of tandem conjugate addition-Michael cyclization, 2a-2e, along with the products of [2 + 2] cycloaddition, 3a-3e. Complete partitioning of the Gilman alkylation and [2 + 2] cycloaddition pathways may be achieved by adjusting the loading of the Gilman reagent, the rate of addition of the Gilman reagent, and the concentration of the reaction mixture. The Gilman alkylation manifold is favored by the rapid addition of excess Gilman reagent at higher substrate concentrations, while the [2 + 2] cycloaddition manifold is favored by slow addition of the same Gilman reagent at lower concentrations and loadings. Notably, [2 + 2] cycloaddition to form 3a-3e is catalytic in Gilman reagent. Kinetic data reveal that the ratio of 2a and 3a changes such that the cycloaddition pathway becomes dominant upon increased consumption of Gilman reagent. These data suggest a concentration-dependent speciation of the Gilman reagent and differential reactivity of the aggregates present at higher and lower concentrations. While the species present at higher concentration induce Gilman alkylation en route to products 2a-2e, the species present at lower concentration provide products of catalytic [2 + 2] cycloaddition, 3a-3e. Moreover, upon electrochemical reduction of the bis(enones) 1a-1e, or chemically induced single-electron transfer from arene anion radicals, the very same [2 + 2] cycloadducts 3a-3e are formed. The collective data suggest that [2 + 2] cycloadducts 3a-3e arising under Gilman conditions may be products of anion radical chain cyclobutanation that derive via electron transfer (ET) from the Me(2)CuLi.LiI aggregate(s) present at low concentration. These observations provide a link between the Gilman alkylation reaction and related ET chemistry and suggest these reaction paths are mechanistically distinct. This analysis is made possible by the recent observation that easily reduced bis(enones) are subject to intramolecular [2 + 2] cycloaddition upon cathodic reduction or chemically induced ET from arene anion radicals, and is herewith showcased as a novel method of testing for the intermediacy of enone anion radicals.  相似文献   

8.
Since the proposal of the dummy ligand concept by Corey, it has been widely accepted that the ligand transfer selectivity of a mixed organocuprate (Me(X)CuLi) depends on the Cu-X bond strength. The present B3LYP density functional studies on the Me(2)(X)Cu(III).OMe(2), pi-allyl Cu(III), and Me(X)Cu(III)LiCl.LiCl reacting with acrolein showed that the ligand transfer selectivity of the conjugate addition depends on two factors, thermodynamic stability (X = tert-butyl, ethynyl, cyano, and thiomethyl groups) and kinetic reactivity ((trimethylsilyl)methyl and vinyl groups) of the Cu(III) intermediate formed by complexation of the cuprate and the alpha,beta-unsaturated carbonyl compound. For the typical dummy ligands (X = alkynyl, cyano, and heteroatom ligands), the trans effect and the strong Li-X affinity are the reasons why these ligands stay on the copper atom. In contrast, for the (trimethylsilyl)methyl and vinyl groups, the selectivity depends on the kinetics of reductive elimination of the Cu(III) intermediate. The (trimethylsilyl)methyl transfer is retarded by repulsive four-electron interaction between the lone pair Cu 3d(xy)() orbital and the C-Si sigma-orbital.  相似文献   

9.
Systematic evaluation of method and basis set on the structure and energetics of organocuprate(I) and organocopper(III) species has been carried out. Various structures of organocuprate(I) and organocuprate(III) complexes were optimized with the HF, MP2, and B3LYP methods, and compared with the structures determined by X-ray crystallography (i.e., Me(2)Cu(I)(-), (CF(3))(4)Cu(III)(-)). Both the MP2 and B3LYP methods reasonably reproduce the X-ray structures while the HF method does not. Using larger basis set and incorporating the relativistic effects for Cu afford the best results. In the studies on the energetics of a Libond;Cu cluster model (Me(2)CuLi. LiCl) and Me(3)Cu model with the MP2, MP3, MP4DQ, MP4SDQ, CCSD(T), and B3LYP methods, the B3LYP method gives energetics similar to those obtained with the CCSD(T) method with much less cost, and hence, is judged to be the best practical method. The studies have shown that B3LYP method with the basis set incorporating the relativistic effective core potentials for Cu and the 6-31G* basis set for the rest is the theoretical method that is the most cost-effective for the studies of the structure and energetics of organocuprate(I) and organocopper(III) species.  相似文献   

10.
We have used a combination of electrospray ionization mass spectrometry and electrical conductivity measurements to analyze solutions of the Gilman cuprates LiCuR2?LiX, with R=Ph, Bu and X=Cl, Br, I, in tetrahydrofuran and have compared our findings with previous results on cyanocuprates LiCuR2?LiCN. Among the various polynuclear organocuprate ions observed, Li2Cu3Ph6?, LiCu4Ph6?, and Cu5Ph6? are of particular interest because aggregates of the same composition are known from X‐ray crystal structures. Control experiments have indicated that the polynuclear organocuprate anions detected in solution are indeed identical to those formed in the solid state. As abundant ions of the type Li2Cu3R6? are found in solutions of Gilman cuprates and cyanocuprates alike, their possible involvement in organocuprate reactions should be considered. For comparison, we have also included solutions of LiCu(R)I, LiCuX2?LiX, LiCuX2, and CuCN?2 LiX in the present study.  相似文献   

11.
Tandem cuprate addition-Dieckmann condensation is featured in the construction of the polyketide metabolite tetracenomycin A(2). Thus, cyclization substrate 11 was treated with Gilman cuprate Me(2)CuLi to afford anthracene 12. The phenolic acetate protecting group of 12ensured its chemoselective oxidation to reveal terminal quinone 13, which intercepts the previously reported synthesis of tetracenomycin A(2).  相似文献   

12.
The origin of the contrasting regioselectivities in allylic substitution found for a heterocuprate MeCu(CN)Li and a homocuprate Me2CuLi was studied using density functional calculations. The gamma-selectivity of MeCu(CN)Li is determined at the oxidative addition stage of the reaction, where the different degree of trans effect of the Me and the CN groups dictates the relative orientation of the methyl group and the leaving acetate group. As the result, the transition state where the acetate group leaves trans to the Me group on the copper atom is favored, and the gamma-selectivity results. The homocuprate Me2CuLi is symmetrical by nature and does not show such regioselectivity.  相似文献   

13.
NMR spectroscopic studies have been performed on the mixed complexes formed by the lithium salt of acetonitrile (LiCH(2)CN) and the chiral lithium amides Li-(S)-N-(2-methoxybenzyl)-1-amino-1-phenyl-2-ethoxyethane (Li-1) and Li-(S)-N-isopropyl-2-amino-1-phenyl-3-methoxypropane (Li-2) in diethyl ether and tetrahydrofuran solvent. In diethyl ether Li-1 and LiCH(2)CN form a mixed dimeric (1:1) complex, while Li-2 and LiCH(2)CN form a mixed trimeric (2:1) complex. The dimer undergoes fast exchange between ketenimine and bridged structures. Both (1)J((15)N,(6)Li) and (1)J((13)C,(6)Li) couplings were observed for the respectively isotopically labeled compounds. In the trimeric complex the CH(2)CN anion also undergoes fast degenerate exchange between ketenimine and bridged structures, and the complex appears C(2)-symmetric on the NMR spectroscopy time scale. Both the dimer and trimer complexes have the bridged acetonitrile anion in common, as indicated by the highly shielded alpha-carbon (13)C NMR shifts (delta -6.1 and -7.4, respectively). In tetrahydrofuran only N-metalated mixed LiCH(2)CN dimers were observed for both Li-1 and Li-2 with the less shielded (13)C NMR shifts of delta -2.5 and -2.2 for the alpha-carbon of LiCH(2)CN of the complexes.  相似文献   

14.
The anti stereochemistry observed in organocuprate SN2′ displacements can be rationalized as a stereoelectronic effect arising from “bidentate” binding involving a d orbital of nueleophilic copper and π* and σ* orbitals of the substrate. The extension of this idea to other reactions of organocuprates including additions to acetylenes and enones is discussed.  相似文献   

15.
Light‐induced transformation of atropisomeric and achiral enones displays divergent reactivity. Photocyclization leading to 3,4‐dihydroquinolin‐2‐one was observed with achiral enone carboxamide, whereas the atropisomeric enone carboxamides underwent hydrogen abstraction leading to spiro‐β‐lactams. Detailed photochemical, photophysical, and theoretical investigations have provided insight into this divergent reactivity and selectivity.  相似文献   

16.
The addition of stannylcuprate reagents such as (Bu3Sn)(PhS)CuLi to alkynones has been found to proceed in high yield and with excellent stereoselectivity for the Z isomer of the product (>95%). The behavior of the stannylcuprates is thus very different from that of their "carbocuprate" counterparts such as Me2CuLi or Me2Cu(CN)Li2 which are nonstereoselective. Furthermore, in contrast to the reactions of (R3Sn)(PhS)CuLi with the corresponding alkynoates, the presence of a proton source in the reaction medium has no effect on the stereoselectivity of the reaction of alkynones.  相似文献   

17.
Enantiomerically enriched bicyclo[m.1.0]alk-3-en-2-ones possessing 8-, 12-, and 15-membered rings were prepared and subjected to additions of nucleophiles. 1,2-Additions of n-butyllithium were highly diastereoselective for all cyclopropyl enones examined. Reactions of (Z)-bicyclo[6.1.0]non-3-en-2-one and (E)-bicyclo[13.1.0]hexadec-3-en-2-one with dimethyloxosulfonium methylide were highly diastereoselective, while reaction of (E)-bicyclo[10.1.0]tridec-3-en-2-one with this reagent was not diastereoselective. In contrast, 1,4-additions of lithium diorganocuprates were highly diastereoselective for the 8- and 12-membered enones but were not diastereoselective for the 15-membered enone. All reactions were chemically efficient. The diastereoselectivities observed for 1,2-additions, which are thought to involve early transition states, can be rationalized by consideration of the low-energy conformations of each cyclopropyl enone. The diastereoselectivities observed for 1,4-additions, which may involve late transition states, do not correlate simply with the lowest energy conformations of these enones.  相似文献   

18.
Conjugate addition of lithium dimethylcuprate to acyclic alpha,beta-unsaturated esters of varying lengths bearing terminal alkyl or phenyl groups leads to a preponderance of syn 1,3-adducts when one methyl is already present. Conversion to enoates, and iteration of cuprate additions also favors syn adducts to give contiguous deoxypropionate units in a growing chain. The effect of end-group variation (Me, i-Pr, phenyl, tert-butyl) in conjunction with the nature of the ester group (Me, tert-butyl, etc.) on the diastereoselectivity of syn and anti products was studied. The results are rationalized in terms of inductor effects related to the minimization of the 1,5-pentane interactions in energetically favored folded conformations and corroborated by homodecoupling NMR studies.  相似文献   

19.
A general method for the preparation of α-phenylselanyl enones is described. Phosphorus ylides react with these α-phenylselanyl enones in a 1,4-addition, leading to cyclopropanes and/or dihydrofurans, depending on the substitution pattern. This unusual reactivity is due to the phenylselanyl moiety, hindering the carbonyl of the enone and making it less prone to 1,2-additions or promoting conjugate addition by electronic effects.  相似文献   

20.
Several types of lithium organocuprates (I) e.g. R2CuLi, RR'CuLi, R3CuLi2, R2CuLi2X (X- CN, SCN), were prepared and used as polymerization initiators for methyl methacrylate in THF at temperatures ranging from −30 to +20 *C. Polymerization initiated by a mixed homocuprate, e.g. lithium n-butyl (thien-2-yl)cuprate was studied in more detail. Kinetic measurements indicated an overall large (−4 to −14 kcal) apparent negative activation energy of polymerization in the −30 + +20 *C temperature range. The reaction was shown to have a kinetic order of about 1.5 with respect to initiator concentration. Conductance measurements indicate that the equivalent conductance increases with initiator concentration in the 10−2 + 10−1 M range indicating the formation of multiple ions formed by intermolecular association. The data are consistent with a polymerization involving aggregated ion pairs of as yet undetermined composition. The negative activation energy is tentatively interpreted as due to the exothermic formation of aggregates existing as highly solvated ion pairs.  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