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 46 毫秒
1.
Mass spectrometric analysis of quaternary nitrogen compounds has shown that three principal thermal processes occur, namely dealkylation, Hofmann degradation and substitution. Only in the cases of dealkylation and the Hofmann degradation has direct chemical evidence been previously obtained. The first part of this paper presents mass spectrometrical and chemical evidence for the pyrolytic formation of a substitution product, e. g. 5 . By means of different deuterated derivatives the mechanisms of fragmentation of the mavacurine derivative ϵ2-dihydromavacurine methochloride ( 8 ) and of the corresponding Hofmann base ( 14 ) have been formulated.  相似文献   

2.
J. Imai  Y. Kondo  T. Takemoto 《Tetrahedron》1976,32(16):1973-1977
Treatment of l(14S)-β-canadine methochloride (1b) and d(14R)-β-canadine methochloride (1c) with organometals gave d- (2b) and l-2,3-methylenedioxy-9,10-dimethoxyochotensanes (2c), respectively. The structures of these derivatives were proved by chemical and spectral means. The CD spectra of 2b showed Davydov split extrema centered at 284 nm with a positive first Cotton effect, while 2c showed the antipodal curve of 2b. Consequently, the absolute configurations of 2b and 2c were concluded the 14R and 14S, respectively.Application of the anionic rearrangement to N-methyltharictricavine chloride (15) led to 2,3 - methylenedioxy - 9,10 - dimethoxy - 13 - methylochotensane (18) together with the Hofmann methines 16 and 17. The stereochemistry of 18 was confirmed in terms of the nuclear Overhauser effects.  相似文献   

3.
The structure of C-alkaloid-O ( 1 =his HCL-salt) was established on the basis of spectral data and comparison with the analogous compound 2 , synthesized from a natural product of known absolute configuration. Internally consistent 13C-NMR. assignments are given for 1, 2 , huntrabrine methochloride (4) and its diacetate derivative 5 .  相似文献   

4.
Single crystal XRD is used to study the crystals of salts of indole alkaloid norfluorocurarine: hydrochloride recrystallized from absolute alcohol, dihydrate hydrochloride recrystallized from water, methochloride monohydrate recrystallized from water, solvate form of methochloride obtained from ethanol, and methobromide monohydrate. Intra- and intermolecular hydrogen bonds are analyzed in these crystals. The crystal structures of norfluorocurarine methochloride and methobromide monohydrates are isomorphic. In norfluorocurarine salts, the orientation of the carbonyl group is determined by the intramolecular C19=O…H-N1 hydrogen bond that is absent in the solvate form with ethanol.  相似文献   

5.
 The optical absorption, photoluminescence, and photoconductivity spectra of some compounds of the formulas [R(CH2) n NH3] x M y X z , [R(CH2) n NH(CH3)2] x M y X z , [R(CH2) n S(CH3)2] x M y X z , [R(CH2) n SC(NH2)2] x M y X z , and [R(CH2) n SeC(NH2)2] x M y X z (R = organic residue; M = Bi(III), Pb(II), Sn(II), Cu(I), Ag(I) etc; X = I, Br, Cl; n, x, y, z = 0, 1, 2, 3, …) are briefly reviewed, and some new results are reported. The position, intensity, and shape of the excitonic bands depend on the dimensionality and size of the inorganic network as well as on the nature of the M, X, R, and onium moieties.  相似文献   

6.
For the 53 neutral atoms from He to Xe in their ground states, the average distances < u> n l , n l in position space and < v> n l , n l in momentum space between an electron in a subshell nl and another electron in a subshell n l are studied, where n and l are the principal and azimuthal quantum numbers of an atomic subshell, respectively. Analysis of 1700 subshell pairs shows that the electron-pair distances < u> n l , n l in position space have an empirical but very accurate linear correlation with a one-electron quantity U n l , n l L r +S r 2/(3L r ), where L r and S r are the larger and smaller of subshell radii < r> n l and < r> n l , respectively. The correlation coefficients are never smaller than 0.999 for the 66 different combinations of two subshells appearing in the 53 atoms. The same is also true in momentum space, and the electron-pair momentum distances < > n l , n l have an accurate linear correlation with a one-electron momentum quantity V n l , n l L p +S p 2/(3L p ), where L p and S p are the larger and smaller of average subshell momenta < p> n l and < p> n l , respectively. Trends in the proportionality constants between < u> n l , n l and U n l , n l and between < > n l , n l and V n l , n l are discussed based on a hydrogenic model for the subshell radial functions. Received: 8 April 1998 / Accepted: 6 July 1998 / Published online: 18 September 1998  相似文献   

7.
The glass transition temperature Tg of nylon 6 decreases monotonically toward a finite value Tgl upon increase of the moisture content. The mechanism of this decrease entails the reversible replacement of intercaternary hydrogen bonds in the accessible regions of the polyamide. The limiting glass transition temperature Tgl is approached when the moisture content approaches Wl, which corresponds to the amount of water required for complete interaction with all accessible amide groups. Denoting with Tg0 the glass transition temperature of the dry polymer, the effect of water on Tg is represented by the equation, Tg = (ΔTg)0 exp{?[ln(ΔTg)0]W/τWl} + Tgl, where (ΔTg)0 = Tg0 ?Tgl, and τ = W(Tgl+1)/Wl. This equation appears to be generally applicable to hydrophilic polymers, since correspondingly calculated data are also in very good agreement with experimental data for polymers such as nylon 66, poly(vinyl alcohol), and polyN-vinylpyrrolidone. The effect of water of Young's modulus E of nylon 6 is represented by an analogous relationship, and the quantity In[(E?El)/(Tg?Tgl)] is a linear function of the moisture content.  相似文献   

8.
Summary Reactions oftrans-[M(N2)2(dppe)2] (A;M=Mo, W;dppe=Ph 2PCH2CH2PPh 2) with ethyldiazoacetate, N2CHCOOEt, yield the bisdiazoalkane speciestrans-[M(N2CHCOOEt)2(dppe)2], upon simple replacement of the dinitrogen ligand by ethyldiazoacetate. However, diazomethane, N2CH2, reacts withA with loss of N2 to give products which we tentatively formulate as containing methylene ligands,trans-[M(CH2)2(dppe)2].
Herstellung von Bisdiazoalkan- und ähnlichen Komplexen aus den Reaktionen von Diazoverbindungen mit Distickstoffkomplexen des Typstrans-[M(N2)2(Ph 2PCH2CH2PPh 2)2] mitM=Mo oder W
Zusammenfassung Die Reaktion vontrans-[M(N2)2(dppe)2] (A:dppe=Ph 2PCH2CH2PPh 2 undM=Mo oder W) mit Ethyldiazoacetat, N2CHCOOEt, ergab nach einfachem Austausch des Distickstoffliganden mit Ethyldiazoacetat die Bisdiazoalkanetrans-[M(N2CHCOOEt)2(dppe)2]. Diazomethan (N2CH2) hingegen reagierte mitA unter Verlust von N2 zu Produkten, die tentativ alstrans-[M(CH2)2(dppe)2] mit Methylenliganden formuliert wurden.
  相似文献   

9.
Reactions of (m- and p-ClC 6 H 4 NH 2 ), (p-BrC 6 H 4 NO 2 ), and (p-ClCOC 6 H 4 NO 2 ) with sodium O,O′-ditolyl/dibenzylphosphorodithionates, (ArO) 2 PS 2 Na, (Ar = o?, m?, and p?CH 3 C 6 H 4 or –C 6 H 5 CH 2 ) in 1:1 molar ratio in refluxing toluene under anhydrous conditions resulted in the formation of the compounds (ArO) 2 PS 2 C 6 H 4 L and (ArO) 2 PS 2 COC 6 H 4 L (L = NH 2 or NO 2 ) in 87–94% yield. These viscous compounds were characterized by elemental analyses, molecular weight determination, and IR and NMR ( 1 H, 13 C, and 31 P) spectroscopic studies, which revealed a monodentate mode of bonding of the dithiophosphate moiety with the carbon of the phenyl ring of the organic moiety leading to a P–S–C linkage.  相似文献   

10.
Disproportionation reactions between (CF 3 CH 2 O) 3 GeNHC 6 H 5 ? n F n and TiCl 4 in petroleum ether (40?60°C) at 0° to ?10°C give (CF 3 CH 2 O) 2 Ge(NHC 6 H 5 ? n F n ) 2 .2TiCl 4 and (CF 3 CH 2 O)Ge(NHC 6 H 5 ? n F n ) 3 . 2TiCl 4 adducts. However, complete disproportionation of (CF 3 CH 2 O) 3 Ge(NHC 6 H 5 ? n F n ) (n = 1,2) occurs at ?55 to ?60°C to give Ge(NHC 6 H 5 ? n F n ) 4 .3TiCl 4 . These complexes give double adducts on reactions with CH 3 NO 2 and CH 3 CN. All the products are characterized by elemental analyses and IR, 1 H, and 19 F NMR spectroscopy. A comparative disproportionation of the germanamines and analogous silanamines is discussed.  相似文献   

11.
The reactions of n-butyl stannonic acid with(PhO) 2 P(O)H leads to the formation of a hexameric tin cage [{(n-BuSn) 3 (PhO) 3 O} 2 {HPO 3 } 4 ].This reaction involves an in situ P─O bond cleavage and the generation of a [HPO 3 ] 2? ion. A direct reaction of six equivalents of n-BuSnO(OH) acid with six equivalents of C 6 H 5 OH and four equivalents of H 3 PO 3 also leads to the formation of same cage structure. A tetranuclear organooxotin cage[(PhCH 2 ) 2 Sn 2 O(O 2 P(OH)-t-Bu) 4 ] 2 has been assembled by debenzylation involving the reaction of (PhCH 2 ) 2 SnCl 2 ,(PhCH 2 ) 2 SnO·H 2 O or (PhCH 2 ) 3 SnCl with two equivalents of t-BuP(O)OH 2 . A half-cage intermediate [(PhCH 2 ) 2 Sn 2 O(O 2 P(OH)-t-Bu) 4 ] has been detected. New organotin cations of the type [n-Bu 2 Sn(H 2 O) 4 ] 2+[2,5-Me 2 -C 6 H 3 SO 3 ]? 2 and {[n-Bu 2 Sn(H 2 O) 3 LSn(H 2 O) 3 (n-Bu) 2 ] 2+[1,5-(SO 3 ) 2 -C 10 H 6 ] 2?} have been obtained in the reactions of n-Bu 2 SnO or (n-Bu 3 Sn) 3 O with 2,5-dimethyl sulfonic acid and 1,5-naphthalene disulfonic acid respectively. These organotin cations form interesting supramolecular structures in the solid state as a result of O─H─···O hydrogen bonding.  相似文献   

12.
The formation of cluster orbitals in CsSn2Br5 is discussed and related more generally to tetragonal compounds of the type AB2X5 (A=monovalent cation; B=Sn, Pb; X=Cl, Br, I). The crystal structures of CsSn2Cl5 and CsSn2Br5 have been solved by single-crystal X-ray diffraction. These compounds are isostructural with each other and a range of AB2X5 structural analogues. In many AB2X5 compounds where B is a subvalent main group metal a tetragonal cell is observed with space group I4/mcm. The structures of CsSn2Br5 and CsSn2Cl5 are layered with polymeric sheets of [Sn2X5]nn separated by the Cs+ cations. Stereochemical considerations suggest that stabilization of this structural form, rather than the more ionic NH4Pb2Cl5 or NaSn2Cl5 structures, is through interaction of the “nonbonding” valence electron pairs on tin with low-lying empty d-orbitals on neighboring X atoms. Electronic structure calculations based on the structural data confirm the likelihood of cluster orbital formation. Crystal data: CsSn2Cl5, tetragonal, I4/mcm, a=8.153(1) Å, c=14.882(4) Å, Z=4, R1=0.0215, wR2=0.0503 [I>2σ(I)], R1=0.0393, wR2=0.0536 (all data); CsSn2Br5, tetragonal, I4/mcm, a=8.483(6) Å, c=15.28(2) Å, Z=4, R1=0.0607, wR2=0.1411 [(I>2σ(I)], R1=0.1579, wR2=0.1677 (all data).  相似文献   

13.
The reaction of [PtCl2(COD)] (COD=1,5-cyclooctadiene) with diisopropyl-2-(3-methyl)indolylphosphine (iPr2P(C9H8N)) led to the formation of the platinum(ii ) chlorido complexes, cis-[PtCl2{iPr2P(C9H8N)}2] ( 1 ) and trans-[PtCl2{iPr2P(C9H8N)}2] ( 2 ). The cis-complex 1 reacted with NEt3 yielding the complex cis-[PtCl{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}] ( 3 ) bearing a cyclometalated κ2-(P,N)-phosphine ligand, while the isomer 2 with a trans-configuration did not show any reactivity towards NEt3. Treatment of 1 or 3 with (CH3)4NF (TMAF) resulted in the formation of the twofold cyclometalated complex cis-[Pt{κ2-(P,N)-iPr2P(C9H7N)}2] ( 4 ). The molecular structures of the complexes 1–4 were determined by single-crystal X-ray diffraction. The fluorido complex cis-[PtF{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}] ⋅ (HF)4 ( 5 ⋅ (HF)4) was formed when complex 4 was treated with different hydrogen fluoride sources. The Pt(ii ) fluorido complex 5 ⋅ (HF)4 exhibits intramolecular hydrogen bonding in its outer coordination sphere between the fluorido ligand and the NH group of the 3-methylindolyl moiety. In contrast to its chlorido analogue 3 , complex 5 ⋅ (HF)4 reacted with CO or the ynamide 1-(2-phenylethynyl)-2-pyrrolidinone to yield the complexes trans-[Pt(CO){κ2-(P,C)-iPr2P(C9H7NCO)}{iPr2P(C9H8N)}][F(HF)4] ( 7 ) and a complex, which we suggest to be cis-[Pt{C=C(Ph)OCN(C3H6)}{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}][F(HF)4] ( 9 ), respectively. The structure of 9 was assigned on the basis of DFT calculations as well as NMR and IR data. Hydrogen bonding of HF and NH to fluoride was proven to be crucial for the existence of 7 and 9 .  相似文献   

14.
A modification of a variation principle due to Delves, is derived which permits the direct calculation of energy differences between states of two different Hamiltonians: [Δ ??] = 〈X0| ??xWx|X1〉 – 〈Y0|??yWy|y1〉 + 〈X0| Δ ??|Y0〉 · 〈X0| Y0?1. Δ ?? = ??y – ??x, |X0〉 and |Y0〉 are the wave functions for the X and Y states and |X1〉 and |Y1〉 are functions defined in the text. The principle is applied to a few simple examples.  相似文献   

15.
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   

16.
Three alkali metal acetylides, namely KNaC2, KRbC2, and NaRbC2, were synthesized and characterized by means of X‐ray powder diffraction. KNaC2 and KRbC2 crystallize as a variant of the anti‐PbCl2‐type structure (Pnma, Z = 4), whereas NaRbC2 crystallizes as a variant of the anti‐PbFCl‐type structure (Pmmn, Z = 2). Based on a simple systematic approach developed by Sabrowsky et al. for inter‐alkali metal chalcogenides all known inter‐alkali metal acetylides can be classified into two classes: variants of the anti‐PbCl2 type structure and variants of the anti‐PbFCl type structure. Acetylides with Q(ABC2) ≤ 1.45 crystallize in the anti‐PbCl2‐type structure, whereas for Q(ABC2) > 1.45 the anti‐PbFCl‐type structure is found (Q(ABC2) = Vm(A2C2)/Vm(B2C2) with Vm(A2C2) > Vm(B2C2); Vm: molar volume, A, B = alkali metals).  相似文献   

17.
In the copolymerization of monomers M1 and M2 which form polymer radicals of chain length n of N1n with electron on a M1 type and N2n with one on a M2 type, it is assumed that the specific rates of termination between N1n and N1n and N1s, N1n and N2s, and N2n and N2s are kα(ns)?a, kβ(ns)?a, and kγ(ns)?a, respectively, where kα, kβ, and kγ are the rate constants of reaction between segment radicals in the respective termination, and a is constant. The relation between kinetic chain length n? and polymerization rate Rp is derived as: 1/n? = 1/n?0 + const. (Rp)A(a), where n?0 is the kinetic chain length of the polymer formed by transfer and A (a) is unity (predominance of transfer) and 1/(1–2a) (no transfer). In the copolymerization between methyl methacrylate (M1) and styrene (M2) at 60°C, when Rp → 0, kr12/k12 + kr21/k21 = 5.9× 10?5 is obtained, where kr12 and kr21 are the rate constants of transfer of N1 to M2 and N2 to M1, and k12 and k21 are the rate constants of propagation of N1 to M2 and N2 to M1. In the absence of transfer, the a value is found to be 0.065 ± 0.008, from the relation between n? and Rp, regardless of the monomer composition. Such a value is also estimated by setting b = 0.72 in a = 0.153 (2b–1), where b is the constant in the Mark-Houwink equation. Further, the value of kβ is found to be 1.18 × 109l./mole-sec, which is comparable with the diffusion-controlled rate of reaction between small molecules. The rate of reaction between segment radicals is fivefold larger than the polymer-polymer termination when transfer predominates.  相似文献   

18.
19.
Positron annihilation lifetime measurements are reported for four monodisperse polystyrenes with molar mass M = 4,000, 9,200, 25,000, and 400,000. The temperature dependences of orthopositronium (o-Ps) lifetime (τ3) and intensity (I3) were measured from 5°C to Tg + 30°C for each sample. From these data, the free volume hole size, 〈vf3)〉, and fractional free volume hps=CI3vf3)〉 were calculated. The temperature dependences of τ3, 〈vf3)〉 and hps show a discrete change in slope at an effective glass transition temperature, Tg,ps, which is measurably below the conventional bulk Tg. This suggests that τ3 is sensitive to large holes which retain their liquid-like mobility in the glassy state. Good agreement was found for T > hg,ps between hps and the theoretical free volume fraction hth deduced from experimental P-V-T data for polystyrene using the statistical mechanical theory of Simha and Somcynsky. Below Tg,ps, deviations between hps and hth are observed, hps falling increasingly below hth as temperature decreases. Whereas hps and hth depend strongly on M in the melt, each essentially independent of M in the glass. A free volume quantity, computed from the bulk volume, which is in good numerical agreement with the Simha-Somcynsky h-function in the melt, gives improved agreement with hps in the glassy state. © 1994 John Wiley & Sons, Inc.  相似文献   

20.
Interaction between two double layers for the Na3PO4 type electrolytes at y 0 > y d > 0 were investigated with the aid of λ parameter methods and the accurate numeral results were given. The maximum of the interaction energies between two double layers for Av+ Bv- A_{v_ + } B_{v_ - } type electrolytes at y 0 > y d > 0 is independent of y 0, but increases with y d . The interaction energies between two double layers for NaCl, CaCl2, FeCl3, Na2SO4 and Na3PO4 type electrolytes at y 0 = 5 and y d = 2 were compared. If the negative or positive ions for Av+ Bv- A_{v_ + } B_{v_ - } type electrolytes have the same charge number, the maximum of their interaction energies increases with the charge number of the ions of opposite charge. If the negative or positive ions of the electrolytes have different charge numbers, the relative heights of the maxima of their interaction energies are indefinite. For the in pairs conjugate type electrolytes like CaCl2 and Na2SO4 or FeCl3 and Na3PO4, the larger the charge number of the negative ions of the electrolytes is, the higher the maximum of their interaction energies is. The results for Na3PO4 type electrolytes at y 0 > y d > 0 can also be applied to FeCl3 type electrolytes at y 0 > y d > 0.  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