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 93 毫秒
1.
Mei-Hsiu Shih   《Tetrahedron》2002,58(52):10437-10445
3-Arylsydnone-4-carbohydroximic acid chlorides (1) could react with N-arylmaleimides (3a–b) or 2-methyl-N-phenylmale-imide (3c) to give 3-(3-arylsydnon-4-yl)-5-aryl-3a,6a-dihydro-pyrrolo[3,4-d]isoxazole-4,6-diones (4a–h) or 6a-methyl-3-(3-arylsydnon-4-yl)-5-phenyl-3a,6a-dihydro-pyrrolo[3,4-d]isoxazole-4,6-diones (4i–l), respectively. However, 3-(arylsydnon-4-yl)-naphtho[2,3-d]isoxazole-4,9-diones (6a–d) were obtained in good yield by the reaction of carbohydroximic acid chlorides 1 with [1,4]naphthoquinone. Furthermore, 2-(3-arylsydnon-4-yl)benzoxazoles (9a–d) and 2-(3-arylsydnon-4-yl)benzothiazoles (9e–h) were obtained via the reaction of carbohydroximic acid chlorides 1 with ortho-substituted aromatic amines 7a and b.  相似文献   

2.
De-Dong Wu  Thomas C. W. Mak 《Polyhedron》1994,13(24):3333-3339
Two polymeric mercury(II) halide adducts of an olefinic double betaine, cis-(p-Me2NC5H4N+)2C2(COO)2 (L), have been prepared and characterized by X-ray crystallography. [{Hg2L2Cl4·6HgCl2}n] (1) crystallizes in the monoclinic space group C2/c with Z = 4, and [{Hg2L2Br4·HgBr2}n] (2) in the triclinic space group P with Z = 1. Complexes 1 and 2 are structurally similar, being composed of centrosymmetric fourteen-membered rings and nearly linear HgX2 (X = Cl, Br) moieties that are further inter-linked by weak HgX [HgCl = 2.930–3.136(9) Å, HgBr = 3.057–3.310(6) Å] and HgO [2.64, 2.75(3) Å] bonds to generate a two-dimensional polymeric network.  相似文献   

3.
Reaction of Na[MCl4] (M=Pd or Pd) with the azo-containing phosphines Ph2P{1-(4-RC6H4N2)-2-OR′-C10H5} {R=Me (I), NMe2 (II); R′=C(O)Me} affords the complexes [MCl2L2] (1–4) in good yield. Complexes 1–4 have all been fully characterised by elemental analysis, 1H-, 13C{1H}-, and 31P{1H}-NMR spectroscopy and UV–visible spectroscopy. The use of 1 in the Heck reaction has been investigated and shown to effect up to 1000 turnovers.  相似文献   

4.
The reactions of the diruthenium carbonyl complexes [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]X (X=BF4 (1a) or PF6 (1b)) with neutral or anionic bidentate ligands (L,L) afford a series of the diruthenium bridging carbonyl complexes [Ru2(μ-dppm)2(μ-CO)22-(L,L))2]Xn ((L,L)=acetate (O2CMe), 2,2′-bipyridine (bpy), acetylacetonate (acac), 8-quinolinolate (quin); n=0, 1, 2). Apparently with coordination of the bidentate ligands, the bound acetate ligand of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ either migrates within the same complex or into a different one, or is simply replaced. The reaction of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ (1) with 2,2′-bipyridine produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)2] (2), [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-bpy)]+ (3), and [Ru2(μ-dppm)2(μ-CO)22-bpy)2]2+ (4). Alternatively compound 2 can be prepared from the reaction of 1a with MeCO2H–Et3N, while compound 4 can be obtained from the reaction of 3 with bpy. The reaction of 1b with acetylacetone–Et3N produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-acac)] (5) and [Ru2(μ-dppm)2(μ-CO)22-acac)2] (6). Compound 2 can also react with acetylacetone–Et3N to produce 6. Surprisingly [Ru2(μ-dppm)2(μ-CO)22-quin)2] (7) was obtained stereospecifically as the only one product from the reaction of 1b with 8-quinolinol–Et3N. The structure of 7 has been established by X-ray crystallography and found to adopt a cis geometry. Further, the stereospecific reaction is probably caused by the second-sphere π–π face-to-face stacking interactions between the phenyl rings of dppm and the electron-deficient six-membered ring moiety of the bound quinolinate (i.e. the N-included six-membered ring) in 7. The presence of such interactions is indeed supported by an observed charge-transfer band in a UV–vis spectrum.  相似文献   

5.
The bimetallic [Pt(NH3)4]2[W(CN)8][NO3]·2H2O is characterised by single-crystal X-ray diffraction [S.G.P21/m(11), a=8.0418(7), b=19.122(2), c=9.0812(6) Å, Z=2]. All platinum centres have the square-plane D4h geometry with average dimensions Pt(1)–N 2.042(2) and Pt(2)–N 2.037(10) Å. The octacyanotungstate anion has the square-antiprismatic D4d configuration with average dimensions W(1)–C 2.164(13), C–N 1.140(12), W(1)–N 3.303(5) Å. The structure exhibits two different mutual orientations of Pt versus W units resulting in Pt(2)–W(1), W(1)* separations of 4.77(2), 4.55(2)* and Pt(1)–W(1) of 6.331(8) Å. A centrosymmetric structure reveals groups of two distinct columns: the first is formed by intercalated NO3 between parallel [Pt(1)(NH3)4]2+ planes and the second consists of [W(CN)8]3− interlayered by, parallel to square faces of W-antiprisms, [Pt(2)(NH3)4]2+. The structure is stabilised through a three-dimensional hydrogen bond network via nitrogen atoms of cyanide ligands, hydrogen atoms of NH3 ligands, water molecules and oxygen atoms of NO3 counteranions. The vibrational pattern and the range of ν(CN) frequencies attributable to the electronic environment of W(V) and W(IV) are consistent with the ground state Pt(II)↔W(V) charge transfer.  相似文献   

6.
Reaction of the cationic complex [WI(CO)(NCMe){Ph2P(CH2)PPh2}(η2-MeC2ME)][BF4] with an equimolar amount of MX (MX = NaCl, NaBr, NaI, KNO2, KNO3, NaNCS or KOH) in acetone at room temperature gave the neutral complex [WIX(CO){Ph2P(CH2)PPh2}(η2-MeC2Me)] (1–7) in good yield. Complexes 1–7 have been characterized by elemental analysis (C, H and N), IR and 1H NMR spectroscopy.  相似文献   

7.
Upon UV irradiation in hexane at 243 K tricarbonyl-η5-cyclohexadienyl-manganese (1) and two equivalents of 2-butyne (2) or diphenylacetylene (4) yield in successive [5 + 2, 3 + 2] cycloadditions tricarbonyl-η2:2:1-1,2,3,10-tetramethyl-tricyclo[5.2.1.04,9]-deca-2,5-dien-10-yl-manganese (6), or tricarbonyl-η2:2:1-1,2,3,10-tetraphenyl-tricyclo[5.2.1.04,9]-deca-2,5-dien-10-yl-manganese (8), respectively. 3-Hexyne (3) reacts with 1 under the same conditions by successive [5 + 2, 3 + 2] cycloadditions and 1,4-H-shift to tricarbonyl-η2:2:1-1,2,3-triethyl-10-ethylidene-tricyclo[5.2.1.04,9]dec-2-en-5-yl-manganse (7). Identical products are also obtained when 1 is first irradiated in THF at 208 K and the thermolabile intermediate, dicarbonyl-η5-cyclohexadienyl-tetrahydrofurane-manganese (11), is treated with an excess of the alkynes 2–4. In contrast, bis(trimethylsily)acetylene (5) substitutes photochemically in 1 only a CO ligand to yield dicarbonyl-η5-cyclohexadienyl-η2-bis(trimethylsily)Acetylene-manganese (9). The crystal and molecular structure of 7 was determined by an X-ray diffraction analysis. Complex 7 crystallizes in the triclinic space group , a = 822.6(2) pm, B = 882.5(2) pm, C = 1344.6(2) pm, = 92.36(2)°, β = 107.13(2)°, γ = 99.71(2)°, V = 0.9152(3) nm3, Z = 2. The complexes 6–9 were studied in solution by IR and NMR spectroscopy. The structures of 6,8 and 9 were elucidated from the NMR spectra. A possible formation mechanism for the complexes 6–9 will be discussed.  相似文献   

8.
The reaction between metallic barium and fluoroisopropanol or alcoholysis of [Ba(OPri)2] produces a pentanuclear fluoroalkoxide. Its X-ray structure determination showed its formulation to correspond to Ba55-OH)[μ3-OCH(CF3)2]42-OCH(CF3)2]4 [OCH(CF3)2](THF)4(H2O)·THF. The metallic core is based on a square pyramid encapsulating an hydroxo ligand. In addition to the barium---alkoxide bonds [2.53(3)–2.86(3) Å] neutral O-donors, four THF [2.82(2)–2.86(3) Å] and one H2O [2.79(3) Å] and secondary barium---fluorine interactions [2.99(2)–3.31(2) Å] ensure high coordination numbers, from 9 to 11 for the metal centers. Hydrogen bonding between the apical fluoroisopropoxide, the water molecule and one THF molecule, non-bonded to a metal center, accounts for the stability of the hydrate and illustrates the Lewis acidity of fluoroalkoxides. Thermal decomposition leads to BaF2.  相似文献   

9.
-Benzotriazolylamides 6a–d afforded N-(benzotriazol-1-ylmethyl)arylimidoyl chlorides (4a–d), which reacted in situ with potassium tert-butoxide to form 3-aryl-1,2,4-triazolo[1,2-a]benzotriazoles (7a–d) (44–68%), representatives of a novel heterocyclic system. The structure of 7a was confirmed by single crystal X-ray analysis.  相似文献   

10.
The new diphenolato complexes [{Mo(NO){HB(dmpz)3}Cl}2Q] where dmpz = 3,5-dimethylpyrazolyl and Q = OC6H4(C6H4O (n = 1 or 2), OC6H4CR=CRC6H4O (R = H or Et), and OC6H4CH=CHC6H4CH=CHC6H4O have been prepared and their electrochemical properties (cyclic and differential pulse voltammetry) compared with previously reported analogues where Q = OC6H4O, OC6H4EC6H4O (E = SO2, CO and S), OC6H4 (CO)C6H4 C6H4(CO)C6H4O and 1,5- and 2,7-O2C10H6. The electrochemical interaction between the redox centres in the new complexes is very weak, in contrast to that in the 1,4-benzenediolato and naphthalendiolato species. The EPR spectra of the reduced mixed-valence species [{Mo(NO){HB(dmpz)3}Cl}2Q] where Q = 1,3- and 1,4-OC6H4O and OC6H4SC6H4O shows that they are valence-trapped at room temperature, whereas those of the dianions [{Mo(NO){HB(dmpz)3}Cl}2Q]2− where Q = 1,4-OC6H4O, OC6H4EC6H4O (E = CO or S) and OC6H4CH=CHC6H4CH=CHC6H4O shows that the unpaired spins on each molybdenum centre are strongly correlated (J, the spin exchange integral AMo, the metal-hyperfine coupling constant). The electrochemical properties and the comproportionation constants for the reaction [{Mo(NO){HB(dmpz)3} Cl}2Q] + [{Mo(NO){HB(dmpz)3}Cl}O]2]2−2[{Mo(NO) {HB(dmpz)3}Cl}2Q] where Q = diphenolato bridge, are compared with related compounds containing benzenediamido and dianilido bridges.  相似文献   

11.
The asymmetric total synthesis of (−)-dehydroclausenamide 1 was accomplished through a concise and efficient synthetic route employing as the key step the novel formation of cis-epoxy amide 7, which was resulted from the reaction of starting material 4 bearing an optically pure erythro vicinal diol unit with 5H-3-oxa-octafluoro pentanesulfonyl fluoride (HCF2CF2OCF2CF2SO2F) in the presence of 1,8-diazabicyclo[5.4.0]-7-undecene (DBU).  相似文献   

12.
The reaction of N-(3,4-dichlorophenethyl)-N-methylamine (1) with 3-chloromethyl-5-phenyl-1,2,4-oxadiazole (2) was investigated. Employment of an equimolar amount of 1 and 2 in the presence of potassium carbonate led to the expected tertiary amine 3 (N-[(3,4-dichlorophenyl)ethyl]-N-methyl-N-[(5-phenyl-1,2,4-oxadiazol-3-yl)methyl]amine), whereas an excess of 1 and prolonged reaction time resulted in ring fission of the oxadiazole system in 3 and finally in the formation of N′-benzoyl-N-[(3,4-dichlorophenyl)ethyl]-N-methylguanidine (4) and N,N′-bis[(3,4-dichlorophenyl)ethyl]-N,N′-dimethylmethanediamine (5). The structures of products 3–5 were determined by means of 1H and 13C NMR-spectroscopy, mass spectrometry and IR-spectroscopy, for 3 (as picrate) and 4 also X-ray structure analysis was employed. A possible mechanism of the reaction pathway leading to compounds 4 and 5 is proposed.  相似文献   

13.
Transamination reactions utilizing the compound mercuric bis(trimethylsilyl)amide, Hg{N(SiMe3)2}2, in tetrahydrofuran (THF), and the metals Na, Mg, Ca, Sr, Ba and Al have been investigated. Thus the THF solvated compounds Na[N(SiMe3)2]·THF and M[N(SiMe3)2]2·2THF, M = Mg, Ca, Sr and Ba (1–4), have been prepared. The X-ray crystal structures of 1 and the related manganese compound Mn[N(SiMe3)2]2·2THF (5) are reported. Interaction of the silylamides, 2–4, with a range of crown ethers apparently proceeded with elimination of silylamine, (Me3Si)2NH, and novel ring opening of the crown ethers, generating species containing a donor alkoxide ligand with a vinyl ether function, presumably, ---O(CH2CH2O)nCH=CH2 (n = 3−5). The silylamides 2–4 were also cleanly converted to the corresponding alkoxides (from 1H NMR data) in reactions with stoichiometric quantities of 3-ethyl-3-pentanol.  相似文献   

14.
The transformation of the benzanilides 1 into 4-arylisochroman-3-acetic acids 8 applying the following sequence of reactions is described. At first, the 3-arylphthalides 3 were obtained via metallation [n-BuLi] of benzanilides 1 and subsequent treatment of the generated bis-lithiated anilides 2 with aromatic aldehydes. Next, the 3-arylphthalides 3 were reduced [LiBH4] to phthalanes 5 and then, via reductive metallation [Li/C10H8] and reaction of the generated bis-lithiated species 6 with dimethylformamide, 3-hydroxy-4-arylisochromans 7 were produced. In the final step the isochromans 7 were treated with 1-methoxy-1-trimethylsilyloxyethene in the presence of titanium tetrachloride and furnished 4-arylisochromans-3-acetic acid methyl esters 8 as trans stereoisomers (Ψ-e/e).  相似文献   

15.
A study has been carried out of the catalytic activity of the systems formed by [HRh{P(OPh)3}4] or [HRh(CO){P(OPh)3}3] with the modifying ligands P(OPh)3, PPh3, diphos and Cp2Zr(CH2PPh2)2 in hydroformylation of hex-1-ene (at p = 5 bar). The best results were obtained with the system [HRh{P(OPh)3}4]+Cp2Zr(CH2PPh2)2 (75–85% yeild of aldehydes).  相似文献   

16.
Norfloxacin, 1-ethyl-6-fluoro-1,4-dihydro-4-oxo-7-(1-piperazinyl)-3-quinoline carboxylic acid (NORH), reacts with aluminium(III) ion forming the strongly fluorescent complex [Al(HNOR)]3+, in slightly acidic medium. The complex shows maximum emission at 440 nm with excitation at 320 nm. The fluorescence intensity is enhanced upon addition of 0.5% sodium dodecylsulphate. Fluorescence properties of the Al-NOR complex were used for the direct determination of trace amounts of NOR in serum. The linear dependence of fluorescence intensity on NOR concentration, at a NOR to Al concentration ratio of 1:10, was found in the concentration range 0.001–2 μg/ml NOR with a detection limit of 0.1 ng/ml. The ability of aluminium (III) ion to form complexes with NOR was investigated by titrations in 0.1 M LiCl medium, using a glass electrode, at 298 K, in the concentration range: 2 × 10−4 ≤ [Al] ≤ 8 × 10−4; 5 × 10−4 ≤ [NOR] ≤ 9 × 10−4 mol/dm3; 2.8 ≤ pH ≤ 8.3. The experimental data were explained by the following complexes and their respective stability constants, log(β ± σ): [Al(HNOR)], (14.60 ± 0.05); [Al(NOR)], (8.83 ± 0.08); [A1(OH)3(NOR)], (−14.9 ± 0.1), as well as several pure hydrolytic complexes of A13+. The structure of the [Al(HNOR)] complex is discussed, with respect to its fluorescence properties.  相似文献   

17.
The four-coordinate tin(II) complex [η4-Me8taa]Sn undergoes oxidative addition of I2 to give six-coordinate [η4-Me8taa]SnI2, in which the iodide ligands exhibit a trans arrangment. Abstraction of I from [η4-Me8taa]SnI2 is facile, as indicated by the rapid formation of the triiodide derivative *[η4-Me8taa]SnI(THF)**I3* upon treatment with I2 in the presence of THF. The molecular structures of [η4-Me8taa]SnI2 and *[η4-Me8taa]SnI(THF)**I3* have been determined by X-ray diffraction.  相似文献   

18.
Two carbon-rich starburst gold(I) acetylide complexes [TEE][Au(PCy3)]4 (3, [TEE]H4=tetraethynylethene) and [TEB][Au(PCy3)]3 (6, [TEB]H3=1,3,5-triethynylbenzene) were prepared and their UV–vis absorption, emission and excitation spectra have been recorded. In fluid CH2Cl2 solutions, 3 exhibits prompt 1(ππ*) fluorescence with λ0–0 and λmax at 413 and 428 nm, respectively, while 6 displays 3(ππ*) phosphorescence with λ0–0 and λmax at 446 and 479 nm, respectively. The crystal structure of 3·CH2Cl2 has been determined.  相似文献   

19.
Addition of 1,4-dithiols to dichloromethane solutions of [PtCl2(P-P)] (P-P = (PPh3)2, Ph2P(CH2)3PPh2, Phd2P(CH2)4PPh2; 1,4-dithiols = HS(CH2)4SH, (−)DIOSH2 (2,3-O-isopropylidene-1,4-dithiol-l-threitol), BINASH2 (1,1′-dinaphthalene-2,2′-dithiol)) in the presence of NEt3 yielded the mononuclear complexes [Pt(1,4-dithiolato)(P-P)]. Related palladium(II) complexes [Pd(dithiolato)(P-P)] (P-P=Ph2P(CH2)3PPh2, Ph2P(CH2)4PPh2; dithiolato = S(CH2)4S, (−)-DIOS) were prepared by the same method. The structure of [Pt((−)DIOS)(PPh3)2] and [Pd(S(CH2)4S)(Ph2P(CH2)3PPh2)] complexes was determined by X-ray diffraction methods. Pt—dithiolato—SnC12 systems are active in the hydroformylation of styrene. At 100 atm and 125°C [Pt(dithiolate)(P-P)]/SnCl2 (Pt:Sn = 20) systems provided aldehyde conversion up to 80%.  相似文献   

20.
The siloxyanilines o-Me3SiOC6H4NH2 (1) and p-RMe2SiOC6H4NH2 (R=H (2); R=Me (3)), and their N-silylated derivatives p-Me3SiOC6H4NHSiMe3 (4) and p-Me3SiOC6H4N(SiMe3)2 (5) have been prepared from ortho- or para-aminophenol and used in the synthesis of imido complexes. Thus, binuclear [{Ti(η5-C5H5)Cl}{μ-NC6H4(p-OSiMe3)}]2 (6) and mononuclear [TiCl2{NC6H4(p-OSiMe3)}(py)3] (7) imido complexes have been obtained from the reaction of 3 and [Ti(η5-C5H5)Cl3] or [TiCl2(NtBu)(py)3], respectively. In contrast, the reaction of 1 with TiCl4 and tBupy affords the titanocycle [TiCl2{OC6H4(o-NH)---N,O}(tBupy)2] (8). Compound 5 has also been used to prepare the niobium imide complex [NbCl3{NC6H4(p-OSiMe3)}(MeCN)2] (9), by its reaction with NbCl5 in CH3CN. These findings have been applied to the synthesis of polynuclear systems. Thus, chlorocarbosilane Si[CH2CH2CH2Si(Me)2Cl]4 (CS–Cl) has been functionalized with the ortho- and para-aminophenoxy groups to give 10 and 11, respectively. The use of 11 has allowed the formation of the tetranuclear compound 12. Attempts to synthesize terminal imido titanium complexes from 10 and TiCl4 in the presence of tBupy and Et3N, give complex 8 and carbosilane CS–Cl.  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