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 46 毫秒
1.
Synthesis of Tris(diorganylphosphino)phosphines with Substituents of Low Steric Hindrance The reaction of MIP(SiMe3)2 (MI = Li, Na, K) with diorganylchlorophosphines forms tris(diorganylphosphino)phosphines P(PR2)3 (R = Cy, Ph, i-Pr, n-Pr, Et, Me). In all cases, the stepwise formation reaction proceeds via di- and triphosphines resulting in iso-tetraphosphines which are stable in solution. Only the compounds with bulky substituents (R = Cy, Ph) can be isolated. By using quantumchemical procedures it is possible to get information about the pyramidal structure and the reaction behaviour. In agreement with these results orbitally controlled reactions occur where nucleophiles attack the central phosphorus atom.  相似文献   

2.
The Morita? Baylis? Hillman (MBH) reactions of (4S,5R,7R,8R)‐ and (4R,5R,7R,8R)‐4‐hydroxy‐7,8‐dimethoxy‐7,8‐dimethyl‐6,9‐dioxaspiro[4.5]dec‐2‐en‐1‐ones ( 2 and 3 , resp.) with aldehydes using various catalysts were studied. A combination of Bu3P/phenol in THF was found being optimum conditions giving the corresponding MBH adducts with high diastereoisomeric ratios. After separation, each stereomerically pure isomer of the MBH adducts was subjected to hydrolysis employing 1% aq. CF3COOH (TFA) in a water bath of an ultrasonic cleaner to afford the corresponding polyhydroxylated cyclopentenones in good yields.  相似文献   

3.
He‐Rng Zeng 《中国化学》2002,20(12):1546-1551
The photoinduced electron‐transfer reaction of N, N, N', N'‐tetra‐(p‐methylphenyl)‐4,4'‐diamino‐1,1'‐diphenyl ether (TPDAE) and fullerenes (C60/C70) by nanosecond laser flash photolysis occurred in benzonitrile. Transient absorption spectral measurements were carried out during 532 nm laser flash photolysis of a mixture of the fullerenes (C60/C70) and TPDAE. The electron transfer from the TPDAE to excited triplet state of the fullerenes (C60/C70) quantum yields and rate constants of electron transfer from TPDAE to excited triplet state of fullerenes (C60/C70) in benzonitrile have been evaluated by observing the transient absorption bands in the near‐IR region where the excited triplet state, radical anion of fullerenes (C60/C70) and radical cations of TPDAE are expected to appear.  相似文献   

4.
Through photocatalysed regiospecific and stereoselective additions of cycloamines to 5‐(R)‐(l)‐menthyloxy‐2 (5H)‐furanone (3), chiral 5‐(R)‐(l)‐menthyloxy‐4‐cycloaminobutyrolactones were synthesized. In the new asymmetric photoaddition of compound 3, the N‐methyl cyclic amines (4) gave novel chiral C? C photoadducts (5) in 24–50% isolated yields with d. e. ≥ 98%. However, the secondary cyclic amines (6) afforded optically active N? C photoadducts (7) in 34–58% isolated yields with d. e. ≥ 98% under the same condition. All the synthesized optically active compounds were identified on the basis of their analytical data and spectroscopic data, such as [α]58920, IR, 1H NMR, 13C NMR, MS and elementary analysis. The photosynthesis of chiral butyrolactones and its mechanism were discussed in detail.  相似文献   

5.
OH+ is an extraordinarily strong oxidant. Complexed forms (L? OH+), such as H2OOH+, H3NOH+, or iron–porphyrin‐OH+ are the anticipated oxidants in many chemical reactions. While these molecules are typically not stable in solution, their isolation can be achieved in the gas phase. We report a systematic survey of the influence on L on the reactivity of L? OH+ towards alkanes and halogenated alkanes, showing the tremendous influence of L on the reactivity of L? OH+. With the help of with quantum chemical calculations, detailed mechanistic insights on these very general reactions are gained. The gas‐phase pseudo‐first‐order reaction rates of H2OOH+, H3NOH+, and protonated 4‐picoline‐N‐oxide towards isobutane and different halogenated alkanes CnH2n+1Cl (n=1–4), HCF3, CF4, and CF2Cl2 have been determined by means of Fourier transform ion cyclotron resonance meaurements. Reaction rates for H2OOH+ are generally fast (7.2×10?10–3.0×10?9 cm3 mol?1 s?1) and only in the cases HCF3 and CF4 no reactivity is observed. In contrast to this H3NOH+ only reacts with tC4H9Cl (kobs=9.2×10?10), while 4‐CH3‐C5H4N‐OH+ is completely unreactive. While H2OOH+ oxidizes alkanes by an initial hydride abstraction upon formation of a carbocation, it reacts with halogenated alkanes at the chlorine atom. Two mechanistic scenarios, namely oxidation at the halogen atom or proton transfer are found. Accurate proton affinities for HOOH, NH2OH, a series of alkanes CnH2n+2 (n=1–4), and halogenated alkanes CnH2n+1Cl (n=1–4), HCF3, CF4, and CF2Cl2, were calculated by using the G3 method and are in excellent agreement with experimental values, where available. The G3 enthalpies of reaction are also consistent with the observed products. The tendency for oxidation of alkanes by hydride abstraction is expressed in terms of G3 hydride affinities of the corresponding cationic products CnH2n+1+ (n=1–4) and CnH2nCl+ (n=1–4). The hypersurface for the reaction of H2OOH+ with CH3Cl and C2H5Cl was calculated at the B3 LYP, MP2, and G3m* level, underlining the three mechanistic scenarios in which the reaction is either induced by oxidation at the hydrogen or the halogen atom, or by proton transfer.  相似文献   

6.
The solubility, diffusivity, and permeability of ethylbenzene in poly(1‐trimethylsilyl‐1‐propyne) (PTMSP) at 35, 45 and 55 °C were determined using kinetic gravimetric sorption and pure gas permeation methods. Ethylbenzene solubility in PTMSP was well described by the generalized dual‐mode model with χ = 0.39 ± 0.02, b = 15 ± 1, and CH = 45 ± 4 cm3 (STP)/cm3 PTMSP at 35 °C. Ethylbenzene solubility increased with decreasing temperature; the enthalpy of sorption at infinite dilution was −40 ± 7 kJ/mol and was essentially equal to the enthalpy change upon condensation of pure ethylbenzene. The diffusion coefficient of ethylbenzene in PTMSP decreased with increasing concentration and decreasing temperature. Activation energies of diffusion were very low at infinite dilution and increased with increasing concentration to a maximum value of 50 ± 10 kJ/mol at the highest concentration explored. PTMSP permeability to ethylbenzene decreased with increasing concentration. The permeability estimated from solubility and diffusivity data obtained by kinetic gravimetric sorption was in good agreement with permeability determined from direct permeation experiments. Permeability after exposure to a high ethylbenzene partial pressure was significantly higher than that observed before the sample was exposed to a higher partial pressure of ethylbenzene. Nitrogen permeability coefficients were also determined from pure gas experiments. Nitrogen and ethylbenzene permeability coefficients increased with decreasing temperature, and infinite dilution activation energies of permeation for N2 and ethylbenzene were −5.5 ± 0.5 kJ/mol and −74 ± 11 kJ/mol, respectively. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 1078–1089, 2000  相似文献   

7.
Poly(5,6‐difluoro‐2,1,3‐benzothiadiazole‐alt‐9,9‐dioctylfluorene) was successfully synthesized via direct arylation polycondensation of 5,6‐difluoro‐2,1,3‐benzothiadiazole and 2,7‐dibromo‐9,9‐dioctylfluorene. The reaction conditions were optimized, and a polymer with number‐average molecular weight (Mn) of 41,000 was obtained by using Pd(OAc)2, PtBu2Me‐HBF4, pivalic acid, K2CO3, and toluene as catalyst, ligand, additive, base, and solvent, respectively. The polycondensation was also performed with 5,6‐dioctyloxy‐2,1,3‐benzothiadiazole or 2,1,3‐benzothiadiazole as the comonomer, and the results indicate that the introduction of electron‐withdrawing fluorine atoms at the ortho‐positions to the C? H bonds is essential for the reactivity of the direct arylation. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 2367–2374  相似文献   

8.
A green and improved method for the synthesis of β‐amido ketones through one‐pot multi‐component reaction of an enolizable ketone, aryl aldehyde, acetonitrile or benzonitrile in the presence of trimethylsilyl chloride using 3‐methyl‐1‐(4‐sulfonic acid)butylimidazolium hydrogen sulfate [(CH2)4SO3HMIM][HSO4], a Br?nsted‐acidic ionic liquid, as an effective and recyclable catalyst is described. The present methodology offers several advantages such as simple procedure with an easy work‐up, relatively short reaction time, and good to excellent yields.  相似文献   

9.
The synthesis of asymmetrically substituted 2,2′:6′,2′′‐terpyridines is reported. First, palladium‐catalyzed C? H arylation of pyridine N‐oxides with substituted bromopyridines gave 2,2′‐bipyridine N‐oxides, which were further arylated in a second step to form 2,2′:6′,2′′‐terpyridine N‐oxides. Yields of up to 77 % were obtained with N‐oxides bearing an electron‐withdrawing ethoxycarbonyl substituent in the 4‐position. Pd(OAc)2 with either P(tBu)3 or P(o‐tolyl)3 was used as the catalyst. Cyclometalated complexes derived from Pd(OAc)2 and these phosphines were also effective. K3PO4 as the base gave better results than K2CO3. Subsequent deoxygenation with H2 and Pd/C as the catalyst gave the asymmetrically substituted 2,2′:6′,2′′‐terpyridines in near quantitative yield. This reaction sequence significantly reduces the number of steps required in comparison with known cross‐coupling methods and therefore allows convenient and scalable access to substituted terpyridines.  相似文献   

10.
A new class of bidentate, aza‐based phosphinic amide ligands of the type RN(H)P(?O)(2‐py)2 (2‐py = 2‐pyridyl) was synthesized within minutes via a one‐pot process including Staudinger reaction of an organic azide (RN3) with 2‐pyridylphosphines, followed by partial, unprecedented hydrolysis under loss of one aromatic substituent. The structure of the unusual‐hydrolysis product H2C?CH(CH2)9N(H)P(?O)(2‐py)2 ( 5a ) was characterized by IR, 1H‐ and 31P‐NMR, as well as by X‐ray crystal‐structure analysis (Figure). The tetrahedral P‐atom was found to be surrounded by a trigonal‐pyramidal arrangement of the substituents. To gain insight into the formation of these novel phosphinic amides, a series of intermediate iminophosphoranes, H2C?CH(CH2)9N?P(Ar)n(2‐py)3 ? n (n = 0–3), compounds 1a – 1f , were synthesized, and their hydrolyses were studied. All tested compounds followed the classical hydrolysis route of P?N cleavage under acidic conditions. Sequential hydrolysis to 5a – 5d only occurred under either basic conditions or in wet MeCN as solvent. Notably, H2C?CH(CH2)9N?P(C6H5)(4‐MeO‐2‐py)2 ( 1c ) was hydrolyzed at a much slower rate compared to its analogue 1b lacking the MeO group. On the contrary, the halogenated compounds H2C?CH(CH2)9N?P(4‐X‐C6H4)3 ( 1f,g ) (X = F, Cl) were hydrolyzed at a notably faster rate relative to the non‐halogenated congener 1e (X = H).  相似文献   

11.
In this study, a silicic acid and tetra isopropyl ortho titanate ceramic precursor and a metallocene polyethylene‐octene elastomer (POE) or acrylic acid grafted metallocene polyethylene‐octene elastomer (POE‐g‐AA) were used in the preparation of hybrids (POE/SiO2? TiO2 and POE‐g‐AA/SiO2? TiO2) using an in situ sol‐gel process, with a view to identifying a hybrid with improved thermal and mechanical properties. Hybrids were characterized using Fourier transform infrared spectroscopy, 29Si solid‐state nuclear magnetic resonance (NMR), X‐ray diffraction, differential scanning calorimetry, thermogravimetry analysis, dynamic mechanical thermal analysis, and Instron mechanical testing. Properties of the POE‐g‐AA/SiO2? TiO2 hybrid were superior to those of the POE/SiO2? TiO2 hybrid. This was because the carboxylic acid groups of acrylic acid acted as coordination sites for the silica‐titania phase to allow the formation of stronger chemical bonds. 29Si solid‐state NMR showed that Si atoms coordinated around SiO4 units were predominantly Q3 and Q4. The 10 wt % SiO2? TiO2 hybrids gave the maximum values of tensile strength and glass transition temperature in both POE/SiO2? TiO2 and POE‐g‐AA/SiO2? TiO2. It is proposed that above this wt %, excess SiO2? TiO2 particles caused separation between the organic and inorganic phases. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 1690–1701, 2005  相似文献   

12.
Five novel pyrazole‐coupled glucosides, 1,5‐diaryl‐1H‐pyrazol‐3‐yl 2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐glucopyranosides 5a – 5e , were synthesized by the phase‐transfer catalytic reaction of 1,5‐diaryl‐1H‐pyrazol‐3‐ols 4a – 4e with acetobromo‐α‐D ‐glucose in H2O/CHCl3 under alkaline conditions, using Bu4N+Br? as catalyst. Then, glucosides 5a – 5c were deacetylated in a solution of Na2CO3/MeOH to yield the 1,5‐diaryl‐3‐(β‐D ‐glucopyranosyloxy)‐1H‐pyrazoles 6a – 6c . Their structures were characterized by 1H,1H‐COSY, 1H‐, 13C‐, and 19F‐NMR spectroscopy, as well as elemental analysis. The structures of 5d and 6c were also determined by single‐crystal X‐ray diffraction analysis. A preliminary in vitro bioassay indicated that compounds 4e and 5d exhibited excellent‐to‐medium fungicidal activity against Sclerotinia sclerotiorum at the dosage of 10 μg/ml.  相似文献   

13.
Oxidative addition of aryl bromides to 12‐electron [Rh(PiBu3)2][BArF4] (ArF=3,5‐(CF3)2C6H3) forms a variety of products. With p‐tolyl bromides, RhIII dimeric complexes result [Rh(PiBu3)2(o/p‐MeC6H4)(μ‐Br)]2[BArF4]2. Similarly, reaction with p‐ClC6H4Br gives [Rh(PiBu3)2(p‐ClC6H4)(μ‐Br)]2[BArF4]2. In contrast, the use of o‐BrC6H4Me leads to a product in which toluene has been eliminated and an isobutyl phosphine has undergone C? H activation: [Rh{PiBu2(CH2CHCH3C H2)}(PiBu3)(μ‐Br)]2[BArF4]2. Trapping experiments with ortho‐bromo anisole or ortho‐bromo thioanisole indicate that a possible intermediate for this process is a low‐coordinate RhIII complex that then undergoes C? H activation. The anisole and thioanisole complexes have been isolated and their structures show OMe or SMe interactions with the metal centre alongside supporting agostic interactions, [Rh(PiBu3)2(C6H4O Me)Br][BArF4] (the solid‐state structure of the 5‐methyl substituted analogue is reported) and [Rh(PiBu3)2(C6H4S Me)Br][BArF4]. The anisole‐derived complex proceeds to give [Rh{PiBu2(CH2CHCH3C H2)}(PiBu3)(μ‐Br)]2[BArF4]2, whereas the thioanisole complex is unreactive. The isolation of [Rh(PiBu3)2(C6H4O Me)Br][BArF4] and its onward reactivity to give the products of C? H activation and aryl elimination suggest that it is implicated on the pathway of a σ‐bond metathesis reaction, a hypothesis strengthened by DFT calculations. Calculations also suggest that C? H bond cleavage through phosphine‐assisted deprotonation of a non‐agostic bond is also competitive, although the subsequent protonation of the aryl ligand is too high in energy to account for product formation. C? H activation through oxidative addition is also ruled out on the basis of these calculations. These new complexes have been characterised by solution NMR/ESIMS techniques and in the solid‐state by X‐ray crystallography.  相似文献   

14.
Preparation of Dithiatetrazocine and Secondary Reactions Li[PhCN2(SiMe3)2] ( 1 ) or PhCN2(SiMe3)3 ( 3 ) react with SCl2 to give in good yields the dithiatetrazocine PhC(NSN)2CPh ( 2 ). By analogy, p-MeC6H4C(NSN)2CC6H4Me-p ( 7 ), p-NO2C6H4C(NSN)2-CC6H4NO2-p ( 8 ), and p-CF3C6H4C(NSN)2CC6H4CF3-p ( 9 ) are obtained from the reaction of p-MeC6H4CN2(SiMe3)3 ( 4 ), Li[p-NO2-C6H4CN2(SiMe3)2] ( 5 ), und Li[p-CF3C6H4CN2(SiMe3)2] ( 6 ) with SCl2. Reaction of 2 /LiCl with AgAsF6 in liquid SO2 leads to [PhCN2S2]+[AsF6] ( 10 ) and 3[PhCN2S2]+2[AsF6]Cl ( 11 ). The structures of 10 and 11 are confirmed by X-ray analyses.  相似文献   

15.
Trimethylsilyl ethers were converted to their corresponding alcohols in the presence Nafion-H® and wet SiO 2 with good-to-excellent yields under mild and heterogeneous conditions.  相似文献   

16.
Treatment of {[(benzyloxy)carbonyl]amino}‐substituted sulfones 1 with 2‐[(trimethylsilyl)oxy]furan ( 2 ) in the presence of InCl3 as a catalyst at room temperature produced the γ‐butenolactone derivatives 3 and 4 containing a protected amino group (Scheme 1). The products were formed in high yields (81–92%) within 3–10 h favoring the anti‐isomer 3 .  相似文献   

17.
Bifunctional derivatives (XMe2Si)2Si(SiMe3)2 (X = H, Cl, or OH) were synthesized for the first time by the reaction of tetrakis(trimethylsilyl)silane with SbCl5. The molecular and crystal structure of bis(hydroxydimethylsilyl)bis(trimethylsilyl)silane was established by X-ray diffraction. The fragmentation of the resulting compounds under electron impact was studied by mass spectrometry. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 461–466, March, 2006.  相似文献   

18.
Vanadium–silver bimetallic oxide cluster ions (VxAgyOz+; x=1–4, y=1–4, z=3–11) are produced by laser ablation and reacted with ethane in a fast‐flow reactor. A reflectron time of flight (Re‐TOF) mass spectrometer is used to detect the cluster distribution before and after the reactions. Hydrogen atom abstraction (HAA) reactions are identified over VAgO3+, V2Ag2O6+, V2Ag4O7+, V3AgO8+, V3Ag3O9+, and V4Ag2O11+ ions, in which the oxygen‐centered radicals terminally bonded on V atoms are active sites for the facile HAA reactions. DFT calculations are performed to study the structures, bonding, and reactivity. The reaction mechanisms of V2Ag2O6++C2H6 are also given. The doped Ag atoms with a valence state of +1 are highly dispersed at the periphery of the VxAgyOz+ cluster ions. The reactivity can be well‐tuned gradually by controlling the number of Ag atoms. The steric protection due to the peripherally bonded Ag atoms greatly enhances the selectivity of the V–Ag bimetallic oxide clusters with respect to the corresponding pure vanadium oxide systems.  相似文献   

19.
A new Briggs? Rauscher oscillating reaction with a tetraazamacrocyclic nickel(II) complex [NiL](ClO4)2 as catalyst and pentane‐2,4‐dione (pe) as the substrate is reported. The ligand L is 5,7,7,12,14,14‐hexamethyl‐1,4,8,11‐tetraazacyclotetradeca‐4,11‐diene. The experimental results indicate that iodine ion may be an important intermediate, and free radicals can be involved in the reaction. A tentative mechanism based on Noyes? Furrow model (NF model) is proposed. Moreover, other factors, such as the variation of concentration of the components and temperature on the oscillator, are discussed.  相似文献   

20.
We present here a novel programmable polymerization route for the synthesis of new indole‐based polymers via a catalyst‐free nucleophilic substitution reaction. The polycondensation of 4‐hydroxyindole with different activated difluoro monomers undergoes in N‐methylpyrrolidinone, affording soluble poly(N‐aryleneindole ether)s (PEINs) with high molecular weights (Mw up to 486,000) in high yields (up to 96%). The structures of the polymers are characterized by means of FT‐IR, 1H NMR spectroscopy and elemental analysis, the results show good agreement with the proposed structures. The resulting polymers are processable and enjoy high glass transition temperatures (Tgs > 180 °C) and thermal stability (Tds > 420 °C). Thin films of PEINs show great mechanical behaviors with high tensile strength up to 104 Mpa, and good optical transparency. In addition, due to the indole moieties in the main chains, all these PEINs are endowed with significantly strong photonic luminescence in chloroform and display highly solvent‐dependent emission bands. Especially, the polymer PEIN‐3 carrying sulfonyl units, shows outstanding blue‐light emission with high quantum yields (45.2%, determined against quinine sulfate). The results obtained by cyclic voltammetry suggest that PEINs possess good electroactivity. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 313–320  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