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.
Reaction solvent was previously shown to influence the selectivity of Pd/PtBu3-catalyzed Suzuki–Miyaura cross-couplings of chloroaryl triflates. The role of solvents has been hypothesized to relate to their polarity, whereby polar solvents stabilize anionic transition states involving [Pd(PtBu3)(X)] (X = anionic ligand) and nonpolar solvents do not. However, here we report detailed studies that reveal a more complicated mechanistic picture. In particular, these results suggest that the selectivity change observed in certain solvents is primarily due to solvent coordination to palladium. Polar coordinating and polar noncoordinating solvents lead to dramatically different selectivity. In coordinating solvents, preferential reaction at triflate is likely catalyzed by Pd(PtBu3)(solv), whereas noncoordinating solvents lead to reaction at chloride through monoligated Pd(PtBu3). The role of solvent coordination is supported by stoichiometric oxidative addition experiments, density functional theory (DFT) calculations, and catalytic cross-coupling studies. Additional results suggest that anionic [Pd(PtBu3)(X)] is also relevant to triflate selectivity in certain scenarios, particularly when halide anions are available in high concentrations.

In the presence of the bulky monophosphine PtBu3, palladium usually prefers to react with Ar–Cl over Ar–OTf bonds. However, strongly coordinating solvents can bind to palladium, inducing a reversal of selectivity.

Oxidative addition is a key elementary step in diverse transformations catalyzed by transition metals.1 For instance, this step is common to traditional cross-coupling reactions, which are among the most widely used methods for small molecule synthesis. During the oxidative addition step of cross-coupling reactions, a low valent metal [usually Pd(0)] inserts into a C–X bond with concomitant oxidation of the metal by two electrons. The “X” group of the C–X bond is commonly a halogen or triflate. Despite a wealth of research into this step,2–5 uncertainties remain about its mechanistic nuances. The mechanistic details are especially pertinent to issues of selectivity that arise when substrates contain more than one potentially reactive C–X bond.6One of the best-studied examples of divergent selectivity at the oxidative addition step is the case of Pd-catalyzed Suzuki couplings of chloroaryl triflates. In 2000, Fu reported that a combination of Pd(0) and PtBu3 in tetrahydrofuran (THF) effects selective coupling of 1 with o-tolylB(OH)2via C–Cl cleavage, resulting in retention of the triflate substituent in the final product 2a (Scheme 1A).7 In contrast, the use of PCy3 (ref. 7) or most other phosphines8 provides complementary selectivity (product 2b) under similar conditions. The unique selectivity imparted by PtBu3 was later attributed to this ligand''s ability to promote a monoligated oxidative addition transition state on account of its bulkiness.5,8 Smaller ligands, on the other hand, favor bisligated palladium, which prefers to react at triflate. The relationship between palladium''s ligation state and chemoselectivity has been rationalized by Schoenebeck and Houk through a distortion/interaction analysis.5 In brief, the selectivity preference of PdL2 is dominated by a strong interaction between the electron-rich Pd and the more electrophilic site (C–OTf). On the other hand, PdL is less electron-rich and its selectivity preference mainly relates to minimizing unfavorable distortion energy by reacting at the more easily-distorted C–Cl bond.Open in a separate windowScheme 1Seminal reports on the effects of (A) ligands and (B) solvents on the selectivity of cross-coupling of a chloroaryl triflate.5,7,9Proutiere and Schoenebeck later discovered that replacing THF with dimethylformamide (DMF, Scheme 1B, entry 1) or acetonitrile caused a change in selectivity for the Pd/PtBu3 system.9,10 In these two polar solvents, preferential reaction at triflate was observed, and PtBu3 no longer displayed its unique chloride selectivity. The possibility of solvent coordination to Pd was considered, as bisligated Pd(PtBu3)(solv) would be expected to favor reaction at triflate. However, solvent coordination was ruled out on the basis of two intriguing studies. First, DFT calculations using the functional B3LYP suggested that solvent-coordinated transition states are prohibitively high in free energy (about 16 kcal mol−1 higher than the lowest-energy monoligated transition structure). Second, the same solvent effect was not observed in a Pd/PtBu3-catalyzed base-free Stille coupling in DMF (Scheme 1B, entry 2). Instead, the Stille coupling was reported to favor reaction at chloride despite the use of a polar solvent. This result appears inconsistent with the possibility that solvent coordination induces triflate-selectivity, as coordination of DMF to Pd should be possible in both the Stille and Suzuki conditions, if it happens at all. Instead, it was proposed that the key difference between the Suzuki and Stille conditions was the absence of coordinating anions in the latter (unlike traditional Suzuki couplings, Stille couplings do not necessarily require basic additives such as KF to promote transmetalation). Indeed, when KF or CsF was added to the Stille reaction in DMF, selectivity shifted to favor reaction at triflate (Scheme 1B, entry 3), thereby displaying the same behavior as the Suzuki coupling in this solvent. On the basis of this and the DFT studies, it was proposed that polar solvents induce a switch in chemoselectivity if coordinating anions like fluoride are available by stabilizing anionic bisligated transition structures (Scheme 1B, right).However, our recent extended solvent effect studies produced confounding results.11 In a Pd/PtBu3-catalyzed Suzuki cross-coupling of chloroaryl triflate 1, we observed no correlation between solvent polarity and chemoselectivity (Scheme 2). Although some polar solvents such as MeCN, DMF, and dimethylsulfoxide (DMSO) favor reaction at triflate, a number of other polar solvents provide the same results as nonpolar solvents by favoring reaction at chloride. For example, cross-coupling primarily takes place through C–Cl cleavage when the reaction is conducted in highly polar solvents like methanol, water, acetone, and propylene carbonate. In fact, the only solvents that promote reaction at triflate are ones that are commonly thought of as “coordinating” in the context of late transition metal chemistry.12 These are solvents containing nitrogen, sulfur, or electron-rich oxygen lone pairs (nitriles, DMSO, and amides). The observed solvent effects were upheld for a variety of chloroaryl triflates and aryl boronic acids.11Open in a separate windowScheme 2Expanded solvent effect studies in the Pd/PtBu3-catalyzed Suzuki coupling.11We have sought to reconcile these observations with the earlier evidence9 against solvent coordination. Herein we report detailed mechanistic studies indicating that coordinating solvents alone are sufficient to induce the observed selectivity switch. In solvents like DMF and MeCN, stoichiometric oxidative addition is favored at C–OTf even in the absence of anionic additives. The apparent contradiction between our observations and the previously-reported DFT calculations and base-free Stille couplings is reconciled by a reevaluation of those studies. In particular, when dispersion is considered in DFT calculations, neutral solvent-coordinated transition structures involving Pd(PtBu3)(solv) become energetically feasible. Furthermore, we find that the selectivity analysis in the Stille couplings is convoluted by low yields, the formation of side products, and temperature effects. When these factors are disentangled, the Stille coupling in DMF displays selectivity similar to the Suzuki coupling in the same coordinating solvent. In light of these new results, anionic bisligated [Pd(PtBu3)(X)] does not appear to be the dominant active catalyst in nonpolar or polar solvents unless special measures are taken to increase the concentration of free halide, such as adding tetraalkylammonium halide salts or crown ethers.  相似文献   

2.
By performing CASPT2 calculations, the lowest energy pathway for oxygen addition to an isolated heme center of a heme-protein is evaluated and found to be reversible (the oxyheme compound is just 14.9 kcal mol(-1) more stable than the deoxyheme + O(2) reactants, and the energy barriers to dissociation are even smaller).  相似文献   

3.
The aim of this work is to better understand the interaction between the confined dihydrogen molecule and armchair (2,2), (3,3) (4,4), (5,5), and (6,6) single‐walled carbon nanotubes (SWNT) using Restricted Hartree–Fock (RHF) and Density Functional Theory (DFT) methods using B3LYP and CAM‐B3LYP functionals. Depending on the calculation method and its orientation inside the nanotube, H2 binds differently. We found that H? H bond length increases when H2 is trapped in CNT (2,2) and decreases for CNT (3,3) and (4,4). The characteristics of confined H2 in (5,5) and (6,6) nanotubes are similar to H2 in a free state. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
The mechanism of addition reaction between carbene and epoxyethane has been investigated employing the MP2 and B3LYP/6-311+G* levels of theory. Geometry optimization, vibrational analysis, and energy property for the involved stationary points on the potential energy surface have been calculated. Based on the calculated results at the MP2/6-311+G* level of theory, it can be predicted that there are two reaction mechanisms (1) and (2). In the first reaction carbene attacks the atom O of epoxyethane to form an intermediate 1a (IM1a), which is a barrier-free exothermic reaction. Then, IM1a can isomerize to IM1b via a transition state 1a (TS1a), where the potential barrier is 48.6 kJ/mol. Subsequently, IM1b isomerizes to a product epoxypropane (Pro1) via TS1b with a potential barrier of 14.2 kJ/mol. In the second carbene attacks the atom C of epoxyethane firstly to form IM2 via TS2a. Then IM2 isomerizes to a product allyl alcohol (Pro2) via TS2b with a potential barrier of 101.6 kJ/mol. Correspondingly, the reaction energies for the reactions (1) and (2) are −448.4 and −501.6 kJ/mol, respectively. Additionally, the orbital interactions are also discussed for the leading intermediate. The results based on the B3LYP/6-311+G* level of theory are paralleled to those on the MP2/6-311+G* level of theory. Furthermore, the halogen and methyl substituent effects of H2C: on the two reaction mechanisms have been investigated. The calculated results indicate that the introductions of halogen or methyl make the addition reaction difficult to proceed.  相似文献   

5.
The nucleophilic addition of the malononitrile anion (MN) to formaldehyde was studied theoretically by the AM1 semiempirical MO method. The addition is found to be endothermic with a late productlike transition state on the reaction coordinate. Additions of MN to a series of carbonyl compounds were studied in order to investigate the substituent effect on the energetics of the title addition and the nucleophilic attack reactivity. The solvent effect was stimulated by hydrogen bonding a single molecule of water to the formaldehyde oxygen and/or to the MN anion. Its influence on the energetics and the transition-state geometry was estimated. The Hammond postulate was satisfied for the studied additions. © 1997 John Wiley & Sons, Inc. Int J Quant Chem 62 : 419–426, 1997  相似文献   

6.
The electronic mechanisms of the cyclic processes of photochemical reductive elimination of H2 from [IrClH2(PH3)3] and thermal oxidative addition of H2 to [IrCl(PH3)3] are investigated theoretically. The geometries of the ground and excited states are optimized using the Hartree-Fock and single excitation configuration interaction methods, respectively, and higher level calculations for the ground and excited states are carried out by the symmetry adapted cluster (SAC)/SAC–configuration interaction method. The present calculation shows that the reductive elimination of H2 from [IrClH2(PH3)3] dose not occur thermally but photochemically through diabatic conversion from the lowest A′ excited state to the ground state (A′), while the oxidative addition of H2 to [IrCl(PH3)3] easily proceeds thermally. The lowest 1A′ excited state involves the nature of the Ir-H2 antibonding. Received: 10 December 1997 / Accepted: 16 January 1998 / Published online: 17 June 1998  相似文献   

7.
In this paper, the high reactivity of silyl macroradicals toward double bonds of olefinic compounds has been explained by means of quantum‐mechanical calculations through their frontier orbital characteristics. In this way, the main orbital interaction corresponds to the overlapping between the SOMO of the disilyl radical and the LUMO of the olefin. In order to obtain more accurate results of differential reactivity, an orbitalic SOMO‐HOMO interaction should be included in addition to the main SOMO‐LUMO one. Also, we theoretically studied the regioselectivity of the addition of silyl radicals to double bonds obtaining similar results as for carbon centered radicals where the reaction takes place on the less hindered carbon of the olefin. Regarding to the geometrical and electronic parameters, it has been shown that carbon radicals have a sp2 geometry and a negative charge on the radical center whilst silyl radicals have a sp3 goemetry and a positive charge. Both factors contribute to the enhanced reactivity of silyl radicals with respect to carbon ones.  相似文献   

8.
9.
Ab initio calculations were used to analyze interactions of BH4 ? with 1?C4 molecules of H2O at the MP2/6-311++G(d,p) and B3LYP/6-311++G(d,p) computational levels. The negative cooperativity for dihydrogen bond clusters containing H2O···H2O hydrogen bonds is more remarkable. The negative cooperativity is increased with increasing the size and also the number of hydrogen bonds in the cluster. The B?CH stretching frequencies show blue shifts with respect to cluster formation. Also greater blue shift of stretching frequencies where predicted for B?CH bonds which did not contribute in dihydrogen bonding with water molecules. The structures obtained have been analyzed with the Atoms in Molecules (AIM) methodology.  相似文献   

10.
The mechanism of dihydrogen addition to Cp(2)Ta(CH(2))(H) was examined using parahydrogen-induced polarization (PHIP), (13)C labeling, and comparison to the related complex Cp(2)Ta(CH(2))(CH(3)). The reaction of para-enriched hydrogen with Cp(2)Ta(CH(2))(H) leads to polarized resonances for both Cp(2)Ta(CH(3))(H)(2) and Cp(2)TaH(3), even at the earliest reaction times. Use of the labeled compound Cp(2)Ta((13)CH(2))(H) shows that the polarized resonances of Cp(2)Ta(CH(3))(H)(2) correspond to the two hydride ligands. The results thus support a mechanistic pathway of H(2) addition to an unsaturated Ta(III) intermediate, [Cp(2)Ta(CH(3))], rather than addition directly across the Ta=C bond. In a same sample comparison, the rates of initial H(2) addition and subsequent C-H reductive elimination for both Cp(2)Ta(CH(2))(H) and Cp(2)Ta(C(6)H(4))(H) were examined. The methylene complex exhibits greater reactivity than the benzyne complex, with the major difference due to the C-H coupling step, in which formation of methane is more facile than that of benzene. The reactivity of the related ethylene hydride complex, Cp(2)Ta(C(2)H(4))(H), with hydrogen was also examined. The PHIP study of this system leads to unusual and unexpected polarization, which is found to be due to a minor impurity in the sample.  相似文献   

11.
The stabilized silene 1,1-bis(trimethylsilyl)-2-adamantylidenesilane (4) has been generated by photolysis of a novel trisilacyclobutane derivative in various solvents and studied directly by kinetic UV spectrophotometry. Silene 4 decays with second-order kinetics in degassed hexane solution at 23 degrees C (k/epsilon = 8.6 x 10(-6) cm s(-1)) due to head-to-head dimerization. It reacts rapidly with oxygen [k(25 degrees C) approximately 3 x 10(5) M(-1) s(-1)] but approximately 10 orders of magnitude more slowly with methanol (MeOH) than other silenes that have been studied previously. The data are consistent with a mechanism involving reaction with the hydrogen-bonded dimer of the alcohol, (MeOH)(2) (k = 40 +/- 3 M(-1) s(-1); k(H)/k(D) = 1.7 +/- 0.2). The stable analogue of silene 4, 1-tert-butyldimethylsilyl-1-trimethylsilyl-2-adamantylidenesilane (5), reacts approximately 50 times more slowly, but via the same mechanism. The mechanism for addition of water and methanol (ROH; R = H, Me) to 4, 5, and the model compound 1,1-bis(silyl)-2,2-dimethylsilene (3a) has been studied computationally at the B3LYP/6-31G(d) and MP2/6-31G(d) levels of theory. Hydrogen-bonded complexes with monomeric and dimeric methanol, in which the Si=C bond plays the role of nucleophile, have been located computationally for all three silenes. Reaction pathways have been characterized for reaction of the three silenes with monomeric and dimeric ROH and reveal significantly lower barriers for reaction with the dimeric form of the alcohol in each case. The calculations indicate that 5 should be approximately 40-fold less reactive toward dimeric MeOH than 4, in excellent agreement with the approximately 50-fold difference in the experimental rate constants for reaction in hexane solution.  相似文献   

12.
13.
Ni-doped germanium clusters have been systematically investigated by using the density functional approach. The growth-pattern behaviors, stabilities, charge transfer, and polarities of these clusters are discussed in detail. Obviously different growth patterns appear between small-sized Ni-doped germanium clusters and middle- or larger-sized Ni-doped germanium clusters. The Ni-convex or substituted Ge(n) frames for small-sized clusters as well as Ni-concaved or encapsulated Ge(n) frames for middle- or large-sized clusters are dominant growth patterns. The calculated fragmentation energies manifest that the magic numbers of stabilities are 5, 8, 10, and 13 for Ni-doped germanium clusters; the obtained relative stabilities exhibit that the Ni-encapsulated Ge(10) cluster is the most stable species of all different-sized clusters, which is in good agreement with available experimental observations of CoGe(10)(-). Natural population analysis shows that different charge-transfer phenomena depend on the sizes of the Ni-doped Ge(n) clusters. Additionally, the properties of frontier orbitals and the polarities of Ni-doped Ge(n) clusters are also discussed.  相似文献   

14.
Density functional theory with the B3LYP hybrid functional and 6–31G* basis set was used to study the geometric and electronic structure of H2C = CHR (R = H, CH3, C2H5, C3H7, C4H9, and C5H11) olefins, their carbocations formed in the addition of the proton to the olefins, R′-S-H aliphatic thiols (R′ = H, CH3, C2H5, and C3H7), the products of the addition of thiols to carbocations, and the final products of the addition of thiols to olefins. The proton affinity of the olefins and the products of the addition of thiols to olefins was calculated. The conclusion was drawn that the limiting stage in the nonradical addition of thiols to olefins catalyzed by acids was proton transfer from the protonated reaction product to the olefin. The theoretical results were compared with the experimental data on the electrophilic addition of polymercaptan to heptene-1.  相似文献   

15.
Pamela Moles 《Tetrahedron》2008,64(40):9448-9463
A theoretical study on the artemisinin decomposition mechanism is reported. The suggested pathways have been reproduced and the appearance of the final products can be explained in a satisfactory way. In addition, several intermediates and radicals have been found as relatively stable species, thus giving support to the current hypothesis that some of these species can be responsible for the antimalarial action of artemisinin and its derivatives.  相似文献   

16.
The mechanism of the boron metathesis reaction of the transition-metal-aminoborylene complex Cp(CO)(2)FeBN(CH(3))(2+) (8) with EX, where EX = H(3)PO (9ap), H(3)AsO (9bp), H(3)PS (9aq), H(3)AsS (9bq), CH(3)CHCH(2) (9cr), (NH(2))(2)CCH(2) (9dr), H(2)CO (9ep), and (NH(2))(2)CO (9dp) is investigated at the B3LYP/LANL2DZ level. The analysis of bonding and charge distribution shows that the Fe-borylene complex (8) is a Fischer-type carbene analogue. The attack of the olefin takes place at the metal end of the M=C bond of the metal-carbene complex in olefin metathesis and proceeds via [2 + 2] cycloaddition, while in boron metathesis, the initial attack of the substrates takes place at the positively charged B atom of the Fe-borylene complex and forms the preferred acyclic intermediate. The energetics of boron metathesis is comparable to that of the olefin metathesis. Substrates that are polar and a have low-lying sigma* molecular orbital (weak sigma bond) prefer the boron metathesis reaction. The relative stability of the metathesis products is controlled by the strength of the Fe-E and B-X bonds of the products 13 and 14, respectively. We have also investigated the possibility of a beta-hydride-transfer reaction in the Fe-borylene complex.  相似文献   

17.
18.
Quantum chemical calculations using density functional theory (B3LYP) were carried out to elucidate the reaction pathways for ethylene addition to the chromium and molybdenum complexes CrO(CH3)2(CH2) (Cr1) and MoO(CH3)2(CH2) (Mo1). The results are compared with previously published results of the analogous tungsten system WO(CH3)2(CH2) (W1). The comparison of the group-6 elements shows that the molybdenum and tungsten compounds Mo1 and W1 have a similar reactivity while the chromium compound has a more complex reactivity pattern. The kinetically most favorable reaction pathway for ethylene addition to Mo1 is the [2+2]Mo,C addition across the MoCH2 double bond which has an activation barrier of only 8.4 kcal/mol. The reaction is slightly exothermic with ΔER = −0.6 kcal/mol. The [2+2]Mo,O addition across the MoO double bond and the [3+2]C,O addition have much higher barriers and are strongly endothermic. The thermodynamically mostly favored reaction is the [1+2]Mo addition of ethylene to the metal atom which takes place after prior rearrangement of the Mo(VI) compound Mo1 to the Mo(IV) isomer Mo1g. The reaction is −19.2 kcal/mol exothermic but it has a large barrier of 34.5 kcal/mol. The kinetically and thermodynamically most favorable reaction pathway for ethylene addition to the chromium homologue Cr1 is the multiple-step process with initial rearrangements Cr1 → Cr1c → Cr1g which are followed by a [1+2]Cr addition yielding an ethylene π complex Cr1g + C2H4 → Cr1g-1. The highest barrier comes from the first step Cr1 → Cr1c which has an activation energy of 14.2 kcal/mol. The overall reaction is exothermic by −26.3 kcal/mol.  相似文献   

19.
20.
用MINDO/3方法求出了N-4-戊烯基硝酮分子内环加成反应的过渡态和反应途径.两个环加成区域异构体是由N-4-戊烯基硝酮的两个不同的构象经过各自的过渡态得到的. 理论分析满意地解释了实验结果.  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