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 453 毫秒
1.
In the presence of a PdI2-based catalytic system, 1,2-diols undergo an oxidative carbonylation process to afford 5-membered cyclic carbonates in good to excellent yields (84-94%) and with unprecedented catalytic efficiencies for this kind of reaction (up to ca. 190 mol of product per mol of PdI2). Under similar conditions, 6-membered cyclic carbonates are obtained for the first time through a direct catalytic oxidative carbonylation of 1,3-diols (66-74% yields).  相似文献   

2.
A novel and convenient approach to furan-3-carboxylic esters 2 is presented, based on palladium-catalyzed direct oxidative carbonylation of readily available 3-yne-1,2-diols 1. The process, corresponding to a sequential combination between a 5-endo-dig heterocyclodehydration step and an oxidative alkoxycarbonylation stage, is catalyzed by PdI2 in conjunction with an excess of KI under relatively mild conditions (100 °C in ROH under 40 atm of a 4:1 mixture of CO-air).  相似文献   

3.
The mechanism of the carbonylation reaction of allyl halides catalyzed by nickel (Ni(CO)4) and palladium (Cl2Pd(PPh3)2) complexes is theoretically investigated at the DFT level using the hybrid B3LYP functional. The favored reaction path to carbonylation corresponds, for both catalysts, to a direct attack of the halogen on the metal. This affords η1 intermediates that can undergo the final carbonylation step. It is also possible to obtain the acyl product (β,γ-unsaturated acyl halides) from η2 and/or η3 intermediates. However, in this case, the barrier of the rate-determining step to carbonylation is much higher. Since a channel on the potential surface connects rather easily the η2 or η3 intermediates to the η1 intermediates, an alternative and competitive path leading to the acyl products can originate from the η2 or η3 intermediates, follow the η23 → η1 transformation, then undergo the final carbonylation step.  相似文献   

4.
The reaction of copper(II) hydroxocarbonate, mandelic acid (H2MANO) and 2,2′-bipyridine (bpy) or 1,10-phenanthroline (phen) in water affords [Cu(bpy)(μ2-MANO)]2 · 8H2O (1), [Cu(bpy)(MANO)] · 4H2O (2) and the opened tetranuclear hydroxo-bridged copper(II) complexes of formulae [Cu43-OH)22-MANO)2(bpy)4](phglyo)2 · 8H2O (3) (phglyo = phenylglyoxylate) or [Cu43-OH)22-OH)2(OH2)2(phen)4](Bza)2(OH)2 · 5H2O (4) (Bza = benzoate), respectively. The compounds have been characterized by spectroscopic techniques and studied by single-crystal X-ray diffractometry. The formation of 3 and 4 takes place in basic media through dehydrogenation or oxidative dehydrogenation followed by in situ oxidative decarboxylation of mandelic acid to phenylglyoxylate or benzoate, respectively. These results indicate that cooperative catalysis of diimine ancillary ligands and copper(II) is essential.  相似文献   

5.
The density functional theory calculations were used to study the influence of the substituent at P on the oxidative addition of PhBr to Pd(PX3)2 and Pd(X2PCH2CH2PX2) where X = Me, H, Cl. It was shown that the Cipso-Br activation energy by Pd(PX3)2 correlates well with the rigidity of the X3P-Pd-PX3 angle and increases via the trend X = Cl < H < Me. The more rigid the X3P-Pd-PX3 angle is, the higher the oxidative addition barrier is. The exothermicity of this reaction also increases via the same sequence X = Cl < H < Me. The trend in the exothermicity is a result of the Pd(II)-PX3 bond strength increasing faster than the Pd(0)-PX3 bond strength upon going from X = Cl to Me. Contrary to the trend in the barrier to the oxidative addition of PhBr to Pd(PX3)2, the Cipso-Br activation energy by Pd(X2PCH2CH2PX2) decreases in the following order X = Cl > H > Me. This trend correlates well with the filled dπ orbital energy of the metal center. For a given X, the oxidative addition reaction energy was found to be more exothermic for the case of X2PCH2CH2PX2 than for the case of PX3. This effect is especially more important for the strong electron donating phosphine ligands (X = Me) than for the weak electron donating phosphine ligands (X = Cl).  相似文献   

6.
A novel cobalt-tetraphenylporphyrin/reduced graphene oxide (CoTPP/RGO) nanocomposite was prepared by a π–π stacking interaction and characterized by ultraviolet–visible absorption spectroscopy (UV–vis), Fourier transform infrared spectroscopy (FTIR) and electrochemical impedance spectroscopy (EIS). The CoTPP/RGO nanocomposite exhibited high electrocatalytic activity both for oxidation and reduction of H2O2. The current response was linear to H2O2 concentration with the concentration range from 1.0 × 10−7 to 2.4 × 10−3 mol L−1 (R = 0.998) at the reductive potential of −0.20 V and from 1.0 × 10−7 to 4.6 × 10−4 mol L−1 (R = 0.996) at the oxidative potential of +0.50 V. The H2O2 biosensor showed good anti-interfering ability towards oxidative interferences at the oxidative potential of +0.50 V and good anti-interfering ability towards reductive interferences at the reductive potential of −0.20 V.  相似文献   

7.
A series of sulfonated copolyimides containing benzimidazole groups (SPIs) were synthesized by random copolymerization of 1,4,5,8-naphthalenetetracarboxylic dianhydride (NTDA), 2-(4-aminophenyl)-5-aminobenzimidazole (APABI), 4,4′-diaminodiphenyl ether-2,2′-disulfonic acid (ODADS) and 9,9-bis(4-aminophenyl)fluorene (BAPF) in m-cresol in the presence of benzoic acid and triethylamine at 180 °C for 20 h. Membranes with good mechanical properties were prepared by solution cast method. Proton exchange treatment resulted in ionic cross-linking and the membranes were further covalently cross-linked by treating them in polyphosphoric acid (PPA) at 180 °C for 6 h. The covalently cross-linked membranes displayed slightly lower ion exchange capacities (IECs) and proton conductivities than the corresponding covalently uncross-linked ones because small part of the sulfonic acid groups had been consumed during the cross-linking process. Fenton’s test (3% H2O2 + 3 ppm FeSO4, 80 °C) revealed that benzimidazole groups played an important role in the radical oxidative stability of the membranes, while the cooperative effect of benzimidazole groups and covalent cross-linking led to much more significant enhancements in the radical oxidative stability of the membranes than each alone. The membrane 4 (ODADS/APABI/BAPF = 2/1/1, by mol), for example, after covalent cross-linking could maintain membrane form within 8 h measurement, which was much longer than that (3 h) before covalent cross-linking under the same conditions. The membrane 5 (ODADS/BAPF = 3/1, by mol) without benzimidazole groups, however, after covalent cross-linking started to break into pieces after 85 min measurement, which was only slightly longer than that (60 min) before cross-linking under the same conditions.  相似文献   

8.
The new naphthofuranoneacetic acid methyl ester IX can be obtained catalytically with fair selectivity from the carbonylation of the diacetylenic alcohol I. The use of the soluble complex Pd(thiourea)4I2 leads to a catalytic reaction resulting from the combination of oxidative carbonylation and hydrogenolysis steps in an additive carbonylation.  相似文献   

9.
Three intact and four degraded hyaluronans were investigated by using chemiluminometry, differential scanning calorimetry, and thermogravimetry. Degradation of hyaluronan was induced by a system containing H2O2 alone (882 mM); 55 mM H2O2plus 1.25 μM CuCl2; NaOCl alone (10 mM); and NaOCl plus CuCl2 and ascorbic acid (10 mM, 0.1 μM, and 100 μM, respectively). The four different oxidative systems yielded biopolymer fragments represented by similar viscosity characteristics. The results obtained by using chemiluminescence and thermoanalytical methods indicate that hyaluronans of similar rheological properties could be distinguished from each other.  相似文献   

10.
A method was developed for the analysis of four aliphatic diamines by capillary zone electrophoresis using pre-column derivatization with naphthalene-2,3-dicarboxaldehyde (NDA)/CN and amperometric detection. The pre-column derivatization reaction conditions including the molar ratio of NDA to amines, the cyanide concentration, the pH value of derivatization buffer, and the reaction time, were investigated. The separation of four derivatives of aliphatic diamines has been optimized by capillary zone electrophoresis (CZE) using end-column amperometric detection with a carbon fiber microelectrode, at a constant potential of 0.7 V versus SCE. The optimum conditions for the separation were 10 mM Tris-H3PO4 (pH 4.0) for the running buffer solution, 15 kV for the separation voltage. The detection limits for diaminopropane, putrescine, cadaverine, diaminohexane were 6.7×10−8, 5.1×10−8, 1.9×10−7 and 3.8×10−7 M, respectively (S/N=3). The proposed method was applied to the determination of aliphatic diamines in a lake water sample by the standard addition method. The recovery of these amines in water was 89.9-107%.  相似文献   

11.
The rhodium catalysed 1,4-carbonylative addition of arylboronic acids to methyl vinyl ketone under carbon monoxide pressure was studied. High yields of 1,4-diketones were obtained using a catalytic system formed from Rh(COD)2BF4 (COD=1,5-cyclooctadiene) and triphenylphosphine even at very low catalyst loading (0.02 mol %). A short synthetic procedure combining this carbonylation reaction with a subsequent cyclisation step affords pyrroles or furans.  相似文献   

12.
The chemistry of bis(3,5-dimethylpyrazolyl)methane complexes of copper(I) has been investigated and a dinuclear copper(I) derivative of formula {Cu2[μ-CH2(3,5-Me2Pz)2]2}(TfO)2 [TfO = trifluoromethanesulphonate anion, ], characterized by an uncommon bridging coordination of the bis(pyrazolyl)methane ligands, has been isolated and characterized by X-ray diffraction methods. Moreover, new olefin derivatives of general formula [Cu[CH2(3,5-Me2Pz)2](olefin)]TfO have been prepared (olefin: coe = cyclooctene, van = 4-vinylanisole, nbe = norbornene), their carbonylation reactions, {Cu[CH2(3,5-Me2Pz)2](olefin)}TfO + CO ? {Cu[CH2(3,5-Me2Pz)2](CO)}TfO + olefin, have been studied gas volumetrically and the thermodynamical parameters of the equilibria for the displacement of the coordinated olefin by carbon monoxide have been determined.  相似文献   

13.
Biphen(OPR2) (with R: Ph, iPr, Cy) is reacted with [Rh(COE)2Cl]2. The corresponding μ-chloro-bridged dimers are received. An X-ray analysis of [Biphen(OPCy2)RhCl]2 is included. This compound shows a dynamic behaviour in solution, ascribed to a monomer/dimer equilibrium. The difference of the Biphen ligands to Milsteins PCP pincer-type ligand is shown. A catalytic cycle for biphenyl metathesis containing the coupling of oxidative addition and reductive elimination of the bridging C-C single bond in the biphenyl fragment using RhI/III complexes and the concept of chelating assistance was calculated using DFT (B3PW91/LANL2DZ). According to the calculations the activation energy of the oxidative addition is about 30 kcal/mol and for the reductive elimination about 19 kcal/mol. The fac-RhIII complex is by far the most stable compound, but the formation of it is kinetically strongly disfavoured. Pre-catalysts (COD)M(Ph-O-PR2) (M: Rh, Ir) were synthesized by pre-coordinating the phosphinite to the metal (X-ray structures of four such compounds included) followed by treatment with 2 equiv. of sec. BuLi (X-ray structures of two such compounds included). In case of Ir this synthesis is complicated by C-H activation (X-ray structure of (COD)Ir(H)(Cl)(2-Br-phenyl-O-(diisopropylphosphinite)) included) and fast oxidative addition of the Ph-C-Halide bond. For (COD)Ir(H)(Cl)(2-phenyl-O-(diisopropylphosphinite)) the C-H activation is reversible and thermodynamic parameters for the ring closure reaction were determined by VT-NMR measurement (ΔH = −21.1 ± 0.5 kJ/mol, ΔS = −62.8 ± 1.7 J/(mol K)). The pre-catalysts were reacted with Biphen(OPR2) to enter the calculated catalytic cycle. With Rh as center metal this reaction works out cleanly to give new complexes with the three P-atoms coordinated to one Rh center. No hemi-labile character was found for these P-donors even at 105 °C in toluene. If (COD)Rh(2-phenyl-O-(diisopropylphosphinite)) is reacted with 2 equiv. of 2-iodo-phenyl-O-(diisopropylphosphinite) oxidative addition of one C-Iodo bond is observed and the corresponding mer-RhIII complex is received. Upon treatment with 2 equiv. of sec. BuLi the resulting product is(Biphen(OPiPr2))RhI(2-phenyl-O-(diisopropylphosphinite)) rather than mer-RhIII(2-phenyl-O-(diisopropylphosphinite))3. Reaction of [Rh(COD)Cl]2 with 3 equiv. of 2-bromo-phenyl-O-(diphenylphosphinite) shows a fast scrambling of the chlorine into all possible ortho positions of the phenolate rings in the final RhIII reaction product.  相似文献   

14.
The oxidative condensation of (5-amino-2,3-dihydro-1,4-phtalazinedione) luminol was carried out under their oxidation by (NH4)2S2O8 and KIO3 in the mixed water-organic (namely dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF) or N-methyl-2-pyrrolidinone (MPD)) solvent under the volume ratio 1:9. The structure of the products was studied by IR and Raman spectroscopy, elemental analysis and the derivatographic method. It was determined that oxidation by KIO3 (E = 1.085 V) occurs on the amide nitrogen atoms, while in the case of (NH4)2S2O8 (E = 2.05 V) it occurs on the amino-group. The structure and thermal stability of the obtained products is determined by the nature of the oxidant. The active decomposition of luminol begins at td = 334.5 °C, while for the specimens of the polyluminols, which were synthesized in the water-DMSO and water-MPD mixtures, td equals 356.7 and 409.1 °C respectively. The worst thermal stability has products of luminol oxidation by KIO3 (td = 282.5 °C). The mechanism of the luminol oxidative polymerization by (NH4)2S2O8 and KIO3 has been proposed.  相似文献   

15.
Reductive carbonylation of rhodium(III) chloride complexes, commercial RhCl3 · nH2O neutralized with BaCO3, (Me2NH2)2[RhCl5(DMF)], (PPh4)[RhCl4(H2O)2], RhCl3(DMF)3, RhCl3(CH3CN)3, RhCl3(CH3CN)2(DMF), [Rh(CO)2Cl3]2, and rhodium(I) complex, Rh(PPh3)3Cl, by N,N-dimethylformamide (DMF) is studied. The data obtained support the conception of direct carbonyl group transfer from DMF molecule to the Rh metal center. The mechanistic scheme of carbonylation process is refined and discussed with regard of new experimental results.  相似文献   

16.
Taking advantage of palladium peculiar “rollover” C,N cyclometallation, it is possible to promote C(3) functionalization of 6-alkyl-substituted-2,2′-bipyridines. The carbonylation reaction of rollover species [Pd(Ln)Cl]2, (HL1 = 6-isopropyl-2,2′-bipy, 1; HL2 = 6-neopentyl-2,2′-bipy, 2; HL3 = 6-ethyl-2,2′-bipy, 3; HL4 = 6-methyl-2,2′-bipy, 4) allowed the synthesis of 2-(pyridin-2-yl)-6-alkyl-nicotinic acids or esters. These nicotinic derivatives are extremely rare and, as far as we know, quite unreported in the case of the 6-substituted molecules.  相似文献   

17.
18.
A novel iridium-catalyzed oxidative esterification reaction of aliphatic aldehydes and olefinic alcohols in toluene was found under mild conditions of [IrCl(cod)]2 (5 mol %) in combination with K2CO3 (10 mol %) at rt.  相似文献   

19.
The complex [(C6H5)2SnCl(HMNA)] (1) where H2MNA is thioamide 2-mercapto-nicotinic acid has been synthesized by reacting a methanolic solution of di-chloro-di-phenyltin(IV) Ph2SnCl2 with an aqueous solution of 2-mercapto-nicotinic acid, containing twofold amount of potassium hydroxide. The Sn/ligand molar ratio is 2:1. The complex was characterized by elemental analysis, FT-IR and Mössbauer spectroscopic techniques. In addition the crystal structure of the molecule was determined by an X-ray diffraction at 293(2) K. C18H14ClNO2SSn is monoclinic, space group P21/n, a = 15.089(3) Å, b = 15.846(3) Å, c = 16.691(3) Å, β = 105.57(3)°, Z = 8. The ligand coordinates to the metal center through the exocyclic sulfur and the heterocyclic nitrogen atoms, forming a four membered ring. The coordination sphere around the tin(IV) ion is completed with two carbon atoms from the two phenyl groups and one chlorine atom. The geometry around the tin atom can be described either as trigonal bipyramidal or tetragonal pyramidal. Finally, the influence of the complex [(C6H5)2SnCl(HMNA)] (1) upon the catalytic peroxidation of linoleic acid to hydroperoxylinoleic acid by the enzyme lipoxygenase (LOX) was also kinetically and theoretically studied and the results compared with the ones of the corresponding binuclear complex [(C6H5)3Sn(MNA)Sn(C6H5)3 · (acetone)] (2) reported in the literature.  相似文献   

20.
The four complexes [Pd(H)(Cl)L2] and [Pd(H)(SnCl3)L2], L = PPh3, PCy3, have been synthesized and fully characterized by multinuclear NMR. They represent the active species of the hydride palladium-catalyzed alkoxycarbonylation of terminal alkenes. Isolation of the model acylplatinum complex, resulting from the carbonylation of dihydromyrcene, clearly shows that SnCl2 as co-catalyst produces a SnCl3 ligand which modulates the metal center electron density.  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