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.
When two different substituents are placed in the nonbridgehead position of adamantane, the two [1A(ax), 1B(eq)] and [1A(eq), 1B(ax)] cyclohexane chair conformers are modeled and features of their NMR spectra can be studied from a single spectrum at 298 K. The effect of [1A(ax), 1B(eq)] and [1A(eq), 1B(ax)] cyclohexane ring substitution on the 1H resonance separation within the γ-CH2s of cyclohexane ring is compared for various substituent pairs; this aim is approached by measuring the 1H chemical shift separation within the 4′,9′-H and 8′,10′-H methylenes from the 1H NMR spectrum of the model 2A,2B-disubstituted adamantane at 298 K.  相似文献   

2.
Earlier we noted that the hydroboration of the trimethylsilyl enol ether of an acyclic ketone results in an elimination of a trimethylsiloxyborane moiety with the subsequent formation of an olefin.1,2 The olefin formed then undergoes hydroboration giving a monoalcohol upon oxidation. (eq 1) We wish to report here on the utility of this sequence, illustrated in eq 1, in the reductive 1,2 transposition of acyclic ketones.3  相似文献   

3.
3-Oxo-1,3-oxathiane (1) and its monomethyl derivatives were prepared by oxidation of the corresponding 1,3-oxathianes. The structural analysis was carried out by1H and13C NMR, FTIR, and mass spectrometry. At 298 K compound1 was a 1 1 (at 173 K a 3 1) mixture of the SO(ax) and SO(eq) chair forms. The major oxidation products of methyl 1,3-oxathianes attained exclusively the SO(ax), Me(eq) chair forms except that of the 5-methyl derivative, which consisted of 7% of the SO(eq), Me(ax) chair conformation in CDCl3 solution. The minor products of oxidation existed in anancomeric SO(eq), Me(eq) chair conformations. The oxidation of 2-methyl- 1,3-oxathiane, however, led to 3,3-dioxo derivative (6) in addition to thetrans [SO(eq)] monoxide. The crystal structures of6 andtrans-3-oxo-5-methyl-1,3-oxathiane were solved by X-ray diffractometry.  相似文献   

4.
A microwave-assisted fluorination method for 1,3-dicarbonyl compounds using Selectfluor® has been developed. 2-Monofluorinated products can be obtained in high yield in neutral reaction conditions with addition of 1 eq. of Selectfluor®. Treatment of 1,3-dicarbonyls with 3 eq. of Selectfluor® in the presence of tetrabutylammonium hydroxide (TBAH) as the base results in the formation of 2,2-difluorinated derivatives only.  相似文献   

5.
Quantum chemical analysis is presented, motivated by Grée and co-workers’ observation of salt effects [Adv. Synth. Catal. 2006, 348, 1149–1153] for SN2 fluorination of KF in ionic liquids (ILs). We examine the relative promoting capacity of KF in [bmim]PF6 vs. [bmim]Cl by comparing the activation barriers of the reaction in the two ILs. We also elucidate the origin of the experimentally observed additional rate acceleration in IL [bmim]PF6 achieved by adding KPF6. We find that the anion PF6 in the added salt acts as an extra Lewis base binding to the counter-cation K+ to alleviate the strong Coulomb attractive force on the nucleophile F, decreasing the Gibbs free energy of activation as compared with that in its absence, which is in good agreement with experimental observations of rate enhancement. We also predict that using 2 eq. KF together with an eq. KPF6 would further activate SN2 fluorination  相似文献   

6.
The synthesis and variable temperature 1H and 13C NMR spectra of three tetrahydro-1,2,4-oxadiazines are reported. The N(4)-Me inversion barriers are 6.8–7.0 (ax→ts) and 7.4–7.9 kcal mol?1 (eq→ts) with ΔG° 0.6–0.9 kcal mol?1. The N(2)-Me inversion barriers are 10.4–11.4 (ax→ts) and 11.6–13.1 kcal mol?1 (eq→ts) with ΔGδ 1.2–1.7 kcal mol?1. The barrier to ring inversion is ca. 12.7 kcal mol?1. “R value” analysis shows the ring to have a 56.5±2δ dihedral angle about the C(5)-(6) bond, indicative of the expected chair conformation.  相似文献   

7.
Transesterification reactions of o-nitrophenyl carbonates have been found to be greatly facilitated by use of p-dialkylaminopyridines as catalysts. A 20-fold rate increase is observed using 0.01 eq. of p-dimethylaminopyridine in place of 2.00 eq of triethylamine. Such rate enhancement is not observed with p-nitrophenyl carbonates: 2.00 eq of Et3N causes reaction to occur 8.5 times faster than 0.01 eq of DMAP. DMAP catalyzed reactions of o-nitrophenyl carbonate have been used to prepare aromatic carbonates of 0° in high yields, and avoid the necessity of using 2 equiv. of base.  相似文献   

8.
Formation of the complex of ethyl viologen in its cationic (Ev+?) and neutral (Ev°) forms with β-cyclodextrin (β-CD) was investigated by means of voltammetric technique in buffer solution of pH 7.00. The number of βCD (n or m) per viologen species (Ev+?) or (Ev°), bonding equilibrium constants as well as bonding rate constants was calculated. The calculated values of $K_{\text{eq}}^{(1)}$ and $K_{\text{eq}}^{ ( 2)}$ (pertaining to the bonding of Ev+? and Ev° with βCD) are 13.6 Mn and 2.1 × 103 M?m , respectively, whereas the calculated values of n and m are 0.54 and 1.25, respectively. The bimolecular rate constant for the Ev°?βCD inclusion complex formation is 3.03 × 103 M?1s?1. These results are supported by the simulation of the experimental cyclic voltammograms. This study also highlights the significance of the proposed electrochemical method as compared to earlier studies on viologen-Cyclodextrin systems.  相似文献   

9.
Detailed insights into the excited-state enol(N*)-keto(T*) intramolecular proton transfer (ESIPT) reaction in 2-(2'-hydroxy-4'-diethylaminophenyl)benzothiazole (HABT) have been investigated via steady-state and femtosecond fluorescence upconversion approaches. In cyclohexane, in contrast to the ultrafast rate of ESIPT for the parent 2-(2'-hydroxyphenyl)benzothiazole (>2.9+/-0.3 x 10(13) s(-1)), HABT undergoes a relatively slow rate (approximately 5.4+/-0.5 x 10(11) s(-1)) of ESIPT. In polar aprotic solvents competitive rate of proton transfer and rate of solvent relaxation were resolved in the early dynamics. After reaching the solvation equilibrium in the normal excited state (N(eq)*), ESIPT takes place with an appreciable barrier. The results also show N(eq)*(enol)<-->T(eq)*(keto) equilibrium, which shifts toward N(eq)* as the solvent polarity increases. Temperature-dependent relaxation dynamics further resolved a solvent-induced barrier of 2.12 kcal mol(-1) for the forward reaction in CH(2)Cl(2). The observed spectroscopy and dynamics are rationalized by a significant difference in dipole moment between N(eq)* and T(eq)*, while the dipolar vector for the enol form in the ground state (N) is in between that of N(eq)* and T(eq)*. Upon N-->N* Franck-Condon excitation, ESIPT is energetically favorable, and its rate is competitive with the solvation relaxation process. Upon reaching equilibrium configurations N(eq)* and T(eq)*, forward and/or backward ESIPT takes place with an appreciable solvent polarity induced barrier due to differences in polarization equilibrium between N(eq)* and T(eq)*.  相似文献   

10.
Kinetic results for the addition of OH? to [Mn(CO)3(η-C6H6)]+ (I) in water (eq. 1, X  OH) obey the expression kobskOH[OH?], and give a kOH value of 290 mol?1 dm3 s?1 at 20.0°C and ionic strength of 0.25 mol dm?3. The analogous reaction of NaCN with I in water fits the two-term expression kobs = kOH[OH?] + kCN[CN?], and leads to a kCN value of 0.8 mol?1 dm3 s?1 at 20.0°C and ionic strength of 0.25 mol dm?3. Interestingly, the related reaction (eq. 1, X  N3) is too rapid to follow by stopped-flow spectrophotometry, indicating the overall rate trend N3? » OH? » CN?. This unusual nucleophilicity order, unexpected on the basis of both basicity and polarizability, is similar to that previously observed for anion addition to free carbonium ions.  相似文献   

11.
The analytical suitability of zinc dithizonate foam for the collection and preconcentration of traces of mercury(II) has been examined. The effect of pH, plasticizer and chelating agent concentration on the collection rate of mercury has been critically investigated. The capacity for mercury(II) of a TBP-plasticized zinc dithizonate foam at pH 6 is 23.4 μeq g?1. The effect of plasticizer on the rate of recovery of mercury with sodium thiosulphate solution is also discussed. The preparation of zinc diethyldithiocarbonate foam is described. Traces of mercury(II) can also be collected rapidly and quantitavely by this foam.  相似文献   

12.
Abstract

34 new fluoro-λ-monophosphazenes and 37 new fluoro-1.3-diaza-2λ55-diphosphetidines have been prepared by applying the STAUDINGER reaction on δ3-phosphorus compounds (eq. (1)).  相似文献   

13.
The Stobbe condensation1 is a classical reaction of the mono-enolate, l, of diethyl succinate, generated in situ by action of alkoxide bases, with aldehydes or ketones (eq 1).  相似文献   

14.
A new inclusion complex, decaazonium tricyclo[28.2.2.213.16]tetratriacontane (L) decatetrafluoroborate decahydrate [(H10L)(BF4)1010H2O] (I), has been prepared and characterized by X-ray crystallography. Crystal data: a = 8.156(2), b = 29.125(6), c = 12.981(3) , = 96.76(3)°, space group P21/c, Z = 2, the final R = 0.0662 for 5376 reflections with I> 2 (I) (MoK radiation). The azamacrocycle of the complex is a 10-charged cation. The components of the supramolecular system in crystal are stabilized by the hydrogen bonds between the macrocycle cations, the water molecules, and the tetrafluoroborate anions. The water molecules make a solvation shell of the macrocycle forming an octahydrate. The tetrafluoroborate anions alternate with the water molecules stitched to the macrocycle, forming polymer chains in the directions parallel to the z axis of the crystal, and serve as bridges between the macrocycle cations related by the c plane. The set of hydrogen bonds including weak ones unite the components of the complex into a three-dimensional framework. All tetrafluoroborate anions in the structure are disordered.  相似文献   

15.
A one-pot synthesis of new polycyclic heterocyclic compounds was carried out via the condensation of 1-carbamoylmethyl-2,3,3-trimethyl-3H-indolium chloride with pyridine- and quinolinecarboxaldehydes. The heating of the aforementioned 3H-indolium salts with 1 eq. of pyridine-2, 3, and 4- or quinoline-4-carboxaldehyde in ethanol in the presence of piperidine as a catalyst provided 9a-[2-(pyridyl)ethenyl]- or 9a-[2-(quinolyl)ethenyl]-9,9a-dihydro-1H-imidazo[1,2-a]indol-2(3H)-one derivatives as the main products. However, reaction outcome was dramatically different for the analogous reactions in acetic acid. In this case, the heating of the chloride with 2 eq. of pyridine-2-carboxaldehyde afforded derivatives of 9a-[3-(pyridin-2-yl)indolizin-2-yl]-9,9a-dihydro-1H-imidazo[1,2-a]indol-2(3H)-one as the major product, while the use of 2 eq. of pyridine-3 and 4- or quinoline-4-carboxaldehyde led to the formation 2-heteroaryl-1-heteroarylmethyl-9H-pyrrolo[1,2-a]indole-3-carboxamides. Plausible pathways for the cyclization reactions are discussed. The structural assignments were based on 1H, 13C and 15N NMR spectroscopy, HRMS and single-crystal X-ray diffraction data.  相似文献   

16.
This report examines the effect of a decrease in solvent viscosity on the simulated folding behaviour of a β‐peptide heptamer in methanol. Simulations of the molecular dynamics of the heptamer H‐β3‐HVal‐β3‐HAla‐β3‐HLeu‐(S,S)‐β3‐HAla(αMe)‐β3‐HVal‐β3‐HAla‐β3‐HLeu‐OH in methanol, with an explicit representation of the methanol molecules, were performed for 80 ns at various solvent viscosities. The simulations indicate that at a solvent viscosity of one third of that of methanol, only the dynamic aspects of the folding process are altered, and that the rate of folding is increased. At a viscosity of one tenth of that of methanol, insufficient statistics are obtained within the 80 ns period. We suggest that 80 ns is an insufficient time to reach conformational equilibrium at very low viscosity because the dependence of the folding rate of a β‐peptide on solvent viscosity has two regimes; a result that was observed in another computational study for α‐peptides.  相似文献   

17.
The reaction o f the lithium enolate o f a-trimethylsilyl esters with aldehydes and ketones has been shown to be an excellent entry into a, B-unsaturated esters as a result of the facile elimination o f the B-oxidosilane intermediate produced in the initial step of the reaction. (eq. 1) More recently it was shown that the bromomagnesium enolate o f trimethylsilyl acetates a1 lowed for the highly stereoselective preparation of (E) a, B-unsaturated esters via the isolable B-hydroxy-a-silyl ester, which was treated with BF30Et2 to effect a trans elimination of the elements of trimethylsilanol5 (eq. 2) If this elimination were carried out with base, however, the stereoselectivity is much lower.  相似文献   

18.
Antonios Kolocouris 《Tetrahedron》2009,65(45):9428-9435
Dynamic NMR spectroscopy and ab initio correlated calculations revealed that the attachment of a spiroadamantane entity at the C-2 position of N-methylpyrrolidine or N-methylpiperidine induces a severe steric crowding around nitrogen, which changes the conformational space of the heterocycle resulting in: (a) the complete destabilization of the N-Me(eq) conformer in spiranic structures; in contrast the N-Me(eq) conformer corresponds to the global minimum in N-methylpyrrolidine or N-methylpiperidine. The spiroadamantane structure raises the energy of the equatorial conformer because of the severe van der Waals repulsion between the N-Me(eq) group and adamantane C-H bonds. (b) The interconversion between the only populated enantiomeric N-Me(ax) conformers ax→[eq]→ax′; the interconversion eq→ax between N-Me(eq) and N-Me(ax) conformers, which are both populated, is observed in N-methylpyrrolidine or N-methylpiperidine. (c) The raising of ring and nitrogen inversion barriers ax→ts by ∼4-6 kcal mol−1. The dynamic NMR study provides evidence that the most important process required for the enantiomerization between the axial N-Me conformers in spiropiperidine 4 and spiropyrrolidine 5 are different, i.e., a nitrogen inversion in 5 (9.10 kcal mol−1) and a ring inversion in 4 (15.2 kcal mol−1). While an enantiomerization interconverts N-Me axial conformers in spiropiperidine 5 and spiropyrrolidine 4, substitution of the pyrrolidine ring of 5 with a C-Me group effects a diastereomerization between two N-Me axial conformers and reduces effectively the nitrogen inversion barrier according to the protonation experiments and the calculations. In general, all the calculations levels used, i.e., the MM3, B3LYP/6-31+G∗∗ and MP2/6-311++G∗∗//B3LYP/6-31+G∗∗, predict correctly the different stability of the local minima; however only MP2/6-311++G∗∗//B3LYP/6-31+G∗∗ was found to be reliable for the calculation of the nitrogen inversion barriers.  相似文献   

19.
Abstract

We have recently discovered direct and high yielding routes to phosphoranylidenephosphines ArP=PMe3 (la, Ar = Dmp; lb, Ar = Mes*, eq. l).1  相似文献   

20.
A new class of bidentate phosphoramidite ligands, based on a spiroketal backbone, has been developed for the rhodium‐catalyzed hydroformylation reactions. A range of short‐ and long‐chain olefins, were found amenable to the protocol, affording high catalytic activity and excellent regioselectivity for the linear aldehydes. Under the optimized reaction conditions, a turnover number (TON) of up to 2.3×104 and linear to branched ratio (l/b) of up to 174.4 were obtained in the RhI‐catalyzed hydroformylation of terminal olefins. Remarkably, the catalysts were also found to be efficient in the isomerization–hydroformylation of some internal olefins, to regioselectively afford the linear aldehydes with TON values of up to 2.0×104 and l/b ratios in the range of 23.4–30.6. X‐ray crystallographic analysis revealed the cis coordination of the ligand in the precatalyst [Rh( 3 d )(acac)], whereas NMR and IR studies on the catalytically active hydride complex [HRh(CO)2( 3 d )] suggested an eq–eq coordination of the ligand in the species.  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