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 593 毫秒
1.
A comparative kinetic study of the reaction of three different hydroxylated liquid poly-butadienes (M?n ? 3000)–R-45M, R-45HT and H-034–with 3-isocyanatomethyl–3,5,5-trimethylclohexylisocyanate (IPDI) and dimer diacid diisocyanate (DDI) was carried out in toluene solution. An analytical method was used to follow the kinetics of the reactions, at four different temperatures. In the second-order plots, a discontinuity was observed in the reaction with IPDI, in contrast to the reactions with DDI which showed straight-line plots. In all studied reactions, the R-45M polybutadiene was about twice as reactive as R-45HT and H-034. The latter hydroxylated polybutadienes (R-45HT and H-034) showed similar reactivities.  相似文献   

2.
A new couloamperometric apparatus has been designed to extend the range of this kinetic technique to the measurement of very high rate constants, 108M?1s?1, by using TFCR-EXSEL conditions (TFCR—very low reactant concentration; EXSEL—salt excess), which give half-lives of a few seconds for very fast second-order reactions. Very low faradaic currents, in the nanoampere range for halogens, corresponding to very low reactant concentrations of 10?8–10?9M, are measured selectively by compensating the eddy currents, principally the residual and the induced currents. When the electroactive species is bromine, the concentration is demonstrated to be linearly related to the limiting reduction current in the very low concentration range. The upper limit of this technique for bromination is at present 3 × 108M?1s?1. The method is applied to the kinetic study of highly reactive enol ethers EtO-C(R) = CH-R′, where R and R′ are H or Me. A value of 2.2 × 108M?1s?1 is obtained for k, the rate constant for free bromine addition to EtO-CH = CH2, by extrapolating the kinetic bromide ion effects to [Br?] = 0. An α-methyl effect (kα-Me/kH)EtO of 15 is found; this is a small decrease in the methyl effect compared to the marked increase in the double bond reactivity. For the enol acetate MeCOO-CH = CH2, whose rate constant is 6 × 102M?1s?1, (kα-Me/kH)OCOMe is 21. The dependence of substituent effects on reactivity is discussed in terms of the Hammond effect on the transition state position and of charge delocalization by group G of olefins G-CH = CH2.  相似文献   

3.
The dilute solution properties of linear, 18-arm, and 270-arm star polybutadienes have been studied in a theta solvent and in a good solvent. Values of the radius of gyration RG, the second virial coefficient A2, the intrinsic viscosity [η], and the diffusion coefficient D0 have been measured for each polymer. The ratios RT/RG, RV/RG, and RH/RG for each type of polymer are used to compare the four dilute solution properties. RT is termed the “thermodynamic radius.” It is the radius of the hard sphere with the same excluded volume as the polymer coil. RT is calculated from A2 by RT = (3A2M2/16ηNA)1/3. RV and RH are equivalent hard spheres defined for the intrinsic viscosity and translational diffusion coefficient, respectively. RT/RG, RV/RG, and RH/RG increase from about 0.7 for linear polymer coils as the number of arms in the star increases. Values of the ratios for the 18-arm stars are less than the value for the hard-sphere, but the values of the ratios of the 270-arm stars are equal to the hard-sphere limit within experimental error.  相似文献   

4.
Three types of metal complexes containing coordinated zwitterionic 8-Quinolinol(oxine) are isolated from the reaction ofMOx 2 (M=divalent Ni, Mn, or Mg; HO x =oxine) and haloacetic acidsRCO2H (R=CF3, CCl3, CHCl2, or CH2Cl) in benzene. These types are:M(O2CR)Ox·HOx forM=Ni,R=CCl3, CHCl2, and CH2Cl and forM=Mn,R=CHCl2.MOx(HOx) (RCO2)MOx·nH2O forM=Ni, Mn, or Mg,R=CF3 andn=1,1, and 4, respectively.MO x (HOx) (RCO2)2 MOx forM=Mn andR=CCl3. These types are compared with the simple mixed chelateMn(O2CCH2Cl)Ox. Interrelated reactions are suggested to explain the formation of these metal complexes and the contributing factors are discussed. The coordination of the zwitterion to the metal ion through its phenolate oxygen and the presence of the triatomic system+N–H...O in the three types of metal complexes are evidenced by typical infrared bands. Analytical and spectral data are in accordance with the suggested formulations.
Koordination von zwitterionischem 8-Chinolinol (Oxin) an gemischten Oxinat-Carboxylat-Komplexen des divalenten Nickel, Mangan und Magnesium
Zusammenfassung Drei Typen von Metallkomplexen mit koordiniertem zwitterionischem 8-Chinolinol (Oxin) wurden aus der Reaktion vonMOx 2 [M=Ni(II), Mn(II), Mg(II); HOx=Oxin] mit Halogen-essigsäurenRCOOH (R=CF3, CCl3, CHCl2, CH2Cl) in Benzol isoliert. Es werden Reaktionswege zur Bildung der Komplexe diskutiert. Die Koordination des Zwitterions über den phenolischen Sauerstoff und die Präsenz der Gruppierung+N–H...O in allen Typen der untersuchten Metallkomplexe wird auf Grund typischer IR-Banden nachgewiesen.
  相似文献   

5.
Ni(0)‐complex promoted dehalogenation polymerization of 1,2‐bis(4‐bromophenyl)ethylene derivatives gave poly(p‐biphenylene vinylene) type polymers, [—C6H2R—CR2 = CR2—C6H2R—)n (P(R1,H) and P(H,R2) ], having substituents (R1 = Me, Et, CHMe2, and n‐C8H17, R2 = Me, Et, n‐C6H13, n‐C11H23, and Ph) at the benzene ring or vinylene group in 90–99% yields. The polymers were soluble in organic solvents such as CHCl3, dimethylformamide, and tetrahydrofuran, and gave Mn of 2.4–5.3 × 103 in gel permeation chromatography analysis. The absorption peak of the polymers appeared at a longer wavelength than that of the corresponding monomers by about 30 nm due to the expansion of the π‐conjugation system. The polymers were photoluminescent in solutions and in their films, emitting blue or green light. P(R1,H)s gave higher quantum yields (Φ = 0.35–0.51) than P(H,R2) s in CHCl3. P(H,R2) s showed a large Stokes shift (9600–13,500 cm−1) in their photoluminescence. Single‐layer and multilayer light emitting diodes using vacuum deposited thin film of P(H,Ph) were prepared. Polymers with long alkyl substituents formed an ordered structure in the solid state as judged from their XRD patterns. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1493–1504, 2000  相似文献   

6.
The syntheses and characterization of four new linear pentadentate ligands and their CoIII complexes are described: N,N′-[(pyridine-2,6-diy)bis(methylene)]bis[sarcosine] (sarmp), N,N′-[(pyridine-2,6-diyl)bis(methylene)]bis[(R)- or (S)-proline] ((R,R)- or (S,S)-promp), N,N′-[(pyridine-2,6-diyl)bis(methylene)]bis[N-(methyl)-(R)- or (S)-alanine] ((R,R)- or (S,S)-malmp); 2,2′-[pyridine-2,6-diyl]bis[(S)- or rac-N-(acetic acid)pyrrolidine] ((S,S)- or rac-bapap). The complexes were characterized and, with but one exception, complex formation is stereospecific: Δ-exo-(R,R) (or Λ-exo-(S,S)) for promp and Λ-(R,R) (or Δ-(S,S)) for bapap. The exception is [Co((R,R)- or (S,S)-malmp)H2O]ClO4 for which two forms are obtained, to which Λ-endo-(R,R) (or Δ-endo-(S,S)) and, tentatively, Δ-unsymmetric-(R,R)- (or Λ-unsymmetric-(S,S)-) configurations are assigned. X-Ray crystal structures are presented for the complexes [Co(sarmp)H2O]ClO4, [Co((S,S)-promp)H2O]ClO4, [Co(rac-bapap)H2O]ClO4 and endo-[Co(rac-malmp)H2O]ClO4. Ligand acid dissociation and CoII and FeII complex-formation constants are reported.  相似文献   

7.
In our ongoing development of ferrocene ligands, 1‐dimethylamino‐2‐(diphenylphosphinothioyl)ferrocene is being used as a convenient building block to obtain racemic or enantiomerically pure ligands. Using this building block in large excess allowed the formation of several by‐products, two of which have already been reported; the structure of a third by‐product, namely 1‐(diphenylphosphinothioyl)‐2‐{[(diphenylphosphinothioyl)sulfanyl]methyl}ferrocene, [Fe(C5H5)(C30H25P2S3)], is presented here. The crystal structure is built up from a ferrocene unit, with one of the cyclopentadienyl (Cp) rings substituted in the 1‐ and 2‐positions by a protected diphenylphosphinothioyl group and a [(diphenylphosphinothioyl)sulfanyl]methyl fragment, –CH2SP(=S)Ph2. There are C—H...S interactions which result in the formation of chains parallel to the c axis. After desulfurization, the crude material was then reacted with Pd and Pt (M) precursors [MCl2(CH3CN)2] to yield two isostructural dinuclear complexes arranged around twofold axes, namely (R,R/S,S)‐bis{μ‐[2‐(diphenylphosphanyl)ferrocen‐1‐yl]methanethiolato‐κ3P,S:S}bis[chloridopalladium(II)] pentane disolvate, [Pd2{Fe(C5H5)(C18H15PS)}2Cl2]·2C5H12, and the platinum(II) analogue, (R,R/S,S)‐bis{μ‐[2‐(diphenylphosphanyl)ferrocen‐1‐yl]methanethiolato‐κ3P,S:S}bis[chloridoplatinum(II)] toluene monosolvate, [Pt2{Fe(C5H5)(C18H15PS)}2Cl2]·C7H8, in which the two metal atoms present a slightly distorted square‐planar geometry formed by two bridging S atoms and P and Cl atoms. The P,S‐chelating ligand results from the rupture of one of the P—S bonds in the starting ligand. These dinuclear complexes display a butterfly geometry. Surprisingly, only the (R,R/S,S) diastereoisomer has been isolated.  相似文献   

8.
The photooxygenation of (4R,4aS,7R)-4,4a,5,6,7,8-hexahydro-4,7-dimethyl-3H-2-benzopyran ( 16 ) was performed in (i) MeOH, (ii) acetaldehyde, and (iii) acetone at ?78°. The products obtained respectively were (i) (2R)-2-[(1S,4R)-4-methyl-2-oxocyclohexyl]propyl formate ( 17 ; 72% yield), (ii) 17 (54.5%), (1R,4R,4aS,7R)-3,4,4a,5,6,7-hexahydro-4,7-dimethyl-1H-2-benzopyran-2-yl hydroperoxide ( 19 ; 16.7%), a 12:1 ratio of (3R,4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,7,10-trimethyl-6H-[2]benzopyrano[1,8a-e]-1,2,4-trioxane ( 20 ) and its C(3)-epimer 21 (17%), together with evidence for the 1,2-dioxetane ( 22 ) originating from the addition of dioxygen to the re-re face of the double bond of 16 , and iii) unidentified products and traces of 22 . Addition of trimethylsilyl trifluoromethanesulfonate (Me3SiOTf) to the acetone solution of 16 after photooxygenation afforded (4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,3,7,10-tetramethyl-6H-[2]benzopyrano[1,8a-e]-1,2,4,-trioxane ( 23 , 40%). The photooxygenation of 16 in CH2Cl2 at ?78° followed by addition of acetone and Me3SiOTf afforded 17 (11%), 23 (59%), and (4aR,7R,7aS,10R,11aR)-7,7a,8,9,10,11-hexahydro-3,3,7,10-tetramethyl-6H-[2]benzopyrano[8a,1-e]-1,2,4-trioxane ( 24 ; 5%. Repetition of the last experiment, but replacing acetone by cyclopentanone, gave 17 (16%), (4′aR,7′R,7′aS,10′R,11′aR)-7′,7′a,8′,9′,10′,11′-hexahydro-7′,10′-dimethylspiro[cyclopentane-1,3′-6′H-[2]benzopyrano[1,8a-e]-1,2,4-trixane] ( 25 ; 61%), and (4′aR,7′R,7′aS,10′R,11′aR)-7′,7′a,8′,9′,10′,11′-hexahydro-7′,10′-dimethylspiro[cyclopentane-1,3′-6′H-[2]benzopyrano[8a,1-e]-1,2,4-trixane] ( 26 , 4%). The X-ray analysis of 23 was performed, which together with the NMR data, established the structure of the trioxanes 20, 21, 24, 25 , and 26 . Mechanistic and synthesis aspects of these reactions were discussed in relation to the construction of the 1,2,4-trioxane ring in arteannuin and similar molecules.  相似文献   

9.
The kninetics of acid-catalyzed acetalization and ketalization of poly(vinyl alcohol) (PVA) were systematically studied in completely homogeneous media with carefully selected solvents. Thus the acetalization reaction was run in water with six aldehydes [R1CHO (R1 = H, CH3, C2H5, n-C3H7, i-C3H7, ClCH2)], whereas the ketalization in dimethylslfoxide with 11 ketones [R2CH3CO (R2 = CH3, C2H5, n-C3H7, i-C3H7, n-C4H9, i-C4H9, tert-C4H9, C6H5CH2, C6H5CH2CH2), cyclopentanone, and cyclohexanone]. The latter was difficult to proceed in aqueous media. Both reactions were reversible and bimolecular and, despite the use of different solvents, gave similar heats of reaction (7.5 kcal/mol) and activation energies (ca. 15 kcal/mol) except for the case of formaldehyde and chloroacetaldehyde; however the equilibrium constants at 25°C showed that the acetalization is thermodynamically much more favored than the ketalization (ca. 5000 vs. 0.01–0.9), probably because of steric hindrance of the ketone substrate. The rate constants of hydrolysis (reverse reactions) for the poly(vinyl acetal) and poly(vinyl ketal) followed the Hammett-Taft equation to give a single p* (=3.60) that is very close to that for the hydrolysis of diethyl acetal and ketal. From these and other data, it was concluded that the polymer hydrolysis, as well as PVA acetalization and ketalization, are all electrophilic reaction where the formation of hemiacetal or hemiketal is the rate-determining step. © 1996 John Wiley & Sons, Inc.  相似文献   

10.
Molecules of phthal­imide [1H‐iso­indole‐1,3(2H)‐dione], C8H5NO2, are linked by N—H?O hydrogen bonds [H?O 2.02 Å, N?O 2.8781 (16) Å and N—H?O 167°] and by C—H?O hydrogen bonds [H?O 2.54 and 2.56 Å, C?O 3.3874 (18) and 3.4628 (19) Å, and C—H?O 149 and 159°] into molecular ribbons, which are pierced by three different ring motifs; there are two centrosymmetric R(8) rings, each containing a single hydrogen bond, N—H?O in one case and C—H?O in the other, and R(9) rings containing all three hydrogen bonds.  相似文献   

11.
Single crystals of (1,3‐diamino‐5‐azaniumyl‐1,3,5‐trideoxy‐cis‐inositol‐κ3O2,O4,O6)(1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol‐κ3O2,O4,O6)lithium(I) diiodide dihydrate, [Li(C6H16N3O3)(C6H15N3O3)]I2·2H2O or [Li(Htaci)(taci)]I2·2H2O (taci is 1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol), (I), bis(1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol‐κ3O2,O4,O6)sodium(I) iodide, [Na(C6H15N3O3)2]I or [Na(taci)2]I, (II), and bis(1,3,5‐triamino‐1,3,5‐trideoxy‐cis‐inositol‐κ3O2,O4,O6)potassium(I) iodide, [K(C6H15N3O3)2]I or [K(taci)2]I, (III), were grown by diffusion of MeOH into aqueous solutions of the complexes. The structures of the Na and K complexes are isotypic. In all three complexes, the taci ligands adopt a chair conformation with axial hydroxy groups, and the metal cations exhibit exclusive O‐atom coordination. The six O atoms of the resulting MO6 unit define a centrosymmetric trigonal antiprism with approximate D3d symmetry. The interligand O...O distances increase significantly in the order Li < Na < K. The structure of (I) exhibits a complex three‐dimensional network of R—NH2—H...NH2R, R—O—H...NH2R and R—O—H...O(H)—H...NH2R hydrogen bonds. The structures of the Na and K complexes consist of a stack of layers, in which each taci ligand is bonded to three neighbours via pairwise O—H...NH2 interactions between vicinal HO—CH—CH—NH2 groups.  相似文献   

12.
Summary 8a-Methoxy-3,4-dihydro-2H-1,4-benzoxazin-6(8aH)-ones2 undergo regio- and stereospecific 1,3-dipolar cycloaddition reactions with diazomethane or diazoethane to yield 3,4,6a,9,9a,9b-hexahydro-pyrazolo[3,4-h][1,4]benzoxazin-6(2H)-ones3, which slowly isomerize in solution to give the 3,4,8,9,9a,9b-hexahydro-pyrazolo[3,4-h][1,4]benzoxazin-6(2H)-ones5. The carbon of the diazoalkane dipole is attached to carbon C-8 of the benzoxazinone. The structures of the obtained products were determined by1H- and13C-NMR spectroscopy. An X-ray crystal structure analysis of3 a was carried out at room temperature:C11H15N3O3,M r =237.26, orthorhombic, Pc21n,a=9.173 (5),b=9.133 (4),c=13.281 (6),V=1112.6 (9) Å3,Z=4,d x =1.416 g/cm–3, =0.93 cm–1,R=4.33%,R w =3.95% (919 observations, 168 parameters).
Herrn Prof. Dr. W. Fleischhacker zum 60. Geburtstag gewidmet  相似文献   

13.
The title compound, C3H12B10O2 or 1-COOH-1,2-closo-C2B10H11, forms centrosymmetric dimers through intermolecular hydrogen bonding between the carboxylic acid groups, resulting in the formation of an eight-membered ring [R(8)]. The C=O bond of the carboxylic acid group almost eclipses the unsubstituted cage C atom, with a C—C—C—O torsion angle of 2.6 (2)°.  相似文献   

14.
The molecules of 3‐amino‐4‐anilino‐1H‐isochromen‐1‐one, C15H12N2O2, (I), and 3‐amino‐4‐[methyl(phenyl)amino]‐1H‐isochromen‐1‐one, C16H14N2O2, (II), adopt very similar conformations, with the substituted amino group PhNR, where R = H in (I) and R = Me in (II), almost orthogonal to the adjacent heterocyclic ring. The molecules of (I) are linked into cyclic centrosymmetric dimers by pairs of N—H...O hydrogen bonds, while those of (II) are linked into complex sheets by a combination of one three‐centre N—H...(O)2 hydrogen bond, one two‐centre C—H...O hydrogen bond and two C—H...π(arene) hydrogen bonds.  相似文献   

15.
In addition to the known C11H16 hydrocarbons multifidene ( 4 ), aucantene ( 2 ), and ectocarpene ( 5 ), the marine brown alga Cutleria multifida produces trace amounts of the C9H12 hydrocarbon 7-melhylcycloocta-1,3,5-triene ( 8 ) and its valence tautomer 7-methylbicyclo[4.2.0]octa-2,4-diene, A second novel C9H12 hydrocarbon is 6-vinyicyclo-hepta-1,4-diene ( 9 ), a lower homologue of ectocarpene ( 5 ). Among the C11H16 hydrocarbons, 7-((1E/Z)-prop-l-enyl)cycloocta-1,4-diene ( 10 / 11 ) is found for the first time. The structure of all new products is confirmed by synthesis and spectroscopic data. The biosynthesis of the new hydrocarbons 8 – 11 is obviously linked to the pathways which lead to the major products giffordene ( 7 ), (6S)-ectocarpene ((6S)- 5 ), and (4R,5R)-aucantene ((4R,5R)- 2 ). Consecutive reactions of certain thermolabile primary products proceed via electrocyclic ring closure, 3,3-sigmatropic rearrangement, or a 1,7-sigmatropic H-shift.  相似文献   

16.
The absolute configuration of the title cis‐(1R,3R,4S)‐pyrrolidine–borane complex, C18H34BNO2Si, was confirmed. Together with the related trans isomers (3S,4S) and (3R,4R), it was obtained unexpectedly from the BH3·SMe2 reduction of the corresponding chiral (3R,4R)‐lactam precursor. The phenyl ring is disordered over two conformations in the ratio 0.65:0.35. The crystallographic packing is dominated by the rarely found donor–acceptor hydroxy–borane O—H...H—B hydrogen bonds.  相似文献   

17.
The geometry of racemic methyl 2‐(4‐methyl‐2‐thio­xo‐2,3‐di­hydro­thia­zol‐3‐yl­oxy)­propanoate, C8H11NO3S2, (I), is characterized by a distorted heterocyclic five‐membered ring and an enantiomorphic N‐alkoxy substituent, which is inclined at an angle of −68.8° to the thia­zole­thione plane in (M)‐(I). The unit cell consists of a 1:1 ratio of R,P‐ and S,M‐configured mol­ecules of (I). The combination of a P configuration at the N—O axis and an R configuration at the asymmetric propanoate Cβ atom on one side, and an S,M configuration on the other side, is considered to originate from steric interactions. The largest substituent at the asymmetric propanoate Cβ atom, i.e. the methoxycarbonyl group, resides above the methyl substituent; the medium‐sized propanoate γ‐methyl substituent points in the opposite direction with respect to the N—O bond, whereas the H atom is located above the C=S double bond of the thiazolethione subunit.  相似文献   

18.
The viscosities, rubbery deformations, densities, and their dependence on temperature have been measured for several series of polybutadienes with molecular weights ranging from 5,000 to 400,000 and differing in proportions of cis and trans structures (cis content from 40 to 95%). On the basis of the viscosity measurements the critical molecular weight Mc has been determined, corresponding to a sharp change in the nature of the viscosity versus molecular weight dependence. Rubbery deformations are displayed pronouncedly in specimens with M > Mc and are closely related to the appearance of non-Newtonian flow. The value of Mc depends on the relative content of cis and trans forms. When M > Mc, the initial viscosity is a parameter sensitive to the microstructure of polybutadienes, so that with at a single molecular weight, depending on the ratio of cis and trans units, the viscosity may vary over a more than tenfold range. The glass transition temperature and activation energy of viscous flow rise regularly with increasing trans content in the polymer chain, these parameters becoming independent of the molecular weight for specimens with M > Mc within a series of polybutadienes of equal microtacticity. Thermomechanical investigations of polybutadienes also made it possible to define more accurately the boundaries of the crystallization region and the dependence of the melting point on the microtacticity. The results obtained are discussed on the basis of modern ideas of polymer structure.  相似文献   

19.
20.
The crystal structure of the title compound, C5H7N2+·C12H10NO4S2, consists of two independent cation–anion pairs, A and B. Within each pair, the H—N—C—N*—H grouping (N*—H is the pyridinium function) and one N—S—O moiety of the anion are linked by N*—H⃛N and N—H⃛O hydrogen bonds to form an antidromic ring motif of type R22(8). The remaining amino donors give rise to N—H⃛O hydrogen bonds, connecting the ion pairs into ABAB– chains. The structure testifies to the persistence of the R22(8) motif in question, which was previously detected as a highly robust supramolecular synthon in a series of onium di(methane­sulfonyl)­amidates. The structure is pseudosymmetric; the anion positions correspond to space group P21/n, but those of the cations do not.  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