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.
A series of chiral non-racemic N1,N10-ethylene bridged flavinium salts 4 was prepared using enantiomerically pure 2-substituted 2-aminoethanols (R?=?isopropyl, phenyl, benzyl, 4-methoxybenzyl, 4-benzyloxybenzyl) derived from amino acids as the sole source of chirality. The flavinium salts were shown to form 10a-hydroperoxy- and 10a-methoxy-adducts with moderate to high diastereoselectivity depending on the ethylene bridge substituent originating from the starting amino acid. High diastereoselectivities (dr values from 80:20 to >95:5) were observed for flavinium salts bearing benzyl substituents attached to the ethylene bridge. The benzyl group preferred the face-to-face (syn) orientation relative to the flavinium unit; thereby effectively preventing nucleophilic attack from one side. This conformation was found to be the most stable according to the DFT calculations. Consequently, the presence of benzyl groups causes intermolecular fluorescence quenching resulting in a significant decrease in the fluorescence quantum yield from 11% for 4a bearing an isopropyl substituent to 0.3% for 4c containing a benzyl group and to a value lower than 0.1% for the benzyloxybenzyl derivative 4e.  相似文献   

2.
The reaction of P4S10 (1) with N,N′-diphenylurea (PhNH)2CO (2) results in new heterocyclic compounds: the pyridinium salt of 1,3-diphenyl-2-sulfido-2-thioxo-1,3-diaza-2λ5-phosphetidine (3) (with a P–N–C–N cycle) and the pyridinium salt of 1,4-diphenyl-2,5-disulfido-2,5-dithioxo-1,4-dithiadiaza-2λ5,5λ5-diphosphinane (4), containing the (P–S–N)2 cycle and the cyclic thiophosphates [pyH]2[P2S8] (5), [pyH]2[P2S7] (6) and [pyH]3[P3S9] (7). A similar reaction, but carried out with N,N′-diphenylthiourea (PhNH)2CS (8), leads to the formation of 4 and 6. pyPS2Cl (9), used as an alternative starting material, also yields compounds 3, 4, 5, and further [pyH][PS2Cl2] (10) and S8 after reaction with 2. Compound 3 reacts with Pd(CH3COO)2, with the formation of the complex [Pd(Ph2N2COPS2)2] (11). The crystal structures of 3 and 7 were determined by single-crystal X-ray diffraction.  相似文献   

3.
Two novel N,N′-dialkylimidazolium thiocyanate-cadmium complexes [(R2Im)2][Cd2(SCN)6] for R=Me (3), and cyclohexyl (4) have been synthesized and characterized by single-crystal X-ray diffraction. Compound 3 crystallizes in the monoclinic unit cell dimensions of 17.468(3), 7.7273(12), 10.6750(16) Å, 104.833(2)°, and space group C2 with two [(Me2Im)2] [Cd2(SCN)6] per unit cell. The two cadmium atoms in 3 are octahedrally coordinated in 4N2S and 2N4S coordination environment, and linking into one-dimensional zigzag chains. Compound 4 belongs to the monoclinic space group Cc with unit cell of dimensions 13.3049(12), 17.5550(16), 20.8012(19) Å, 101.494(2)°, and four [(Cy2Im)2][Cd2(SCN)6]·C3H6O per unit cell. The cadmium atoms in 4 are all 3N3S hexa-coordinated with six bridging SCN ions in an fac configuration and form infinite zigzag polymeric chains. The infinite chains in 3 form an approximate hexagonal array, making triangular channels which are occupied by N,N′-dimethylimidazolium ions, whereas the chains in 4 form layered structure, and the layers are stacked perpendicularly with respect to the orientation of the infinite anionic chains alternatively. N,N′-dicyclohexylimidazolium cations and solvent molecules fill in between layers.  相似文献   

4.
Treatment of 4-nitrophenyl chloroformate with alkylammonium hydrochloride salts and solid anhydrous Na2CO3 in either CH2Cl2 or CH3CN gave 4-nitrophenyl N-methylcarbamate and other N-alkylcarbamate analogues in excellent yields. Of particular interest is the observation that 4-nitrophenyl N-methylcarbamate, a safer alternative to the highly toxic methyl isocyanate, is obtained in quantitative yield (?95% pure as determined by 1H NMR) after simple filtration and solvent evaporation.  相似文献   

5.
The reaction of primary amines RNH2 (R: Me, Et, iPr, tBu and Ph) with 1,2-dibromoethane gave N,N′-disubstituted ethylenediamines R-NH-CH2CH2-NH-R (1) in yields ranging from 10% (1a; R=Me) to 70% (1d, R=tBu; 1e, R=Ph). Piperazines and N-substituted polyethyleneimines were identified (1H NMR, 13C NMR and EI-MS) as side products of the reaction and isolated by fractional distillation. The piperazines 2 are formed in yields of 3-10% and can be separated from the diamines 1 in all cases, except for R=Me and Ph. The polyamine homologues RNH-[CH2CH2NR]n-H (3-5) were isolated in yields ranging from 0.1% (n=4, R=iPr) to 14% (n=2, R=iPr). The yields of 1 increase with the size of the substituent R, no obvious trend exists for the yields of the side products.  相似文献   

6.
Imino- and bis-iminopyridine ligands bearing N-ter-butylhydroxy groups were synthesized by the condensation reaction between a N-ter-butylhydroxy substituted aniline and 2-formylpyridine, 2-acetylpyridine and 2,6-bis-acetylpyridine. The N-ter-butylaminoxy radicals obtained after oxidation using PbO2 or Ag2O were studied by EPR spectroscopy in solution. Indeed, complete decomposition was observed during isolation of these radical derivatives. The N-ter-butylhydroxy substituted ligands obtained were complexed with Mn2+, Ni2+, Zn2+ and Gd3+ salts; the obtained complexes were characterized and oxidized to give the aminoxy analogs, which were studied using IR, UV and EPR spectroscopies.  相似文献   

7.
Protonation of N-alkyl-2-aza[3]ferrocenophanes by HCl and NH4PF6 affords hexafluorophosphate salts having a trialkylammonium group. Structures of the protonated and unprotonated N-(p-methylbezyl)-2-aza[3]ferrocenophanes were determined by X-ray crystallography. Variable temperature 13C{1H} NMR spectra of the N-protonated N-hexyl-2-aza[3]ferrocenophane revealed inversion of the nitrogen of the 2-aza[3]ferrocenophane on the NMR time scale probably via partial deprotonation of the nitrogen atom. Cyclic voltammograms of the N-protonated compounds exhibited reversible redox peaks at higher potentials than those of the corresponding neutral ferrocenophanes.  相似文献   

8.
The homogeneous polymerization of 3-(N-2-methacryloyloxyethyl-N,N-dimethyl)ammonatopropanesulfonate (MDAPS) with potassium peroxydisulfate (KPS) was kinetically in situ investigated in water by means of FT-near IR spectroscopy. The overall activation energy of the polymerization was calculated to be 16.0 kcal/mol. The initial polymerization rate (Rp) at 40 °C was expressed by Rp=k[KPS]0.65[MDAPS]1.0. The presence of alkaline metal salts was observed to accelerate the polymerization. The order of acceleration at 40 °C was CsCl > KCl > NaCl > LiCl when the chloride salts were used. NaCl showed higher acceleration effect than NaF. NaBr and NaI exhibited retardation and inhibition effect, respectively, because of reduction of KPS and its primary radical with bromide and iodide ions. The polymerization of MDAPS with KPS in water in the presence of NaCl at 2.0 mol/l gave Rp=k[KPS]0.70[MDAPS]1.4 at 40 °C. The overall activation energy of the polymerization in the presence of NaCl was estimated to be 11.6 kcal/mol being considerably lower value compared with that in its absence. The syndiotacticity of poly(MDAPS) tended to increase with rising temperature and decrease in the presence of NaCl.  相似文献   

9.
The valence stability of tin in its complexes with 1-hydroxyethylene-diphosphonate (HEDP) and with N,N′,N′-trimethylenephosphonate-polyethyleneimine (PEI-MP) was investigated. With particular interest in the possible interconversion between Sn2+ and Sn4+, the complexes were monitored with the aid of 31P NMR spectroscopy. The extent of complex formation with both ligands was evaluated for systems with tin in their respective oxidation states. The Sn2+-complexes underwent initial, but limited oxidation upon preparation, and beyond which were rather stable, irrespective of pH or time. Both Sn2+- and Sn4+-complexes were found to exist in solution without change. Oxidation of Sn2+ was achieved by addition of hydrogen-peroxide and was partially reversed by the addition of glutathione (GSH). The amount of H2O2 needed for complete oxidation of the Sn2+- into Sn4+-complexes was determined for both ligands, as well as the time taken for that oxidation.  相似文献   

10.
Na2[(VIVO)2(ttha)]·8 H2O (ttha = triethylenetetraamine–N,N,N′,N″,N′″,N′″–hexaacetate ion), prepared by treating [VO(H2O)5][(VO)2(ttha)]·4 H2O with Na6(ttha), has been characterized by single crystal X-ray diffraction, infrared spectroscopy, UV–Vis absorption spectroscopy, electron spin resonance spectroscopy, and modeled by density functional theory (DFT). The X-ray structure revealed a distorted octahedral geometry around each vanadium center. The electronic absorption spectrum of [(VO)2(ttha)]2− (aq) features absorptions at ca. 200 nm (ε > 13900 L mol−1 cm−1), 255 nm (ε = 3480 L mol−1 cm−1), 586 nm (ε = 33 L mol−1 cm−1), and 770 nm (ε = 38 L mol−1 cm−1). The time-dependent density functional theory (TDDFT) calculated electronic absorption spectrum was remarkably similar to the actual spectrum, and TDDFT predicts absorption peaks at 297, 330, 458, 656, and 798 nm. TDDFT assigned the peak at 798 nm to be the α spin HOMO → LUMO transition. Hence, the peak at 770 nm in the actual spectrum is most likely the α spin HOMO → LUMO transition. Moreover, the TDDFT calculations revealed that the α spin HOMO and LUMO are partly comprised of d orbitals on both vanadium centers, and the first derivative electron spin resonance spectrum also suggests that the two unpaired electrons in [(VO)2(ttha)]2− are localized near the vanadium centers.  相似文献   

11.
The synthesis and characterization of binuclear ruthenium complexes [{(η6-C6H6)Ru}2(μ-bsh)2] (1), [{(η6-C10H14)Ru}2(μ-bsh)2] (2), [{(η6-C6Me6)Ru}2(μ-bsh)2] (3), and rhodium complex [{(η5-C5Me5)RhCl}2(μ-bsh)] (4) (bsh=N,N-bis(salicylidine)-hydrazine dianion) are reported. The complexes have been fully characterized by analytical and spectral techniques and unusual coordination mode of the ligand H2bsh has been confirmed by single crystal X-ray analysis of the complex 2. Structural data revealed extensive inter- and intra-molecular C-H?O and C-H?π interactions and involvement of methyl and isopropyl hydrogen from the p-cymene in hydrogen bonding.  相似文献   

12.
New polar vanadium tellurite enantiomers have been synthesized under mild hydrothermal conditions through the use of sodium metavanadate, sodium tellurite and enantiomerically pure sources of either R-3-aminioquinuclidine or S-3-aminioquinuclidine. [R-C7H16N2][V2Te2O10] and [S-C7H16N2][V2Te2O10] contain [V2Te2O10]n2n layers constructed from [(VO2)2O(TeO4)2] monomers. Steric effects associated with the hydrogen-bonding network between the [V2Te2O10]n2n layers and [C7H16N2]2+ result in polar structures and crystallization in the space group P21 (no. 4). Electron localization functions were calculated to visualize the tellurite stereoactive lone pairs. Both iterative and non-iterative Hirshfeld techniques were evaluated as means to determine atomic partial charges, with iterative Hirshfeld charges more accurately representing charge distributions in the reported enantiomers. These charges were used to calculate both component and net dipole moments. [R-C7H16N2][V2Te2O10] and [S-C7H16N2][V2Te2O10] exhibit dipole moments of 17.37 and 16.62D, respectively. [R-C7H16N2][V2Te2O10] and [S-C7H16N2][V2Te2O10] both display type 1 phase-matching capabilities and exhibit second harmonic generation activities of ∼50×α-SiO2.  相似文献   

13.
New tetraimidazolinium and tetrabenzimidazolium salts have been prepared. Upon reaction with tBuOK, they generate carbene ligands, which were associated in situ to [RuCp*(MeCN)3]PF6 to produce new ruthenium catalysts that are active for the substitution of allylic substrates by amines, phenols, and carbonucleophiles. The influence of the N-heterocyclic core as well as that of the N-substitutents at the periphery of the salts, on reactivity and regioselectivity have been examined.  相似文献   

14.
New Schiff-base copper and cobalt complexes, [Cu(L1)], [Cu(L2)] and [Co(L1)], [Co(L2)] (where L1 = N-N′-bis(3,5-di-tert-butylsalicylaldimine)-1,4-cyclohexane bis(methylamine) and L2 = N-N′-bis(3,5-di-tert-butylsalicylaldimine)-1,8-diamino-3,6-dioxaoctane), were synthesized and characterized using elemental analysis, IR spectra, UV–Vis spectra, magnetic susceptibility measurements, 1H and 13C NMR spectroscopy, thermal analysis and molar conductance (ΛM). Their electro-spectrochemical properties were investigated using cyclic voltammetric (CV) and thin-layer spectroelectrochemical techniques in a dichloromethane solution (CH2Cl2). The CV of [Cu(L2)] showed a lower oxidation potential than that of [Cu(L1)] under the same experimental conditions. The oxidation wave (II) of [Cu(L2)] was accompanied by an EC process (II′), which was not observed for [Cu(L1)]. Also, [Cu(L2)] exhibited a reduction process, but [Cu(L1)] did not. These results indicate that the Cu(II) ion in [Cu(L2)] is coordinated by N2O4 donor sites while [Cu(L1)] presents a square-planar structure with N2O2 donor sites. Both oxidation processes for [Co(L1)] and [Co(L2)] are based on the cobalt center, and they are assigned to Co(II)/Co(III) couples. The spectroelectrochemical results indicate that the oxidized species of [Cu(L2)] is similar to that of [Cu(L1)], the only difference being that the absorption bands of the oxidized species for [Cu(L2)] shift to lower energy compared with those of [Cu(L1)] because of their different coordination environment. The geometry of [Cu(L2)] changed into square-planar after the complex was totally oxidized and the neutral complex was only recovered following the EC process, as observed from the CV of [Cu(L2)]. For the two cobalt complexes, the bands corresponding to the π → πtransitions disappeared and new bands with small red shifts and of lower intensity were observed during the oxidation process. These new bands are attributed to the LMCT transition as observed in the case of the oxidation processes of the cobalt complexes.  相似文献   

15.
The C3v tris-methoxy calix[6]arene was selectively mono-alkylated by dibromoethane yielding a key intermediate for the design of disymmetrically O-substituted calix[6]arenes. Indeed, subsequent reactions with various functional groups afforded novel calix[6]arene-based biomimetic ligands that present a mixed donor N2S or N3CO2 environment in an efficient way.  相似文献   

16.
The salts, [Ph2B{OCH2CH2N(Me)(CH2)n}2][Ph4B3O3] (n = 4, 5), were prepared in moderate yields in MeOH solution from reaction of Ph2BOBPh2 with [N(CH2)n(Me)(CH2CH2OH)][OH] and PhB(OH)2 in a 1:2:4 ratio. The reactions also lead to Ph3B3O3. Both salts were characterized by NMR (1H, 13C, 11B) IR, and single-crystal XRD studies. The salts are comprised of cationic monoborates (zwitterionic, 2N+ and 1B) and tetraphenylboroxinate anions.  相似文献   

17.
Reactions of tetrafluoroethylene, chlorotrifluoroethylene and 1,2-dichlorodifluoroethylene with N-potassium salts of imidazole, 2-methylbenzimidazole, 3,5-dimethylpyrazole, 1,2,4-triazole, and benzotriazole lead to the formation of the corresponding N-(1,1,2,2-tetrafluoroethyl), N-(2-chloro-1,1,2-trifluoroethyl), and N-(2-chloro-1,2-difluorovinyl) azoles. Treatment of N-(2-chloro-1,2-difluorovinyl) and N-(2-chloro-1,1,2-trifluoroethyl) derivatives of azoles with tetramethylammonium fluoride is a useful synthetic method for the preparation of heterocycles with 1,2,2,2-tetrafluoroethyl group attached to nitrogen.  相似文献   

18.
A Pd2dba3/P(i-BuNCH2CH2)3N catalyzed one-pot synthesis of unsymmetrically substituted trans-4-N,N-diarylaminostilbenes and both symmetrically and unsymmetrically substituted N,N-diarylaminostyrene derivatives is reported. The procedure involves two or more palladium catalyzed sequential coupling reactions (an amination and an inter-molecular Heck reaction) in one-pot using the same catalyst system with two different aryl halides, including aryl chlorides and hetero aryl halides as the coupling partners.  相似文献   

19.
We found a suitable condition for the effective alkynylation of N-tosylimines with aryl acetylenes. The reaction of N-tosylimines and aryl acetylenes in the presence of ZnBr2 and DIEA (N,N-diisopropylethylamine) in CH3CN afforded the desired N-tosyl propargylamines in moderate to good yields.  相似文献   

20.
A simple and highly efficient protocol for the AcOH-catalyzed three-component reaction among nitroalkenes, difluoroethylamine and tert-butyl nitrite through cascade aza-Michael addition/N-nitrosation has been developed. A range of CF2HCH2-containing N-nitrosoamines were obtained in good to excellent yields. This approach features easily available cheap materials, without using additional organic solvents, simple operation under room temperature, gram scalable preparation and functionally diverse products.  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