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.
Electron attachment reactions and negative ion mass spectra which were obtained under negative chemical ionization conditions have been examined for a series of 21 nickel(II) bis-chelates of formula Ni[R1CXCHCYR2]2. Three ligand donor atom sets (X, Y), respectively O4, O2S2, S4 were investigated for each of the substituent combinations, viz.: R1=CH3, CF3 or C2H5O, R2=CH3; R1=C6H5, CH3 or CF3, R2=C6H5; and R1 = R2 = tert?C4H9. While the ligand substituent combinations exerted considerable influence over the various ion decomposition reactions, the relative molecular ion stabilities were largely dependent on the ligand donor atom sets and followed the sequence O4? O2S2>S4 for most substituent combinations. Rationalizations are offered in terms of reductive electron capture reactions involving metal-based orbitals, as well as the increasing stabilities of reaction products as sulphur is incorporated into the ligand donor atom sets. A comparison is also given of negative ion mass spectral data obtained under electron impact conditions as well as negative chemical ionization conditions when methane was used as an electron energy moderating gas.  相似文献   

2.
Reaction conditions for the C? C cross‐coupling of O6‐alkyl‐2‐bromo‐ and 2‐chloroinosine derivatives with aryl‐, hetaryl‐, and alkylboronic acids were studied. Optimization experiments with silyl‐protected 2‐bromo‐O6‐methylinosine led to the identification of [PdCl2(dcpf)]/K3PO4 in 1,4‐dioxane as the best conditions for these reactions (dcpf=1,1′‐bis(dicyclohexylphosphino)ferrocene). Attempted O6‐demethylation, as well as the replacement of the C‐6 methoxy group by amines, was unsuccessful, which led to the consideration of Pd‐cleavable groups such that C? C cross‐coupling and O6‐deprotection could be accomplished in a single step. Thus, inosine 2‐chloro‐O6‐allylinosine was chosen as the substrate and, after re‐evaluation of the cross‐coupling conditions with 2‐chloro‐O6‐methylinosine as a model substrate, one‐step C? C cross‐coupling/deprotection reactions were performed with the O6‐allyl analogue. These reactions are the first such examples of a one‐pot procedure for the modification and deprotection of purine nucleosides under C? C cross‐coupling conditions.  相似文献   

3.
An efficient and straightforward protocol for one-pot, three-component reaction of aryl glyoxal monohydrates 1a-h , 5-amino-1-aryl-3-methylpyrazoles 2a , b and 4-hydroxyquinoline-2(1H)-one ( 3 ) or 2-hydroxy-1,4-naphthoquinone ( 4 ) using silver nanoparticles (AgNPs) as a high performance nanocatalyst in H2O/EtOH at 60°C afforded the corresponding polyfunctionalized benzo[h]pyrazolo[3,4-b][1,6]naphthyridines 5a-h and benzo[g]pyrazolo[3,4-b]quinolines 6a-i , respectively. Excellent catalytic activity, high yields, employing green media and green nanocatalyst, cost-effective and simple procedure are some notable advantages of using AgNPs as a noble metal nanocatalyst in this synthetic strategy. The structures of fused heterocycles were confirmed by their Fourier transform infrared, proton nuclear magnetic resonance (1H-NMR), and 13C-NMR spectral data and microanalysis.  相似文献   

4.
The reactions of CH3 radicals with O(3P) and O2 have been studied at 295 K in a gas flow reactor sampled by a mass spectrometer. For the reaction between CH3 and O, conditions were such that [O] » [CH3] and the methyl radicals decayed under pseudo-first-order conditions giving a rate coefficient of (1.14 ± 0.29) × 10?10 cm3/s. The reaction between CH3 and O2 was studied in separate experiments in which CH3 decayed under pseudo-first-order conditions. In this case, the rate coefficient obtained increased with increasing concentration of the helium carrier gas. This was varied over the range of 2.5–25 × 1016 cm?3, resulting in values for the apparent two-body rate coefficient ranging from 1 × 10?14 to 5.2 × 10?14 cm3/s. No evidence was found for the production of HCHO by a direct two-body process involving CH3 + O2, and an upper limit of 3 × 10?16 cm3/s was placed on the rate coefficient for this reaction. The experimental results for the apparent two-body rate coefficient exhibit the curvature one would expect for an association reaction in the fall-off region. Calculations used to extrapolate these measurements to the low-pressure limit yield a value for k0 of (3.4 ± 1.1) × 10?31 cm6/s, which is more than a factor of 2 higher than previous estimates.  相似文献   

5.
Thanks to the superior redox potential of platinum(II) complex compared with that of Ru(bpy)32+ in the excited state, an efficient and selective visible‐light‐induced CDC reaction has been developed by using a catalytic amount (0.25 %) of 1 . With the aid of FeSO4 (2 equiv), the corresponding amide could not be detected under visible‐light irradiation (λ=450 nm), but the desired cross‐coupling product was exclusively obtained under ambient air conditions. A spectroscopic study and product analysis revealed that the CDC reaction is initiated by photoinduced electron‐transfer from N‐phenyltetrahydroisoquinoline to the complex. An EPR (electron paramagnetic resonance) experiment provides direct evidence on the generation of superoxide radical anion (O2? . ) rather than singlet oxygen (1O2) under irradiation of the reaction system, in contrast to that reported in the literature. Combined, the photoinduced electron‐transfer and subsequent formation of superoxide radical anion (O2? . ) results in a clean and facile transformation.  相似文献   

6.
The chemiluminescence of the Cypridina luciferin analog, 2-methyl-6-(p-methoxyphenyl)-3,7-dihydroirnidazo[1,2-a]pyrazin-3-one (MCLA), with O2 (1Δg) generated by the retro-Diels-Alder reaction of 3-(4′-methyI-l'-naphthyl)-propionic acid endoperoxide was studied in an aqueous solution with pH 7.12 at 37°C. The retro-Diels-Alder reaction occurs with a first-order rate constant of (4.16 ± 0.13) × 10?4/s to quantitatively yield O2 (1Δg) and 3-(4′-methyl-l'-naphthyl (-propionic acid. MCLA consumed equimolar amounts of O2 (1Δg) with a second-order rate constant (6.96 ± 0.27) × 108/M/s to emit light in an aqueous solution with pH 7.12 at 37°C. The chemiluminescence spectrum was identified as the fluorescence spectrum of 2-acetylamino-5-(p-methoxyphenyl)pyrazine (OMCLA), a major chemiluminescence reaction. Chemiluminescence spectra and product yields for MCLA reactions with O21Δg, with O2 (3Σ?g) and with superoxide anion radicals are identical, suggesting that all of these reactions occur via a common MCLA-2-hydrope-roxide intermediate formed by a combination of MCLA radicals and superoxide anion radicals. We have established practical use of NEPO as an O2 (1Δg) source and MCLA as a biological probe for detecting O2 (1Δg).  相似文献   

7.
The kinetics of hydroquinone-inhibited oxidation of acrylic acid and methyl methacrylate was studied volumetrically by measuring the oxygen uptake in the presence of an initiator (azobisisobutyronitrile) at T = 333 K and P O 2 = 1 and 0.21 atm. The oxidation of acrylic acid inhibited by 4-methoxyphenol was studied under the same conditions for comparison. The rate constants of the reactions of the peroxyl radicals of acrylic acid (∼CH2CH(COOH)O2·) and methyl methacrylate (∼CH2CMe(COOMe)O2·) with hydroquinone (HOC6H4OH) (1.20 × 105 and 7.16 × 104 l mol−1 s−1, respectively) and of the reaction of peroxyl radicals of acrylic acid with 4-methoxyphenol (p-CH3OC6H4OH) (3.25 × 104 l mol−1 s−1) were measured. The stoichiometric inhibition factors f were determined. The reaction between the semiquinone radical and oxygen, O2 + HOC6H4O·, plays an important role, decreasing the factor f and the efficiency of the inhibition effect of hydroquinone. The rate constants of this reaction were calculated from kinetic data: k = 5.72 × 102 (in acrylic acid) and 4.60 × 102 l mol−1 s−1 (in methyl methacrylate).  相似文献   

8.
The reaction of 3-amino-1,2,4-triazole ( 1 ) with substituted methyl cinnamates 2a-h leads selectively to the formation of 7-aryl-6,7-dihydro[1,2,4]triazolo[1,5-a]pyrimidin-5(4H)-ones 3a-h . The structure eluci dation of the products is based on ir, 1H and 13C nmr measurements and X-ray diffraction.  相似文献   

9.
Reactions of cerium(III) nitrate, Ce(NO3)3?6 H2O, with different carboxylic acids, such as pivalic acid, benzoic acid, and 4‐methoxybenzoic acid, in the presence of a tridentate N,N,N‐donor ligand, diethylenetriamine (L1), under aerobic conditions yielded the corresponding cerium hexamers Ce6O8(O2CtBu)8(L1)4 ( 1 ), Ce6O8(O2CC6H5)8(L1)4 ( 2 ), and Ce6O8(O2CC6H4‐4‐OCH3)8(L1)4 ( 3 ). Hexamers 1 , 2 , and 3 contain the same octahedral CeIV6O8 core, in which all interstitial oxygen atoms are connected by μ3‐oxo bridging ligands. In contrast, treatment of the CeIV precursor (NH4)2Ce(NO3)6 (CAN) with pivalic acid and the ligand L1 under the same conditions afforded Ce6O4(OH)4(O2CtBu)12(L1)2 ( 4 ), exhibiting a deformed octahedral CeIV6O4(OH)4 core containing μ3‐oxo and μ3‐hydroxo moieties in defined positions. In contrast to the formation of 1 – 3 , the use of N‐methyldiethanolamine (L) in the reaction with Ce(NO3)3?6 H2O and pivalic acid afforded a previously reported CeIII dinuclear cluster, Ce2(O2CtBu)6L2, even in the presence of dioxygen. ESI‐MS analysis of the reaction mixture clearly indicated the importance of the ligand L1 in promoting oxidation of the CeIII aggregates, [Cen(O2CtBu)3n(L1)2], which is necessary for the formation of CeIV hexamers.  相似文献   

10.
A strategy for the vanadium-catalyzed dehydrative C- and S-alkylation by nucleophilic substitution of benzhydrols with arenes and thiols is reported. The alkylation was achieved with the divanadium oxoperoxo complex [K3(V+5)2(O22−)4(O2−)2(μ-OH)] in water under air. The newly developed transformation could accommodate a broad substrate scope, including (hetero)arenes and thiols (34 examples). Both the symmetrical and unsymmetrical benzhydrols furnished excellent yields of the alkylated product under mild reaction conditions. The scope of this strategy was further extended to synthesize bis-benzylated arenes (poly-arylated products) in high yields and regioselectivities. The green metrics determination of all the alkylated products suggests the technical and environmental benefits of the present protocol. The longevity experiment reveals the catalytic activity was maintained over seven cycles. To understand the mechanism of the present reaction, spectroscopic and kinetic studies were undertaken. This simple protocol, which affords the desired products with water as the by-product, can be achieved under mild conditions without needing a base or other additives.  相似文献   

11.
The chemistry of the HC(O)CO radical, produced in the oxidation of glyoxal, has been studied under conditions relevant to the lower atmosphere using an environmental chamber/Fourier Transform infrared spectrometric system. The chemistry of HC(O)CO was studied over the range 224–317 K at 700 Torr total pressure and was found to be governed by competition between unimolecular decomposition [to HCO and CO, reaction (5)] and reaction with O2 [to form HO2 and 2CO, reaction (6a), or HC(O)C(O)O2, reaction (6b)]. The rate coefficient for decomposition relative to that of reaction with O2 increases with increasing temperature. Assuming a value for k6 of 10−11 cm3 molecule−1 s−1, the following expression for the unimolecular decomposition is obtained at 700 Torr, k5 = 1.4+9/−1.1 × 1012 exp(−3160 ± 500/T). The rate coefficients for reactions (6a) and (6b) are about equal, with no strong dependence on temperature. The reaction of HC(O)C(O)O2 with NO2 was also studied. Final product analysis was consistent with the formation of HCO, CO2, and NO3 as the major products in this reaction; no evidence for the PAN‐type species, HC(O)C(O)O2NO2, was found even at the lowest temperature studied (224 K). The UV‐visible absorption spectrum of glyoxal is also reported; results are in substantive agreement with previous studies. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 149–156, 2001  相似文献   

12.
Photochemical processes involving singlet oxygen (O2(a1Δ)), oxygen atoms, and ozone are critical in determining atmospheric ozone concentrations. Here we report on kinetic measurements and modeling that examine the importance of the reactions of vibrationally excited ozone. Oxygen atoms and O2(a1Δ) were produced by UV laser photolysis of ozone. Time‐resolved absorption spectroscopy was used for O3 concentration measurements. It was found that vibrationally excited ozone formed by O + O2 + M → O3(ν) + M recombination reacts effectively with O2(a1Δ) and O atoms. The reaction O3(υ) + O2(a1Δ) → O + 2O2 results in a reduction of the ozone recovery rate due to O atom regeneration, whereas the reaction O3(υ) + O → 2O2 removes two odd oxygen species, resulting in incomplete ozone recovery. The possible impact of these reactions on the atmospheric O2(a1Δ) and O3 budgets at altitudes in the range of 80–100 km is considered.  相似文献   

13.
The oxidation of methylhydrazine (monomethylhydrazine MMH) is studied in a strictly single-phase gaseous medium, in a reconstituted surrounding atmosphere. In order to work under such conditions, a specific apparatus was assembled to monitor the evolution of the reactants through time. The reaction kinetics was studied at 50°C to avoid condensation of the water formed and for O2/MMH mole ratios between 1 and 4. Under these conditions, the partial pressures of O2 were between 0.05 and 0.18 bar (4% MMH per volume). The main reaction products were monitored through time and identified by gas chromatography coupled with mass spectrometry (GC/MS). The products were: N2, CH4, CH3-NH-N=CH2 (formaldehyde monomethylhydrazone), NH3, H2O, CH3OH, and CH2=N-N=N-CH3 (2,3,4-triazapenta-1,3-diene). The formation of nitrosamines was not observed under these experimental conditions. The rate laws related to the disappearance of the reagents were clearly established and can be described by 2 consecutive parallel reactions of order 2 whose rate constants are: k 1 = 7.57 × 10−2 bar−1 min−1 and k 2 = 0.5 bar−1 min−1. Analysis of the products permitted establishing an approximated balance of the global reaction.  相似文献   

14.
In order to synthesize poly-(fluorinated alkanesulfonamides) a series of model experiments were carried out: (1) reactions of fluorinated alkanesulfonyl fluorides with amines, (2) reactions of fluorinated alkanesulfonyl chloride with amines and (3) reactions of sodium salts of fluorinated alkanesulfonamides with alkyl iodides of fluorinated alkanesulfonic acid esters. Seventeen new fluorinated alkanesulfonamides were prepared in good yields, namely: RFO(CF2)2SO2NR1R2 (1a-h), R1R2NSO2RFSO2NR1R2 (2a-h) and [Cl (CF2)4O(CF2)2SO2NH(CH2)3]2 (3). Reaction of RFSO2NH2 with equivalent amount of NaOCH3 and methyl iodide was shown to give both the N-mono- and N,N-di-substituted amides. Consequently the N-monosubstituted alkanesulfonamides were chosen as monomers for syntheses of the poly-(fluorinated alkanesulfonamides) and two new polymers were synthesized. The effect of the condition of the polycondensation on M?n of the polymers were discussed and elemental composition, 19F NMR, IR, M?n, Tg, tensile strength, thermal and chemical stabilities of the polymers were measured. Several new perfluoroalkanesulfonyl chlorides CISO2RFSO2Cl (4a-c) and fluorinated alkanesulfonic acid esters (6a-d) were synthesized. However, reaction of CFCl2CF2O(CF2)2SO2F with AlCl3 was found to give Cl3CCF2O(CF2)2SO2F (5) instead of the expected sulfonyl chloride.  相似文献   

15.
Butane‐1‐sulfonic acid immobilized on magnetic Fe3O4@SiO2 nanoparticles (Fe3O4@SiO2‐Sultone) was easily prepared via direct ring opening of 1,4‐butanesultone with nanomagnetic Fe3O4@SiO2. The prepared reagent was characterized and used for the efficient promotion of the synthesis of barbituric acid and pyrano[2,3‐d] pyrimidine derivatives. All reactions were performed under mild and completely heterogeneous reaction conditions affording products in good to high yields. The catalyst is easily isolated from the reaction mixture by magnetic decantation and can be reused at least eight times without significant loss in activity.  相似文献   

16.
The mechanism of the collision-induced fragmentation of peracetylated methyl-α-D-glucopyranoside was investigated using deuterium-labelled acetates and sequential mass spectrometry. Loss of the substituent at C(1), the anomeric carbon, yields an ion of m/z 331, [C14H19O9]+. This ion further dissociates via two pathways, the first including m/z 271, [C12H15O7]+, 169, [C8H9O4]+ and 109, [C6H5O2]+, and the second including m/z 211, [C10H11O5]+, 169, [C8H9O4]+ and 127 [C6H7O3]+. The first path proceeds via loss of acetate at C(3), followed by a single-step concerted loss of acetates from C(2) and C(4), and ending with loss of acetate from C(6). The second path proceeds predominantly via loss of acetates from C(3) and C(4), elimination of ketene from the C(2)-acetate and finally loss of ketene from the acetate at C(6). This path is also characterized by an ill-defined series of parallel decomposition reactions involving acetates from other sites on the molecule. At low collision energy, and in the absence of collision gas (unimolecular reaction conditions), the former pathway predominates; m/z 331 dissociates via loss of acetate at C(3), followed by a single-step concerted loss of acetates from C(2) and C(4).  相似文献   

17.
A divanadium‐substituted phosphotungstate, [γ‐PW10O38V2(μ‐OH)2]3? ( I ), showed the highest catalytic activity for the H2O2‐based epoxidation of allyl acetate among vanadium and tungsten complexes with a turnover number of 210. In the presence of I , various kinds of electron‐deficient alkenes with acetate, ether, carbonyl, and chloro groups at the allylic positions could chemoselectively be oxidized to the corresponding epoxides in high yields with only an equimolar amount of H2O2 with respect to the substrates. Even acrylonitrile and methacrylonitrile could be epoxidized without formation of the corresponding amides. In addition, I could rapidly (≤10 min) catalyze epoxidation of various kinds of terminal, internal, and cyclic alkenes with H2O2 under the stoichiometric conditions. The mechanistic, spectroscopic, and kinetic studies showed that the I ‐catalyzed epoxidation consists of the following three steps: 1) The reaction of I with H2O2 leads to reversible formation of a hydroperoxo species [γ‐PW10O38V2(μ‐OH)(μ‐OOH)]3? ( II ), 2) the successive dehydration of II forms an active oxygen species with a peroxo group [γ‐PW10O38V2(μ‐η22‐O2)]3? ( III ), and 3) III reacts with alkene to form the corresponding epoxide. The kinetic studies showed that the present epoxidation proceeds via III . Catalytic activities of divanadium‐substituted polyoxotungstates for epoxidation with H2O2 were dependent on the different kinds of the heteroatoms (i.e., Si or P) in the catalyst and I was more active than [γ‐SiW10O38V2(μ‐OH)2]4?. On the basis of the kinetic, spectroscopic, and computational results, including those of [γ‐SiW10O38V2(μ‐OH)2]4?, the acidity of the hydroperoxo species in II would play an important role in the dehydration reactivity (i.e., k3). The largest k3 value of I leads to a significant increase in the catalytic activity of I under the more concentrated conditions.  相似文献   

18.
Ding Liu 《Acta Physico》2008,24(9):1584-1588
Photodegradation of a textile dye X3B and photoreduction of dichromate (Cr(VI)) in an acidic aqueous solution were studied under 320 nm cut-off UV light irradiation in the presence of two polyoxometalates (POM), H3PW12O40 (PW), and H4SiW12O40 (SiW). The reactions in POM-X3B-Cr(VI) system were faster than those in POM-X3B, POM-Cr(VI), and X3B-Cr(VI) systems. For all reactions, PW was more photoactive than SiW. The reaction rates were proportional to the initial concentration of each component. The effects of N2, O2, and air were small but regular, indicating Cr(VI) photoreduction by a reduced POM. Quenching experiments with H2O2 and ethanol revealed that X3B photodegradation mainly occurred through hydroxyl radical (OH). It was proposed that the production of OH and a reduced POM by the reaction between H2O and excited POM* was the rate determining step, with which all evidence could be well interpreted. Different effects of POM concentration in a two- or three-component system on the reaction rates suggested that the reaction between H2O and excited POM* was reversible.  相似文献   

19.
Summary. A series of benzoxazoles, benzothiazoles, benzimidazoles, and oxazolo[4,5-b]pyridines was efficiently synthesized from the reactions of o-aminophenols, o-aminothiophenol, o-phenylenediamines, and 2-amino-3-hydroxypyridine with orthoesters in the presence of catalytic amounts of Bi(III) salts, such as Bi(TFA)3, Bi(OTf)3, and BiOClO4 · xH2O under solvent-free conditions. The remarkable features of this new protocol are high conversion, very short reaction times, cleaner reaction profiles under solvent-free conditions, straightforward procedure, and use of relatively non-toxic catalysts.  相似文献   

20.
Boron Lewis acid-catalyzed and catalyst-free hydroboration reactions of imines are attractive due to the mild reaction conditions. In this work, the mechanistic details of the hydroboration reactions of two different kinds of imines with pinacolborane (HBpin) are investigated by combining density functional theory calculations and some experimental studies. For the hydroboration reaction of N-(α-methylbenzylidene)aniline catalyzed by tris[3,5-bis(trifluoromethyl)phenyl]borane (BArF3), our calculations show that the reaction proceeds through a boron Lewis acid-promoted hydride transfer mechanism rather than the classical Lewis acid activation mechanism. For the catalyst- and solvent-free hydroboration reaction of imine, N-benzylideneaniline, our calculations and experimental studies indicate that this reaction is difficult to occur under the reaction conditions reported previously. With a combination of computational and experimental studies, we have established that the commercially available BH3 ⋅ SMe2 can serve as an efficient catalyst for the hydroboration reactions of N-benzylideneaniline and similar imines. The hydroboration reactions catalyzed by BH3 ⋅ SMe2 are most likely to proceed through a hydroboration/B−H/B−N σ-bond metathesis pathway, which is very different from that of the reaction catalyzed by BArF3.  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