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 80 毫秒
1.
Reactivity in the Systems A/Cu/M/O (A = Na–Cs and M = Co, Ni, Cu, Ag); Synthesis and Crystal Structures of K3Cu5O4 und Cs3Cu5O4 The systems A/Cu/M/O with A = Na–Cs and M = Co, Ni, Cu, Ag have been investigated with preparative, thermoanalytical and in situ X‐ray techniques to study the reactivity. For the redox reaction Co/CuO in the presence of Na2O the intermediate, NaCuO, has been characterized. K3Cu5O4 was obtained by annealing intimate mixtures of K2O and CuO (molar ratio 1 : 1) in Ag containers at 500 °C. Cs3Cu5O4 could be synthezised by reaction of KCuO2 with Cs2O (molar ratio 1 : 1) in Cu containers at 500 °C. Both compounds crystallize in the space group P21/c with Z = 4 isotypic to Rb3Cu5O4 [IPDS data, Mo–Kα; K3Cu5O4: a = 946.0(1), b = 735.61(6), c = 1401.3(2) pm, β = 107.21(1)°; 2249 F2(hkl), R1 = 7.09%, wR2 = 11.42%; Cs3Cu5O4: a = 1027.7(1), b = 761.42(7), c = 1473.4(2) pm, β = 106.46(1)°, 1712 F2(hkl), R1 = 6.04%, wR2 = 14.22%]. Force constants obtained from FIR experiments for the deformation mode δ(O–Cu–O), the Madelung Part of the Lattice Energie, MAPLE, Effective Coordination Numbers, ECoN, calculated via Mean Effective Ionenradii, MEFIR, are given.  相似文献   

2.
The sterically hindered Schiff bases (L3–L5), prepared from 3,5‐dicumenyl salicylaldehyde and chiral amino alcohols, were used in combination with Ti(OiPr)4 for asymmetric oxidation of aryl methyl sulfides with H2O2 as terminal oxidant. Among the ligands L3–L5, L4 with a tert‐butyl group in the chiral carbon of the amino alcohol moiety gave the best result with 89% yield and 73% ee for the sulfoxidation of thioanisole under optimal conditions [with 1 mol% of Ti(OiPr)4 in a molar ratio of 100:1:1.2:120 for sulfide:Ti(OiPr)4:ligand:H2O2 in CH2Cl2 at 0 °C for 3 h]. The reaction afforded good yield (84%) with a moderate enantioselectivity (62% ee) even with a lower catalyst loading from 1.0 to 0.5 mol%. The oxidations of methyl 4‐bromophenyl sulfide and methyl 4‐methoxyphenyl sulfide with H2O2 catalyzed by the Ti(OiPr)4–L4 system gave 79–84% yields and 54–59% ee of the corresponding sulfoxides in CH2Cl2 at 20 °C. The chiral induction capability of the cumenyl‐modified sterically hindered Schiff bases for sulfoxidation was compared with the conventional Schiff bases bearing tert‐butyl groups at the 3,5‐positions of the salicylidenyl unit. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

3.
Thiomethylation of amino alcohols using formaldehyde and hydrogen sulfide   总被引:1,自引:0,他引:1  
Three-component condensations of formaldehyde with hydrogen sulfide and hydroxylamine or amino alcohols at a ratio of 3:2:1 give 47–73% of N-hydroxy(or hydroxyalkyl)-1,3,5-dithiazinanes. The reactions of CH2O with H2S and hydroxylamine or α-amino alcohols at a ratio of 4:3:1 involve both amino and hydroxy groups, leading to the corresponding dithiazinanes. 4-Aminobutan-1-ol reacts with CH2O and H2S under analogous conditions only at the amino group to form 4-(1,3,5-dithiazinan-5-yl)butan-1-ol.  相似文献   

4.
The reaction of isoprene with aniline, catalyzed by Pd (acac)2–(RO)3P‐CF3COOH, (1:4:4) (R = Me, Et, acac = (CH3CO)2CH‐) in MeCN solution, results in high (up to 89 mol.%) selectivity of N–(3‐methyl‐2‐buten‐1‐yl) aniline. The presence of telomeric products in the reaction mixture is observed at a P/Pd ratio of 1:2 and 1:1. The use of (1,1,1‐trifluoro, 4‐perfluorocyclo hexyl ‐2,4‐butanedionato) palladium as the catalyst gives rise to 92 mol% mol.selectivity of telomers by the favored tail‐to‐head and head to head coupling.  相似文献   

5.
A procedure is developed for the selective photometric determination of selenium(IV) in bottled drinking water by the oxidation of Methylene Blue in 1 M HCl to colorless decomposition products and of selenium(VI) by its interaction with the specified reagent at pH 5–6 with the formation of a colored ion pair. The limits of detection are 1 and 0.8 µg/L, respectively. At the concentration of selenium(IV) 2 µg/L, the admissible weight ratios are: SeO42-, Br3- (1: 20); Br (1: 60); I, IO3- and IO4- (1: 100). At equal concentration of selenium(VI), the following species: SeO42-(1: 20); Br3-, Br, I, IO3-, and IO4- (1: 100) do not interfere with the determination. Other anions and cations present in highly mineralized waters do not interfere with the determination. The relative error of determination is 8–10% in the concentration range 2–10 µg/L of selenium(IV) and selenium(VI) and does not exceed 5% in their concentration range of 10–100 µg/L.  相似文献   

6.
Corich core–Ptrich shell/C prepared by thermal decomposition and chemical reduction methods were treated by 20% H2SO4 aqueous solution and used as the electrocatalysts for the oxygen reduction reaction (ORR). The particle size range of Corich core–Ptrich shell (molar ratio of 0.92:1) on carbon powder support decreased from 3–8 to 1–6 nm when the time for the electrocatalysts immersed and treated with 20% H2SO4 aqueous solution increased from 0 to 4 h. Using Corich core–Ptrich shell (molar ratio of 0.92:1)/C treated with 20% H2SO4 from 0 to 4 h as the working electrode, the open circuit potential of ORR in 0.5 M HClO4 aqueous solution increased from 0.9995 to 1.0155 V, and the current density, mass activity, and specific activity at the overpotential of 0.1 V increased from 0.619 mA cm?2, 6.184 A g?1, and 18.614 μA cm?2 to 0.912 mA cm?2, 15.544 A g?1, and 23.413 μA cm?2, respectively.  相似文献   

7.
method has been developed for the selective photometric redox determination of periodate and iodate ions in bottled drinking water based on redox reactions of analytes with Methylene Blue with different duration of processes, products of which form the analytical signal. The limits of detection for periodate and iodate ions are 0.5 and 0.2 µg/L, respectively. The allowable weight ratios for concomitant ions for (at the analyte concentration 2 µg/L) are as follows: I, Br, IO3, BrO3, ClO, CIO, CIO2, CIO3 and CIO4(1: 100); and for IO3 (1 µg/L) are: BrO32, NO (1: 60), CIO, CIO2, (1: 100), and I, Br, IO4, CIO3, and CIO4 (1: 200). The HCIO3, Cl, and SO42- anions and Ca2+, Mg2+, Na+, K2+, and NH4+ cations are macrocomponents of drinking water and at total concentrations up to 10 g/L do not affect the results of analysis. In the concentration range 1–10 µg/L of IO4 andIO3, the total error of determination is 5–7%.  相似文献   

8.
The isoplethic sections in the diagram of the quaternary system Na+ , Mg2+ //Cl , SO2– 4 –H2 O were established at 25 and 30oC by analytical and conductometric measurements. Three compounds can be observed in the isoplethic sections: NaCl, Na2 SO4 and MgNa2 (SO4 )2 4 H2 O. Seven fields are determined, relating to the precipitation of one, two or three salts. The solubility range of MgNa2 (SO4 )2 4 H2 O is wide, while the liquidus curve of Na2 SO4 is very short. The compositions, expressed in Jänecke coordinates, at the eutonic and peritonic points, respectively, were: 42.70% Cl and 745% H2 O; 79.47% Cl and 787% H2 O; 71.6% Cl and 744% H2 O at 25°C; and 48.80% Cl and 715% H2 O; 80.20% Cl and 778% H2 O; 70.14% Cl and 707% H2 O at 30°C.  相似文献   

9.
Controlled radical polymerization of 4‐vinylpyridine (4VP) was achieved in a 50 vol % 1‐methyl‐2‐pyrrolidone/water solvent mixture using a 2,2′‐azobis(2,4‐dimethylpentanitrile) initiator and a CuCl2/2,2′‐bipyridine catalyst–ligand complex, for an initial monomer concentration of [M]0 = 2.32–3.24 M and a temperature range of 70–80 °C. Radical polymerization control was achieved at catalyst to initiator molar ratios in the range of 1.3:1 to 1.6:1. First‐order kinetics of the rate of polymerization (with respect to the monomer), linear increase of the number–average degree of polymerization with monomer conversion, and a polydispersity index in the range of 1.29–1.35 were indicative of controlled radical polymerization. The highest number–average degree of polymerization of 247 (number–average molecular weight = 26,000 g/mol) was achieved at a temperature of 70 °C, [M]0 = 3.24 M and a catalyst to initiator molar ratio of 1.6:1. Over the temperature range studied (70–80 °C), the initiator efficiency increased from 50 to 64% whereas the apparent polymerization rate constant increased by about 60%. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 5748–5758, 2007  相似文献   

10.
The interaction of rare-earth metals (Pr, Nd, Sm, Y, Er, and Yb) with selenium at Ln: Se = 1: 2, 2: 3, 3: 4, and 1: 1 under polythermal (293–1270 K) and isothermal (770 and 1170 K) heat treatment conditions was studied by differential thermal and X-ray powder diffraction analyses. While heating the reaction mixture, the exothermic events in the interaction between elements occur within the temperature ranges 620–750 K (Pr), 580–870 K (Sm), and 620–920 K (Er). X-ray powder diffraction analysis detected the formation of the following phases in the homogeneous state: at 770 K, LnSe2 and Yb2Se3, and at 1170 K, Ln3Se4, Ln2Se3, and YbSe. Scanning electron microscopy showed that particles of the selenide phases are formed by 1 × 2–5-μm fragments, which constitute a spongy structure of agglomerates tens to hundreds of micrometers in size.  相似文献   

11.
Supercritical fluid extraction (SFE) of the volatile oil from Santolina chamaecyparissus L. flower heads was performed under different conditions of pressure, temperature, mean particle size and CO2 flow rate. This oil was compared with the essential oil isolated by hydrodistillation (HD). The SFE volatile and essential oils were analysed by GC and GC‐MS. The range of the main volatile components obtained with HD and SFE were, respectively: 1,8‐cineole (25–30% and 7–48%), camphor (7–9% and 8–14%), borneol (7–8% and 2–11%), terpinen‐4‐ol (6–7% and 1–4%), terpinolene (1–4% and 1–7%) and isobornyl acetate (1–2% and 1–11%). The chemical composition of the extracts was greatly influenced by the conditions of pressure and temperature used. In fact, it was possible to enrich the sesquiterpene fraction by increasing the pressure from 8 to 9 MPa, while changing the temperature from 40 to 50°C at 90 bar enriched of the volatiles in n‐alkanes.  相似文献   

12.
The reactions of 1-aryl-2,2,2-trifluoroethanone oximes with tetrasulfur tetranitride (S4N4) in toluene at reflux gave 5-aryl-5-trifluoromethyl-4H-1,3,2,4,6-dithiatriazines 2 , 1-aryl-2,2,2-trifluoroethanonyliden-aminosulfenamides 3 and sulfur in 0–37%, 7–53%, and 2–41% yields, respectively. Treatment of 2 with tributyltin hydride in the presence of azobisisobutyronitrile in benzene at 80° afforded 3 in excellent yields.  相似文献   

13.
Copoly(4,4′-oxanilideterephthalamide—4,4′-phenyleneterephthalamide) (A-202/PPD) was synthesized by reaction of 4,4′-diaminooxanilide, p-phenylenediamine, and terephthaloyl chloride in organic solvents. Copolymer inherent viscosities in H2SO4 as high as 10.3 were obtained. Isotropic copolymer solutions (4%—5% concentration) of A-202/40%–80% PPD were spun to fibers with tenacity/elongation/modulus at 1% extension in the 13–14 gpd/1.5%–2%/700–1000 gpd range. Oxamide and amide stabilities in 98–100% H2SO4 and 20% oleum were compared. Poly(4,4′-oxanilideterephthalamide) (A-202), A-202/PPD copolymers, and poly(4,4′-phenyleneterephthalamide) (PPT) were unstable in 20% oleum, but all proved relatively stable in 100% H2SO4. However, the oxamide linkage proved less stable than the amide linkage in 98% H2SO4. A-202 and A-202/PPD copolymers formed stable anisotropic spinning solutions in 1% oleum at 10–20% concentrations. Dynamic mechanical analyses (Vibron) showed no glass transition temperature (Tg) below 200°C. Dilatometric measurement of A-202/50% PPD revealed a Tg at 257°C. Differential thermal analyses of A-202/40–80% PPD exhibited endotherms at 470–480°C. Thermogravimetric analyses showed no significant weight loss below 400°C.  相似文献   

14.
MgO-MgAl2O4 nanocomposite was prepared from the co-precipitation of Mg(NO3)2 and Al(NO3)3 salts, characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), energy-dispersive X-ray spectroscopy (EDS) and fourier transform infrared spectroscopy (FTIR) techniques and evaluated in the synthesis of thirty five derivatives of benzo[4,5]thiazolo[3,2-a]chromeno[4,3-d]pyrimidin-6-ones (d1-d34) via the multi-component reaction of 4-hydroxycoumarins, aldehydes, and 2-aminobenzothiazole derivatives under solvent free condition. The catalytic activity of MgO-MgAl2O4 nanocomposite and the synthesis of the above mentioned compounds were investigated under thermal solvent free (times: 1.4–3 h; yields: 75–95%), ultrasonic irradiation (US) conditions (times: 1–2.5 h; yields: 69–97%) and using high-speed ball milling (HSBM) technique (times: 0.7–2.5 h; yields: 67–97%). In all cases, the products were obtained in excellent yields. Nuclear Magnetic Resonance (NMR) and MASS spectroscopy were used to characterize the structure of the desired product. The mechanism for the preparation of compounds d1-d34 was proposed and confirmed by 1H NMR investigations.  相似文献   

15.
The reactions of S4N4 with phosphines of the type PhPR2 ( 1–5 ) and PhR′PR [ 6–11 (R′ = dicyclohexylamino), R: 1,6 = pyrrolidino; 2,7 = piperidino; 3,8 = morpholino; 4,9 = N-methylpiperazino; 5,10 = hexamethylenimino; and 11 = anilino] afford the phosphiniminocyclotrithiazene derivatives, 12–22 at room temperature. Only in the case of the phosphine 10 was the disubstituted derivative 23 isolated (in 65% yield). The trithiazene derivatives of the chiral phosphines in refluxing CH3CN produce the acyclic compounds →PN–S3N, 24–27 in ca. 60% yield. Norbornadiene reacts at room temperature with the cyclotrithiazene derivatives to give the addition products 28–35 . →P=N–S3N3 derivatives are found to be stable in 2M NaOH. © 1997 John Wiley & Sons, Inc. Heteroatom Chem 8 : 225–232, 1997.  相似文献   

16.
Diacetone Alcohol as Complex Ligand. Crystal Structures of [MnBr2{O=C(Me)CH2–C(Me)2OH}2] and [M{O=C(Me)CH2–C(Me)2OH}2][MCl4] with M = Fe, Co, and Zn The metal halides MnBr2 and MCl2 (M = Fe, Co, Zn) react with diacetone alcohol (4-hydroxy-4-methyl-2-pentanon) forming the title compounds, which are characterized by IR spectroscopy and crystal structure analyses. [MnBr2{O=C(Me)CH2–C(Me2)OH}2] ( 1 ): Space group C2/c, Z = 4, lattice dimensions at 293 K: a = 1189.2(4), b = 1317.2(3), c = 1200.0(3) pm, β = 102.25(3)°, R1 = 0.0256. In 1 the manganese atom is coordinated in a distorted octahedral fashion by the two cis bromine atoms and by the four oxygen atoms of the two diacetone alcohol chelating molecules. The distances Mn–[OH] (223.8 pm) and Mn–[O=C] (222.1 pm) are only slightly different. [M{O=C(Me)CH2–C(Me)2OH}2][MCl4] [M = Fe ( 2 ), Co ( 3 ), Zn ( 4 )]: 2 and 3 crystallize isotypically with each other in the space group Pc, Z = 4. Lattice dimensions for 2 at 293 K: a = 865.8(3), b = 926.3(2), c = 1401.5(1) pm, β = 104.19(2)°, R1 = 0.0421. Lattice dimensions for 3 at 293 K: a = 872.3(1), b = 925.7(1), c = 1394.2(3) pm, β = 104.79(2)°, R1 = 0.0481. As in 1 , the metal atoms of the [M{O=C(Me)CH2–C(Me)2OH}2]2+ ions in 2 and 3 are chelated in a distorted octahedral fashion by two diacetone alcohol molecules and associated cis via two μ-Cl atoms of the [MCl4]2– anions to form strands. [Zn{O=C(Me)CH2–C(Me)2OH}2][ZnCl4] ( 4 ): Space group C2/c, Z = 4. Lattice dimensions at 213 K: a = 1582.27(13), b = 1356.15(13), c = 941.93(7) pm, β = 107.283(10)°, R1 = 0.0328. The zinc atom of the dication in 4 is associated in a distorted octahedral fashion by the two diacetone alcohol chelating molecules in the equatorial positions and trans by two μ-Cl atoms of the [ZnCl4]2– ions to form strands.  相似文献   

17.
On the Oxidative Addition of 1-Halogenalk-1-ynes – Synthesis and Structure of Phenylalkynylpalladium Complexes [Pd(PPh3)4] ( 2 ) reacts with IC≡CPh and ClC≡CPh in the sense of an oxidative addition to give trans-[Pd(C≡CPh)X(PPh3)2] (X = I: 3 a , X = Cl: 3 b ). As side products trans-[PdX2(PPh3)2] (X = I: 4 a , X = Cl: 4 b ; < 10%) and PhC≡C–C≡CPh ( 5 ; X = I: ca 30%, X = Cl: < 4%) are formed. 3 a and 3 b were characterized by NMR (1H, 13C, 31P) and IR spectroscopies as well as by X-ray single-crystal structure analyses. In the crystals of 3 a and 3 b isolated molecules were found. The Pd–C≡C–Ph unit is linear in 3 a and approximately linear in 3 b [Pd–C≡C 174.2(6)°, C≡C–C 179,0(7)°].  相似文献   

18.
Five novel fluorene‐containing polymers, poly[(9,9‐dimethylfluoren‐2‐yl)acetylene] ( PFA1 ), poly[(1‐pentyl‐2‐(9,9‐dimethylfluoren‐2‐yl)acetylene) ( PFA2 ), poly[1‐decyl‐2‐(9,9‐dimethylfluoren‐2‐yl)acetylene] ( PFA3 ), poly[1‐phenyl‐2‐(9,9‐dimethylfluoren‐2‐yl)acetylene] ( PFA4 ), and poly[1‐(3,4‐difluorophenyl)‐2‐(9,9‐dimethylfluoren‐2‐yl)acetylene] ( PFA5 ) were synthesized by the polymerization of the corresponding fluorene‐substituted acetylenic monomers ( M1–M5), using WCl6, MoCl5, and TaCl5 as catalysts and n‐Bu4Sn as a cocatalyst. The synthesized polymers were thermally stable and readily soluble in common organic solvents. The degradation temperatures for a 5% weight loss of the polymers were ∼352–503 °C under nitrogen. PFA1–PFA5 show emission peaks from 402 to 590 nm. Besides, their electroluminescent properties were studied in heterostructure light‐emitting diodes (LEDs), using PFA2–PFA5 as an emitting layer. The PFA5 device revealed an orange‐red emission peak at 602 nm with a maximum luminescence of 923 cd/m2 at 8 V. A device with the ITO/PEDOT/ a mixture of PFA2 (98 wt %) and PFA5 (2 wt %)/Ca/Al showed near white emission. Its maximum luminance and current efficiency are 450 cd/m2 at 15 V and 1.3 cd/A, respectively. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 519–531, 2006  相似文献   

19.
Phosphoraneiminato‐Acetato Complexes of Cobalt and Cadmium with M4N4 Heterocubane Structure The phosphoraneiminato‐acetato complexes [M(NPEt3)(O2C–CH3)]4 with M = Co and Cd are formed from the anhydrous metal(II) acetates with excess Me3SiNPEt3 at 180 °C. By crystallization from diethyl ether blue, moisture sensitive single crystals of [Co(NPEt3) · (O2C–CH3)]4 can be obtained, while colourless single crystals of [Cd(NPEt3)(O2C–CH3)]4 · 2 CH2Cl2 originate from dichloromethane solution. In vacuo the intercalary CH2Cl2 is released. The complexes are characterized by their IR spectra and by crystal structure analyses. In both complexes the metal atoms are associated via μ3–N bridges of the (NPEt3) groups to form heterocubanes. In the cobalt complex the acetato ligands are bonded in a semichelate fashion with a short Co–O and a long Co–O bond each (Co–O distances in average 199.5 and 257.4 pm). In the cadmium complex the acetato groups form almost symmetrical chelates (Cd–O distances in average 232.1 and 237.8 pm); this leads to a distorted trigonal‐bipyramidal arrangement at the cadmium atoms. [Co(NPEt3)(O2C–CH3)]4: Space group P 1, Z = 4, lattice dimensions at –60 °C: a = 1110.1(2), b = 2051.3(5), c = 2169.5(4) pm, α = 100.03(2)°, β = 103.404(15)°, γ = 97.63(2)°, R = 0.0480. [Cd(NPEt3)(O2C–CH3)]4 · 2 CH2Cl2: Space group C2/c, Z = 4, lattice dimensions at –80 °C: a = 1550.2(1), b = 2101.1(1), c = 1706.1(1) pm, β = 91.09(1)°, R = 0.0311.  相似文献   

20.
The conformational equilibria of 1-phenyl-1-silacyclohexane 1, 3-phenyl-1,3-thiasilacyclohexane 2, 1-methyl-1-phenyl-1-silacyclohexane 3, and 3-methyl-3-phenyl-1,3-thiasilacyclohexane 4 have been studied for the first time by low temperature 13C NMR spectroscopy at 103 K. Predominance of the equatorial conformer of compound 1 (Pheq/Phax=78%:22%) is much less than in its carbon analog, phenylcyclohexane (nearly 100% of Pheq). And in contrast to 1-methyl-1-phenylcyclohexane, the conformers with the equatorial Ph group are predominant for compounds 3 and 4: at 103 K, Pheq/Phax ratios are 63%:37% (3) and 68%:32% (4). As the Si–C bonds are elongated with respect to C–C bonds, the barriers to ring inversion are only between 5.2–6.0 (ax→eq) and 5.4–6.0 (eq→ax) kcal mol?1. Parallel calculations at the DFT and MP2 level of theory (as well as the G2 calculations for compound 1) show qualitative agreement with the experiment. The additivity/nonadditivity of conformational energies of substituents on cyclohexane and silacyclohexane derivatives is analyzed. The geminally disubstituted cyclohexanes containing a phenyl group show large deviations from additivity, whereas in 1-methyl-1-phenyl-1-silacyclohexane and 3-methyl-3-phenyl-1,3-thiasilacyclohexane the effects of the methyl and phenyl groups are almost additive. The reasons for the different conformational preferences in carbocyclic and heterocyclic compounds are analyzed using the homodesmotic reactions approach.  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