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 109 毫秒
1.
The reactions of diazomethane, diazoethane, and (trimethylsilyl)diazomethane with (S)-2-p-tolylsulfinylcyclopent-2-en-1-one have been studied. The sulfinyl group increases the reactivity and controls the π-facial and endo/exo selectivities. The π-facial selectivity can be inverted in the presence of Yb(OTf)3, which makes possible the stereodivergent synthesis of both diastereoisomeric pyrazolines. Completely stereoselective denitrogenation of optically pure pyrazolines into cyclopropanes was achieved under substoichiometric Yb(OTf)3 catalysis.  相似文献   

2.
A highly selective copper-catalyzed trifunctionalization of allenes has been established based on diborylation/cyanation with bis(pinacolato)diboron (B2pin2) and N-cyano-N-phenyl-p-toluenesulfonamide (NCTS). The Cu-catalyzed trifunctionalization of terminal allenes is composed of three catalytic reactions (first borocupration, electrophilic cyanation, and second borocupration) that provide a densely functionalized product with regio-, chemo- and diastereoselectivity. Allene substrates have multiple reaction-sites, and the selectivities are determined by the suitable interactions (e.g., electronic and steric demands) between the catalyst and substrates. We employed DFT calculations to understand the cascade copper-catalyzed trifunctionalization of terminal allenes, providing densely-functionalized organic molecules with outstanding regio-, chemo- and diastereoselectivity in high yields. The selectivity challenges presented by cumulated π-systems are addressed by systematic computational studies; these give insight to the catalytic multiple-functionalization strategies and explain the high selectivities that we see for these reactions.  相似文献   

3.
TiCl4‐induced Baylis–Hillman reactions of α,β‐unsaturated carbonyl compounds with aldehydes yield the (Z)‐2‐(chloromethyl)vinyl carbonyl compounds 5 , which react with 1,4‐diazabicyclo[2.2.2]octane (DABCO), quinuclidine, and pyridines to give the allylammonium ions 6 . Their combination with less than one equivalent of the potassium salts of stabilized carbanions (e.g. malonate) yields methylene derivatives 8 under kinetically controlled conditions (SN2’ reactions). When more than one equivalent of the carbanions is used, a second SN2’ reaction converts 8 into their thermodynamically more stable allyl isomers 9 . The second‐order rate constants for the reactions of 6 with carbanions have been determined photometrically in DMSO. With these rate constants and the previously reported nucleophile‐specific parameters N and s for the stabilized carbanions, the correlation log k (20 °C)=s(N + E) allowed us to calculate the electrophilicity parameters E for the allylammonium ions 6 (?19<E <?18). The kinetic data indicate the SN2’ reactions to proceed via an addition–elimination mechanism with a rate‐determining addition step.  相似文献   

4.
Y. Becker  A. Eisenstadt  Y. Shvo 《Tetrahedron》1976,32(17):2123-2126
A proposition regarding the capability of a complex π-allyl system to undergo a displacement reaction was advanced (eqn 2) and examined experimentally. A unique structural feature of this system is the four rather than three C atoms which intervene between the displaced and entering groups. The experiment was carried out using lactone 1c which is internally complexed and was properly labelled with deuterium. The product of its reaction with aniline has been found to be rearranged with respect to the D label, thus substantiating an SN2′-type mechanism in a complexed π-allyl system (eqn 2). From an intramolecular comparison study of the reactivity of an allyl and complexed π-allyl system in SN2′-type displacement reactions using compound 10, it was concluded that the latter is more reactive than the former. Furthermore, the π-allyl lactone 12, in which the displaced and entering groups are separated by three C atoms, was also found to undergo a displacement reaction.  相似文献   

5.
Kinetic studies for the azo‐coupling reactions of 3‐ethoxythiophene 1 with a series of 4‐X‐substituted diazonium cations 2a‐e (X = OCH3, CH3, H, Cl, and NO2) have been investigated in acetonitrile at 20°C. The second‐order rate constants have been employed to determine the nucleophilicity parameters N and s of the thiophene 1 according the Mayr equation. Thus, the nucleophile‐specific parameters N and s of thiophene 1 have been derived and compared with the reactivities of other C‐nucleophiles in acetonitrile (pyrroles, furan, indoles, etc.). The Yukawa–Tsuno plot resulted in an excellent correlation (R2 = 0.9980) with an r value of 0.89, suggesting that the nonlinear Hammett plot observed in the present work is due to resonance demand of the π–electron donor substituent of on the –N2+ moiety. Importantly, using the concept of global electrophilicity (ω) proposed by Parr, we successfully predict the electrophilicity parameters E of seven substituted diazonium cations whose experimental data are available.  相似文献   

6.
Abstract

Dinuclear Pd(II) halides that contain bridging π-conjugated groups, trans,trans-[(PR3)2(X)Pd–Y–Pd(X)(PR3)2] (X?=?Br; YH2 = terpyridine, fluorenone, benzil, benzthiadiazole), were prepared by the oxidative addition of corresponding dihalo π-conjugated reagents to [Pd(styrene)(PR3)2]. Similar reactions involving dihalobenzil, dihalobithiophene, or dihaloterthiophene afforded dinuclear Pt(II) halides containing bridging π-conjugated groups. Additionally, when the dihalosilole derivatives {2,5-dibromo-1,1-dimethyl (or diphenyl)-3,4-diphenylsilole} reacted with [Pd(styrene)(PR3)2], mono or dinuclear Pd(II) complexes bearing a dimethyl (or diphenyl)-3,4-diphenylsilole group were obtained. π-Conjugation extension reactions of dinuclear bithiophene-bridged Pd(II) halides with HC≡C–R {R?=?SiPh3, C(O)OMe} in the presence of CuI and HNEt2 led to the unexpected formation of bis(acetylide) Pd(II) complexes of the form, [Pd(C≡C–R)2(PR3)2] and bithiophene. In contrast, treatment of the dinuclear Pd(II) halides with two equiv of organic isocyanide resulted in isocyanide insertion into the Pd???C bonds to afford π-conjugation-extended dinuclear Pd(II) compounds bearing a π-conjugated moiety.  相似文献   

7.
The reaction rates of 2-chloro-3,5-dinitropyridine 1 with a series of arylthiolates 2a-h in methanol have been measured at 25°C. The products are the corresponding 2-thioaryl-3,5-dinitropyridine 3a-h. Good Hammett correlation with ρ value −1.19 was obtained suggesting an elimination-addition mechanism SNAr and the formation of Meisenheimer-like intermediates. Plot of log k2 vs. pKa values of arylthiols gave straight line with β=0.38 indicating that the π-bond breaking in the pyridine ring is so much advanced over bond making between the nucleophile and the carbon that bears the chlorine atom. Excellent correlation between log k2 and log K (carbon basicity of arylthiolates) was obtained. © 1997 John Wiley & Sons. Inc. Int J Chem Kinet 29: 515–521, 1997.  相似文献   

8.
A chemoselective C(sp2)−C(sp2) coupling of sufficiently electron-deficient fluorinated arenes and functionalized N-aryl-N’-silyldiazenes as masked aryl nucleophiles is reported. The fluoride-promoted transformation involves the in situ generation of the aryl nucleophile decorated with various sensitive functional groups followed by a stepwise nucleophilic aromatic substitution (SNAr). These reactions typically proceed at room temperature within minutes. This catalytic process allows for the functionalization of both coupling partners, furnishing highly fluorinated biaryls in good yields.  相似文献   

9.
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   

10.
In order to investigate the mechanism of mercuration reaction of substituted ben-zylideneanilines, kinetic measurements of these reactions at different temperatures (40-60℃) inmethanol-l,4-dioxane (1/1, V/V) were carried out and Hammett ρ value for C-phenyl substituentsof-0.61 for the N-(substituted benzylidene)-4-toluidine series was obtained. Thermodynamicparameters E_a, △S~≠ were obtained for the reaction of different. N-(substituted benzylidene)-4-toluidines. It was found that this ortho-mercuration was brought about by an internal cyclometal-lation process involving the imino-moiety.  相似文献   

11.
Changes in the activation parameters of amination along mechanisms S N 2, S N Ar, Ad N and in reactions with acyl group transfer may serve as an additional investigation method for these reactions in solution. The approximation utilizes the substituents effect in benzene and pyridine derivatives on the changes in the activation parameters ΔX (X = H, S, G) in the framework of equations of Hammett’s type for the estimation of the resulting reaction constants δΔX . The single linear dependences of the reaction constants of internal enthalpy δΔH int on the reaction constants δΔG and Hammett’s ρ show that the substituents effect in the leaving and non-leaving groups and in the nucleophiles on the amination mechanisms is governed by the δΔH int , when unique stage of the process determines its rate.  相似文献   

12.
Enantiomerically pure bis‐heterocycles containing a (S)‐proline moiety have been prepared starting from (S)‐N‐benzylprolinehydrazide ( 2b ). The reactions with isothiocyanates or butyl isocyanate in refluxing MeOH led to the corresponding thiosemicarbazide 5 and semicarbazide 9 with a N‐benzylprolinoyl residue. The structure of the tert‐butyl derivative 5d was established by X‐ray crystallography. Base‐catalyzed cyclization of 5 and 9 led to (S)‐3‐(pyrrolidin‐2‐yl)‐1H‐1,2,4‐triazole‐5(4H)‐thiones 6 and the corresponding 5(4H)‐one 8 , respectively, whereas, in concentrated H2SO4, compounds 5 undergo cyclization to give (S)‐5‐amino‐2‐(pyrrolidin‐2‐yl)‐1,3,4‐thiadiazoles 7 . Furthermore, 2b reacted with hexane‐2,5‐dione in boiling iPrOH to yield the (S)‐N‐(2,5‐dimethylpyrrol‐1‐yl)prolinamide 10 . In the case of the bis‐heterocycle 8 , treatment with HCOONH4 and Pd/C in MeOH gave the debenzylated product 12 .  相似文献   

13.
The kinetics of (3+2) cycloaddition reactions of 18 different donor–acceptor cyclopropanes with the same aldehyde were studied by in situ NMR spectroscopy. Increasing the electron density of the donor residue accelerates the reaction by a factor of up to 50 compared to the standard system (donor group=phenyl), whereas electron‐withdrawing substituents slow down the reaction by a factor up to 660. This behavior is in agreement with the Hammett substituent parameter σ. The obtained rate constants from the (3+2) cycloadditions correlate well with data from additionally studied (3+n) cycloadditions with a nitrone (n=3) and an isobenzofuran (n=4). A comparison of the kinetic data with the bond lengths in the cyclopropane (obtained by X‐ray diffraction and computation), or the 1H and 13C NMR shifts, revealed no correlation. However, the computed relaxed force constants of donor–acceptor cyclopropanes proved to be a good indicator for the reactivity of the three‐membered ring.  相似文献   

14.
Density Functional Theory studies of square-planar PtII pincer structures, (4-Z-NCN)PtCl ([4-Z-NCN]=[4-Z-2,6-(Me2NCH2)2C6H2-N,C,N], Z=H, NO2, CF3, CO2H, CHO, Cl, Br, I, F, SMe, SiMe3, tBu, OH, NH2, NMe2), enable characterisation of mesomerism for the pincer-Pt interaction. Relationships between Hammett σp substituent parameters of Z and DFT data obtained from NBO6 and AOMix computation are used to probe the interaction of the 5dyz orbital of platinum with π-orbitals of the arene ring. Analogous computation for 2,6-(Me2CH2)2C6H3Z (Z=H, CF3, CHO, Cl, Br, I, F, SMe, SiMe3, tBu, OH, NH2) and (4-H-NCN)PtZ allows an estimation of the relative substituent effects of “(CH2NMe2)2PtZ” on π-delocalisation in the pincer system.  相似文献   

15.
Rhodium-catalyzed C−H insertions and cyclopropanations of donor/acceptor carbenes have been used for the synthesis of positional analogues of methylphenidate. The site selectivity is controlled by the catalyst and the amine protecting group. C−H functionalization of N-Boc-piperidine using Rh2(R-TCPTAD)4, or N-brosyl-piperidine using Rh2(R-TPPTTL)4 generated 2-substitited analogues. In contrast, when N-α-oxoarylacetyl-piperidines were used in combination with Rh2(S-2-Cl-5-BrTPCP)4, the C−H functionalization produced 4-susbstiuted analogues. Finally, the 3-substituted analogues were prepared indirectly by cyclopropanation of N-Boc-tetrahydropyridine followed by reductive regio- and stereoselective ring-opening of the cyclopropanes.  相似文献   

16.
The kinetics and mechanism of nucleophilic aromatic substitution reactions of 4‐chloro‐7‐nitrobenzofurazan 1 with 4‐X‐substituted anilines 2a–g (X = OH, OCH3, CH3, H, I, Cl, and CN) are investigated in a dimethyl sulfoxide (Me2SO) solution at 25°C. The Hammett plot of log k1 versus σ is nonlinear for all the anilines studied due to positive deviations of the electron‐donating substituents. However, the corresponding Yukawa–Tsuno plot resulted in a good linear correlation with σ+r (σ+?σ). The corresponding Brønsted‐type plot is also nonlinear, i.e., the slope (βnuc) changes from 1.60 to 0.56 as the basicity of anilines decreases. These results indicate a change in a mechanism from a polar SNAr process for less basic nucleophiles (X = I, Cl, and CN) to a single electron transfer for more basic nucleophiles (X = OH, OCH3, and CH3). The satisfactory log k1 versus Eo correlation obtained for the reactions of 1 with anilines 2a–d in the present system is consistent with the proposed mechanism. Interestingly, the βnuc = 1.60 value measured for 1 in Me2SO reflects one of the highest coefficients Brønsted ever observed for SNAr reactions. © 2013 Wiley Periodicals, Inc. Int J Chem Kinet 45: 152–160, 2013  相似文献   

17.
In the title compound, (4‐O2NC6H4)2S2 or C12H8N2O4S2, the mol­ecules lie across twofold rotation axes. A single type of C—H?O hydrogen bond, with C?O = 3.394 (3) Å and C—H?O = 158°, links the mol­ecules into continuous two‐dimensional sheets built from a single type of R44(44) ring. These sheets are linked by aromatic π?π stacking interactions to form a continuous three‐dimensional framework.  相似文献   

18.
A quantum chemical investigation of the Bu4N[Fe(CO)3(NO)]‐catalyzed Cloke–Wilson rearrangement of vinyl cyclopropanes is reported. It was found that allylic C?C bond activation can proceed through a SN2′ or SN2‐type mechanism. The application of the recently reported intrinsic bond orbital (IBO) method for all structures indicated that one Fe?N π bond is directly involved. Further analysis showed that during the reaction oxidation occurs at the NO ligand exclusively.  相似文献   

19.
In order to explore the existence of α‐effect in gas‐phase SN2@N reactions, and to compare its similarity and difference with its counterpart in SN2@C reactions, we have carried out a theoretical study on the reactivity of six α‐oxy‐Nus (FO?, ClO?, BrO?, HOO?, HSO?, H2NO?) in the SN2 reactions toward NR2Cl (R = H, Me) and RCl (R = Me, i‐Pr) using the G2(+)M theory. An enhanced reactivity induced by the α‐atom is found in all examined systems. The magnitude of the α‐effect in the reactions of NR2Cl (R = H, Me) is generally smaller than that in the corresponding SN2 reaction, but their variation trend with the identity of α‐atom is very similar. The origin of the α‐effect of the SN2@N reactions is discussed in terms of activation strain analysis and thermodynamic analysis, indicating that the α‐effect in the SN2@N reactions largely arises from transition state stabilization, and the “hyper‐reactivity” of these α‐Nus is also accompanied by an enhanced thermodynamic stability of products from the n(N) → σ*(O?Y) negative hyperconjugation. Meanwhile, it is found that the reactivity of oxy‐Nus in the SN2 reactions toward NMe2Cl is lower than toward i‐PrCl, which is different from previous experiments, that is, the SN2 reactions of NH2Cl is more facile than MeCl. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
The bis‐thionooxalamic acid esters trans‐(±)‐diethyl N,N′‐(cyclohexane‐1,2‐diyl)bis(2‐thiooxamate), C14H22N2O4S2, and (±)‐N,N′‐diethyl (1,2‐diphenylethane‐1,2‐diyl)bis(2‐thiooxamate), C22H24N2O4S2, both consist of conformationally flexible molecules which adopt similar conformations with approximate C2 rotational symmetry. The thioamide and ester parts of the thiooxamate group are significantly twisted along the central C—C bond, with the S=C—C=O torsion angles in the range 30.94 (19)–44.77 (19)°. The twisted scis conformation of the thionooxamide groups facilitates assembly of molecules into a one‐dimensional polymeric structure via intermolecular three‐center C=S...NH...O=C hydrogen bonds and C—H...O interactions formed between molecules of the opposite chirality.  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