首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 487 毫秒
1.
(±)-[1-hydro-8H-HDP]cobalt(I) 1
  • 1 Full name of 1: [2,2,3,3,7,7,8,8,12,12,13,13,17,17,18,18,-hexadecamethyl-2,3,7,8,12,13,17,18,-octahydro-1H,23H-10,20-diaza-porphinato]cobalt(I); full name of 2a : dibromo[2,2,3,3,7,7,8,8,12,12,13,13,17,17,18,18-hexadecamethyl-2,3,7,8,12,13,17,18-octahydro-1H,21H-10,20-diazaporphinato]cobalt(III).
  • 2 For the nomenclature of [HDP]-complexes see addendum in [2].
  • is obtained by chemical or electrochemical four-electron reduction of (±)-dibromo- or (±)-dicano[1-hydroxy-8H-HDP]cobalt(III) 2a or 2b 4, respectively. The crystal nad molecular structure of 1 was determined by combination of X-ray analysis and MS, 1H-, and 13C-NMR spectroscopy. Square-planar coordinated Co(I) lies closely to the best plane through the four N-atoms which form the first coordination sphere. Thermodynamic data for the coordination of axial bases with the cation of [1-hydroxy-8H-HDP]cobalt 2 in its different metal oxidation states were determined. The pathway of the overall four-electron reduction of 2a to 1 was elucidated: it involves a two-electron reduction of the central metal, a two-electron reduction of the macrocycle accompanied by elimination of the OH-group and final protonation at C(1). Evidence for an intramolecular electron transfer between the central metal and the macrocycle is presented.  相似文献   

    2.
    Synthesis and Structure of Porphine-type Nickel Complexes Containing Axially Oriented Sidechains The structure of (±)-[1, 11-dimethoxy-10 H-HDP (2-)]nickel ( 3 ), the product of the thermodynamically controlled addition of methanol to [6 H-HDP]nickel-bis-(tetrafluoroborate) ( 1 ), was determined by X-ray analysis. The two methoxy groups in 3 are cis-oriented. The syntheses and spectroscopic properties of [1, 11-diethoxy-10 H-HDP (2-)]nickel ( 5 ), [1, 11-dineopentyloxy-10 H-HDP (2-)]nickel ( 6 ) as well as the bridged complexes [1, 11-pentamethylenedioxy-10 H-HDP (2-)]nickel ( 4 ) and [1, 11-(E)-2,3-(dimethyl-2-butenylenedioxy)-10 H-HDP (2-)]nickel ( 7 ) are described. Reaction of 1 with bromide ions or 4-methylpyridine leads to the formation of the corresponding hexacoordinated, paramagnetic complexes dibromo [6 H-HDP]-nickel (II) ( 8 ) and bis (4-methylpyridine)[6 H-HDP]nickel-bis (tetrafluoroborate) ( 9 ).  相似文献   

    3.
    The deuterioformylation of (Z)- or (E)-2-butene catalyzed by [DIOP]Pt(SnCl3)-Cl
  • 1 DIOP=2,3-O-isopropylidene-2,3-dihydroxy-1,4-bis(diphenylphosphino)butane.
  • gives predominantly erythro- or threo-1,3-[2H]2-2-methylbutanal respectively. Hence, hydroformylation by this catalytic system must take place with cis-stereochemistry.  相似文献   

    4.
    The reversible isomerization of cis-hepta-1,3-diene to cis-2-trans-4-heptadiene via a 1,5 hydrogen shift has been investigated kinetically at nine temperatures in the range of 475° to 531°K. Equilibrium is reached near 94% reaction. Some cis-2-cis-4-heptadiene is also formed, but at a rate some 60 times slower than the cis,trans isomer. A least-squares analysis of the data yielded the Arrhenius equation for the isomerization of the cis-hepta-1,3-diene: Possible errors in the equilibrium constant measurements are discussed, and employing an equilibrium constant calculated by using group additivity estimates together with the values of k1, we obtained for the reverse reaction where .  相似文献   

    5.
    Equilibrium constants for the reaction CH3COCH2CH3 + I2 ? CH3COCHICH3 + HI have been computed to fit the kinetics of the reaction of iodine atoms with methyl ethyl ketone. From a calculated value of S(CH3COCHICH3) = 93.9 ± 1.0 gibbs/mole and the experimental equilibrium constants, ΔH(CH3COCHICH3) is found to be ?38.2 ± 0.6 kcal/mole. The Δ(ΔH) value on substitution of a hydrogen atom by an iodine atom in the title compound is compared with that for isopropyl iodide. The relative instability of 2-iodo-3-butanone (3.4 kcal/mole) is presented as further evidence for intramolecular coulombic interaction between partial charges in polar molecules. The unimolecular decomposition of 2-iodo-3-butanone to methyl vinyl ketone and hydrogen iodide was also measured in the same system. This reaction is relatively slow compared to the formation of the above equilibrium. Rate constants for the reaction over the temperature range 281°–355°C fit the Arrhenius equation: where θ = 2.303RT kcal/mole. The stability of both the ground and transition states is discussed in comparing this activation energy with that reported for the unimolecular elimination of hydrogen iodide from other secondary iodides. The kinetics of the reaction of hydrogen iodide with methyl vinyl ketone were also measured. The addition of HI to the double bond is not rate controlling, but it may be shown that the rate of formation of 1-iodo-3-butanone is more rapid than that for 2-iodo-3-butanone. Both four- and six-center transition complexes and iodine atom-catalyzed addition are discussed in analyzing the relative rates.  相似文献   

    6.
    Secondary hydrogen-deuterium isotope effects have been observed in the mass spectra of cis-4-t-butylcyclohexyl iodide, 5-iodononane and 2-iodopropane. Under conditions which suppress competing and second generation fragmentations, β-deuterium substitution decreases the intensity ratio \documentclass{article}\pagestyle{empty}\begin{document}$ ([{\rm M} - {\rm I]}^{\rm + } /[{\rm M]}^{\mathop + \limits_ \cdot } ) $\end{document}, a result analogous to a normal isotope effect. The decrease is larger in the spectrum of cis-4-t-butylcyclohexyl iodide-trans-2-d than in the spectrum of the cis-2-d derivative. Since these effects parallel those in the better understood solvolysis reaction, both effects may have a common origin. In contrast, deuteration of more remote positions in cis-4-t-butylcyclohexyl iodide and 5-iodononane increases the indicated intensity ratio, an apparent inverse isotope effect. Although similar effects have been observed in solvolysis reactions, the mass spectral effect may be attributable to an increase in the nonfixed energy available for fragmentation. These results suggest that secondary isotope effects can be readily measured in certain cases, and that they may eventually become useful probes into the mechanisms of mass spectral fragmentations.  相似文献   

    7.
    Supported Organometallic Complexes. VI. Characterization und Reactivity of Polysiloxane-Bound (Ether-phosphane)ruthenium(II) Complexes The ligands PhP(R)CH2D [R = (CH3O)3Si(CH2)3; D = CH2OCH3 ( 1b ); D = tetrahydrofuryl ( 1c ); D = 1,4-dioxanyl ( 1d )] have been used to synthesize (ether-phosphane)ruthenium(II) complexes, which have been copolymerized with Si(OEt)4 to yield polysiloxane-bound complexes. The monomers cis,cis,trans-Cl2Ru(CO)2(P ~ O)2 ( 3b ) and HRuCl(CO)(P ~ O)3 ( 5b ) were treated with NaBH4 to form cis,cis,trans-H2Ru(CO)2(P ~ O)2 ( 4b ) and H2Ru(CO)(P ~ O)3 ( 6b ), respectively (P ~ O = η1-P coordinated; = η2- coordinated). Addition of Si(OEt)4 and water leads to a base catalyzed hydrolysis of the silicon alkoxy-functions and a precipitation of the immobilized counterparts 4b ′, 6b ′. The polysiloxane matrix resulting by this new sol gel route has been described under quantitative aspects by 29Si CP-MAS NMR spectroscopy. 4b ′ reacts with carbon monoxide to form Ru(CO)3(P ~ O)2 ( 7b ′). Chelated polysiloxane-bound complexes Cl2Ru( )2 ( 9c ′, d ′) and Cl2Ru( )(P ~ O)2 ( 10b ′, c ′) have been synthesized by the reaction of 1b–c with Cl2Ru(PPh3)3 ( 8 ) followed by a copolymerization with Si(OEt)4. The polysiloxane-bound complexes 9c ′, d ′ and 10b ′, c ′ react with one equivalent of CO to give Cl2Ru(CO)( )(P ~ O) ( 12b ′– d ′). Excess CO leads to the all-trans-complexes Cl2Ru(CO)2(P ~ O)2 ( 14b ′– d ′), which are thermally isomerized to cis,cis,trans- 3b ′– d ′. The chemical shift anisotropy of 31P in crystalline Cl2Ru( )2 ( 9a , R = Ph, D = CH2OCH3) has been compared with polysiloxane-bound 9d ′ indicating a non-rigid behavior of the complexes in the matrix.  相似文献   

    8.
    The synthesis of racemic stereoisomeric compounds with the 5,5,9-trimethyldecalin skeleton and an oxygen function at C(1), C(2), or C(3) is described
  • 1 Although racemic decalins are described, only the enantiomer related to steroids is drawn. The projection of the decalins was chosen so as to place the angular methyl group above the plane of the molecule and the oxygen function at C(1), C(2) or C(3) on the left side, as represented by formula 1–6 . The relative configuration of the substituents in decalins is designated by using the convention of the steroid series: β, meaning on the same side as the angular methyl group at C(9) and α, meaning on the side opposite from the angular methyl group. The prefix cis or trans refers to the fusion of the decalin ring system, not to the position of the substituents.
  • . A novel general one-step synthesis of 2-decalones by means of acid catalyzed cyclization of acyclic or monocyclic precursors has been developed.  相似文献   

    9.
    The angular alkylation of cis-decalin epoxides like 5 or 7 can be achieved at C(8a)
  • 1 For convenience, the arbitrary numbering given for 5 (Scheme I) is used throughout the General Part; for systematic names, see Exper. Part.
  • in good yield by using CuI and a large excess of Grignard reagents without an sp3 centre at C(2). The reaction proceeds via a carbenium-ion intermediate which is stabilised by homoconjugative interaction with the adjacent double bond. Due to 1,3-diaxial strain in the alkoxides resulting from alkylation or reduction at C(4a) of the epoxides 5 or 7 , the nucleophile is delivered selectively to C(8a). Grignard reagents possessing H-atoms at C(β), transfer a hydride to the epoxide yielding the trans-decalol 11 (Grignard reduction). The angular alkylation of 5 with allylic and benzylic Grignard reagents proceeds with good yield.  相似文献   

    10.
    A diastereoselective synthesis of 1,2,3,3a,4,5‐hexahydropyrrolo[1,2‐a]quinoline‐5‐carboxylic esters has been developed using a tandem reduction‐double reductive amination reaction. The nitro dicarbonyl cyclization substrates were synthesized by alkylation of methyl (2‐nitrophenyl)acetate with 2‐bromomethyl‐1,5‐hexadiene derivatives, followed by ozonolysis. Catalytic hydrogenation of each substrate gave the target heterocycle, along with a deacylated product and an adduct resulting from capture of the intermediate hydroxylamine by the side chain carbonyls. The product ratio varied dramatically with the catalyst and the hydrogen pressure. Cyclization to the title compounds was highly diastereoselective, producing each hexahydropyrrolo[1,2‐a]‐quinoline as a single stereoisomer with the all‐cis geometry. The competing processes have not been observed in previous heterocyclization studies but can be attributed to greater strain in the system, which slows the final ring closure.  相似文献   

    11.
    The thermal reaction of hydrogen–butene-2-cis mixtures has been studied in a static system at low extent of reaction around 500°C. Hydrogen does not affect the thermal reaction itself of the olefin, but gives rise to new stoichiometries of hydrogenolysis and hydrogenation, which are specified: The reaction is described in terms of a molecular and free-radical mechanism. It is shown that the key process for the hydrogenolysis–hydrogenation reaction is (1) and that the rate constant of this process can be determined from either propylene, or methane, or butene-1 formations: with θ = 4.57 × 10?3 T kcal/mol. Other rate constants are estimated and agree with literature data.  相似文献   

    12.
    Tetramethyl 3a,7a‐dihydro‐1‐methyl‐1H‐indole‐2,3,3a,4‐tetracarboxylate which is an 1:2 adduct of 1‐methylpyrrole and dimethyl acetylenedicarboxylate underwent isomerization catalyzed by sodium methoxide to form a 5,7a‐dihydro‐1‐methyl‐1H‐indole‐2,3,4,5‐tetracarboxylate, its 5,6‐dihydro isomer, and a ring opening product which is an azonine derivative. Fully aromatized esters such as 1‐methylindole‐2,3,4‐triester, 1‐methylindole‐2,3,4,5‐tetraester and 1‐methyl‐2,3,4,6‐tetraester were also isolated. An indole compound which could be formed by conjugative addition of the methoxide ion was also isolated.  相似文献   

    13.
    The title reaction has been investigated in the temperature range 667–715K. The only reaction products were trifluorosilyl iodide and hydrogen iodide. The rate law was obeyed over a wide range of iodine and trifluorosilane pressures. This expression is consistent with an iodine atom abstraction mechanism and for the step log k1(dm3/mol·sec) = (11.54 ± 0.17) ? (130.5 ± 2.2 kJ/mol)/RT In 10 has been deduced. From this the bond dissociation energy D(F3Si? H) = (419 ± 5) kJ/mol (100.1 kcal/mol) is obtained. The kinetic andthermochemical implications of this value are discussed.  相似文献   

    14.
    Chemical Transport in the System Mn? O in Consideration of the Oxygen Coexistence Pressure (I) The chemical transport of the coexistent phases Mn2O3? Mn3O4 and Mn3O4? MnO with Cl2, Br2, I2, HCl, HBr, and HI was analysed thermodynamically and experimentally. The mentioned transport agents are able to transport the following phases:
  • 1 Index (o) bedeutet obere, (u) untere Phasengrenze (index (o) – upper phase boundary, (u) – lower phase boundary).
  • .  相似文献   

    15.
    Detection of the adduct radical by ESR spectroscopy and after‐effect ESR measurements of the adduct radical concentrations in the photosensitized polymerization of styrene (St) in the presence of dimers of α‐methylstyrene (MSD) and methyl methacrylate have revealed that the dominant mechanism of adduct radical loss changes from bimolecular termination to fragmentation as the temperature is increased beyond 90 °C for St/MSD.

      相似文献   


    16.
    The cycloadditions of methyl propynoate and methyl vinyl ketone to 5,6-dimethylidene-2-norbornanone ( 6 ) are para′-regioselective
  • 1 “Para” (p) designs in this paper the 4, 9-disubstituted tricclo[6.2.1.02, 7] undecane- and 4, 9-disubsstituted tricycle[6.2.202,7] dodecane derivatives, “meta” (m) design the corresponding 4, 10-disubstituted compounds.
  • . A smaller para -regioselectivity is observed for the addition of methyl propynoate to 5,6-dimethylidene-2bicyclo[2.2.2]octanone ( 10 ). No regioselectivity is observed with 5,6-dimethylidene-2exo-norbornyl alcohol ( 3 ), acetate ( 5 ) and 5,6-dimethylidene-2exo-bicyclo[2.2.2]octanol ( 9 ). PMO arguments based on the shape of the HOMO's and subHOMO's of the dienes allow to rationalize these observations. Unpredictable para′- or ‘meta’regioselectivities are found for the Diels,-Alder additions of 5,6-dimethylidene-2endo-norbornyl alcohol ( 2 ), acetate ( 4 ) and 5,6-dimethylidene-2endo-bicyclo[2.2.2]octanol ( 8 ). The carbonyl group of β,γ unsaturated ketones such as 6 and 10 can act as an electron donating homoconjugated substituent. The n(CO) ? σ[C(1), C(2)] ? π[C(5), C(6)] hyperconjugative interaction can override the usual electron-withdrawing effect of this function.  相似文献   

    17.
    Summary: Well‐defined flexible, annealed, and quenched polyelectrolytes with two cationic sites per monomeric unit are synthesized. The synthetic scheme involves the synthesis of narrowly distributed poly(ptert‐butoxystyrene) precursors by anionic polymerization high vacuum techniques, hydrolysis to poly(p‐hydroxystyrene), and subsequent quantitative functionalization by a Mannich‐type reaction, to yield annealed polyelectrolytes with two dimethylamino groups per monomer. In a last step, the dimethylamino groups are converted into quaternary ammonium salts by reaction with methyl iodide to give high‐charge‐density quenched cationic polyelectrolytes. The polymers are molecularly characterized by NMR and FT‐IR spectroscopy, while their solution behavior is studied by potentiometric titrations, turbidimetry, and fluorescence spectroscopy as a function of pH, in the case of the annealed polyelectrolytes, as well as by viscometry in the case of the quenched polyelectrolytes.

    Schematic diagram of the synthesis of the polyelectrolytes.  相似文献   


    18.
    The thermal reaction of trans‐1‐methyl‐2,3‐diphenylaziridine (trans‐ 1a ) with aromatic and cycloaliphatic thioketones 2 in boiling toluene yielded the corresponding cis‐2,4‐diphenyl‐1,3‐thiazolidines cis‐ 4 via conrotatory ring opening of trans‐ 1a and a concerted [2+3]‐cycloaddition of the intermediate (E,E)‐configured azomethine ylide 3a (Scheme 1). The analogous reaction of cis‐ 1a with dimethyl acetylenedicarboxylate ( 5 ) gave dimethyl trans‐2,5‐dihydro‐1‐methyl‐2,5‐diphenylpyrrole‐3,4‐dicarboxylate (trans‐ 6 ) in accord with orbital‐symmetry‐controlled reactions (Scheme 2). On the other hand, the reactions of cis‐ 1a and trans‐ 1a with dimethyl dicyanofumarate ( 7a ), as well as that of cis‐ 1a and dimethyl dicyanomaleate ( 7b ), led to mixtures of the same two stereoisomeric dimethyl 3,4‐dicyano‐1‐methyl‐2,5‐diphenylpyrrolidine‐3,4‐dicarboxylates 8a and 8b (Scheme 3). This result has to be explained via a stepwise reaction mechanism, in which the intermediate zwitterions 11a and 11b equilibrate (Scheme 6). In contrast, cis‐1,2,3‐triphenylaziridine (cis‐ 1b ) and 7a gave only one stereoisomeric pyrrolidine‐3,4‐dicarboxylate 10 , with the configuration expected on the basis of orbital‐symmetry control, i.e., via concerted reaction steps (Scheme 10). The configuration of 8a and 10 , as well as that of a derivative of 8b , were established by X‐ray crystallography.  相似文献   

    19.
    Treatment of 3,5-dimethoxy-1,2,4-triazine ( 1a ) with methyl iodide was found to give depending on the reaction time triazinium iodide 2a , triaziniumolates 4a and 6a as well as methoxytriazinones 7a and 8a . Thermolysis of 2a gave triaziniumolates 4a and 6a . Reaction of 2a , 4a or methoxytriazinone 9a with methyl iodide in acetonitrile yielded as the sole product 6a . Reaction of 3-methoxy-5-methylthio-1,2,4-tri-azine (1b ) with methyl iodide gave triazinium iodide 2b and methylthio triazinone 7b . Hydrolysis of 2a,b afforded 4a . Reaction of 5-methoxy-3-methylthio-1,2,4-triazine ( 1c ) with methyl iodide gave triazinium iodide 2c , triaziniumolate 4b , triazinium iodide 5b and triazinone 8b . Hydrolysis of 2c yielded 4b and its thermolysis gave a mixture of 4b and 5b . Reaction of 2c , 4b and triazinone 9b with methyl iodide afforded 5b . Treatment of 3,5-bis(methylthio)-1,2,4-triazine ( 1d ) with methyl iodide was found to give a mixture of N1 and N2 methiodides 2d and 3d which gave on hydrolysis 4b and 8b , respectively. Methylation of 6-methyl derivatives 1c-g gave analogous results, however the proportions of N1 methylated products were lower and the reaction rates higher in comparison to their respective lower homologues 1a,c,d . The structures of the mesoionic dimethyl derivatives were assigned from uv, ir, 1H nmr and electron impact mass spectra. The structural assignments were eventually confirmed by quantum chemical calculations of net charge distributions, bond lengths and ipso angles of the C5?O bonds.  相似文献   

    20.
    Treatment of methyl propiolate and 2‐(thiophen‐2‐yl)benzaldehyde, 2‐(thiophen‐3‐yl)benzaldehyde or 2‐(furan‐3‐yl)benzaldehyde with tetrabutylammonium iodide/zirconium (IV) chloride or treatment of methyl acrylate and the same aldehydes with 1,4‐diazabicyclo[2,2,2]octane and triethanolamine induce an aldol‐type reaction to furnish Baylis‐Hillman adducts β‐iodo‐α‐(hydroxymethyl)acrylates or α‐(hydroxy‐methyl)acrylates, respectively. These can be used for the preparation of indenothiophenes and indenofurans having acrylic acid unit by intramolecular Friedel‐Crafts reaction with sulfuric acid in tetrachloromethane.  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-23

    京公网安备 11010802026262号