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 31 毫秒
1.
5β-androstan-3-ones carrying a 6α-OH group show in their mass spectra a key-ion indicating the loss of water and C-1 to C-4 as C4H5O? particle. 6β-OH isomers lose instead C-1 to C-4 in form of C4H7O?.In 6α-hydroxy-androstan-3-ones differentiation between the connection of the A/B-ring system is possible, because in 5α-isomers the loss of C-3 to C-7 occurs as a C5H6O2 particle, while the 5β-isomers lose the same C atoms as a C5H7O? unit.Compounds with a 6β-OH group in an A/B trans connected ring system show a tendency for thermal water elimination. After rearrangement of the double bond in 4,5 position the typical fragments for 3-keto-Δ4-steroids are obtained.Occasionally a strong influence of a 6-OH group on fragmentation reactions in the D-ring system is observed: The presence of a 6α-OH group in an androstan-3,17-dione enhances the loss of C-16 and C-17 in the form of acetaldehydenol. Also the connection of the A/B-ring system may have a considerable influence on this type of reaction: In 6,17β-dihydroxy-androstan-3-ones only by trans connection of the A/B-ring system, C-16 and C-17 are lost with high probability after water elimination.  相似文献   

2.
A new procedure for the solid-phase synthesis of 2,6- and 2,7-diamino-4(3H)-quinazolinones is described. The method involves coupling of 2,4,6- and 2,4,7-trichloroquinazoline to a solid support via benzyl alcohol type linkers, subsequent displacement of chlorine at C-2 then at the C-6 or C-7 positions by amines (Fig. 1) and the cleavage of the products from the resin. The palladium-catalyzed amination of C-6 and C-7 positions with a representative set of amines in the presence of 2-(di-t-butylphosphino)biphenyl (DTBPBP), P(t-Bu)3 and 2,2′-bis(diphenylphosphino)-1,1′-binaphthyl (BINAP) ligands has been investigated. This method should prove to be a useful tool for constructing combinatorial libraries containing the 4(3H)-quinazolinone moiety.  相似文献   

3.
The syntheses of [Au(CC-4-C6H4CC-4-C6H4NN-4-C6H4NO2)(PPh3)] (3), trans-[Ru(CC-4-C6H4-CC-4-C6H4NN-4-C6H4NO2)Cl(dppm)2] (4), [Ru(CC-4-C6H4CC-4-C6H4NN-4-C6H4NO2)(dppe)(η-C5Me5)] (5), and [Ni(CC-4-C6H4NN-4-C6H4NO2)(PPh3)(η-C5H5)] (6) are reported, together with a single-crystal X-ray diffraction study of 4. Quadratic nonlinearities for 36 and [Ru(CC-4-C6H4NO2)(dppe)(η-C5Me5)] (7) have been determined at 1.064 μm and 1.300 μm by the hyper-Rayleigh scattering (HRS) technique, comparison to related complexes revealing that β values increase on introduction of azo group and π-system lengthening.  相似文献   

4.
The complexes trans-[Ru(PMe3)4(CCPh)2] and trans-[Ru(PMe3)4(CCC6H4C6H4CCSnMe3)2] have been prepared from the reaction between trans-[Ru(PMe3)4Cl2] and an excess of either Me3SnCCPh or Me3SnCCRCCSnMe3 (R = p-C6H4C6H4), respectively. However, if only one equivalent of the latter reagent is used the rod-like polymeric species trans-[-Ru(PMe3)4CCRCC-]n can be isolated.  相似文献   

5.
The tautomerism of pyrazolones unsubstituted at position 3(5) has been investigated by 13C- and 1H NMR spectroscopic methods. Apart from chemical shift considerations and NOE effects the magnitude of the geminal 2J[pyrazole C-4,H3(5)] spin coupling constant permits the unambiguous differentiation between 1H-pyrazol-5-ol (OH) and 1,2-dihydro-3H-pyrazol-3-one (NH) forms. Whereas 1H-pyrazol-5-ols and 2,4-dihydro-3H-pyrazol-3-ones (CH-form) exhibit 2J values of approximately 9-11 Hz, in 1,2-dihydro-3H-pyrazol-3-ones this coupling constant is considerably reduced to 4-5 Hz. This can be mainly attributed to the removal of the lone-pair at pyrazole N−1 in the latter due to protonation or alkylation. According to the data obtained, 2-substituted 4-acyl-1,2-dihydro-3H-pyrazol-3-ones exist predominantly as pyrazol-5-ols in CDCl3 or benzene-d6 solution, whereas in DMSO-d6 also minor amounts of NH tautomer may contribute to the tautomeric composition. 2,4-Dihydro-2-phenyl-3H-pyrazol-3-one (1-phenyl-2-pyrazolin-5-one) exists in benzene-d6 solely in the CH-form, in CDCl3 as a mixture of CH and OH-form, whereas in DMSO-d6 a fast equilibrium between OH and NH isomer (with the former far predominating) is probable. For 11 compounds, including neutral and protonated molecules, we have calculated at the B3LYP/6-311++G** level, the 2J(1H,13C) coupling constants which are in good agreement with those measured experimentally.  相似文献   

6.
This communication describes an iron-catalyzed route for the synthesis of 6-substituted chromeno[3,2-c]quinolin-7-one. The method developed does not require any pre-functionalization to execute the pivotal coupling reaction at the C-3 position of flavones. The final step involves the consecutive application of imine formation, Csp2-Csp2 coupling and oxidation reaction, with aromatic aldehydes and 2-(2-aminophenyl)-4H-chromen-4-one as the reactants. Presence of electron donating/withdrawing groups was well tolerated in the aldehydes and the method developed could also be extended to other substituted 2-(2-aminophenyl)-4H-chromen-4-one’s. This is the first report of synthesis of 6-substituted chromeno[3,2-c]quinolin-7-one’s via direct functionalization of the C-3 site of flavones.  相似文献   

7.
The acid-mediated reaction of [{Co2(CO)6(μ-η2-HOCH2CC-)}2] (1) with the meta- and para-substituted aminothiophenols, 3-NH2-C6H4SH and 4-NH2-C6H4SH, affords the straight chain species, [{Co2(CO)6(μ-η2-(3-NH2-C6H4S)CH2CC-)}2] (2) and [{Co2(CO)6(μ-η2-(4-NH2-C6H4S)CH2CC-)}2] (3), respectively. The molecular structure of 3 reveals the presence of two isomeric forms differing in the relative disposition of the S-aryl groups. Conversely, reaction of 1 with the ortho-substituted aminothiophenol, 2-NH2-C6H4SH, furnishes the 10-membered macrocyclic species [{Co2(CO)6}2{cyclo-μ-η2:μ-η2-CH2C2C2CH2SC6H3-NH-2}] (4) along with the linear chain complex [{Co2(CO)6(μ-η2-(2-NH2-C6H4S)CH2CC-)}2] (5). On the other hand, treatment of 1 with the ortho-substituted mercaptopyridine, 2-SH-C5H4N, in the presence of HBF4 gives the salt [{Co2(CO)6(μ-η2-(2-S-C5H4NH)CH2CC-)}2](BF4)2 (6a) in good yield; work-up in the presence of base affords the neutral complex [{Co2(CO)6(μ-η2-(2-S-C5H4N)CH2CC-)}2] (6b). Single crystal X-ray diffraction studies have been reported on 3-5 and 6a.  相似文献   

8.
Zita Zalán 《Tetrahedron》2005,61(22):5287-5295
By condensation of 1-(2′-aminoethyl)-1,2,3,4-tetrahydroisoquinoline derivatives with substituted benzaldehydes, 1,6-unsubstituted and diastereomers of 1-methyl- or 6-methyl-substituted 4-aryl-9,10-dimethoxy-1,3,4,6,7,11b-hexahydro-2H-pyrimido[6,1-a]isoquinolines were prepared. The ring-chain tautomeric equilibria of most of these compounds in CDCl3 at 300 K were found to be shifted nearly totally towards either the cyclic or the open tautomeric forms, while the (6R*,11bR*)-6-methyl substituted compounds proved to be three-component tautomeric mixtures, the equilibria of which could be characterized by a Hammett-type equation. The conformational equilibria of the cyclic forms turned out to be strongly influenced by the 1- and 6-methyl substituents and the configurations of the substituted carbons (C-1 or C-6 and C-4) relative to C-11b.  相似文献   

9.
Treatment of the dihaloiron-chelating phosphine complex, Fe(DEPE)2Cl2 (1) with 2.5 equivalents of Me3SnCC6H5 (2) gives the monomeric complex, tras-Fe(DEPE)2(CCC6H5)2 (3, whereas the reaction of equimolar quantities of 1 and Me3SnCCRCCSnMe3 (R = p-C6H4 (4); p-C3)2C6H2 (5)) yields soluble and high molecular weight rigid rod polymeric species, trans-[-Fe(DEPE)2(−CCRCC-)]n (6,7).  相似文献   

10.
Isoxazolines 2 from the cycloaddition of imidazoline 3-oxides 1 with DMAD undergo rearrangement to 3,4-dihydro-2H-imidazol-1-ium-1-(1,2-bis-methoxycarbonyl-2-oxo-ethanides) 3, which spontaneously undergo elimination to give 3H-imidazol-1-ium-1-(1,2-bis-methoxycarbonyl-2-oxo-ethanides) 5 or 1H-imidazoles 6 when heated in toluene at reflux. The presence of the aromatic ring at C-6 decelerated the conversion and enhanced the yield of 5. Solvents more polar than toluene (e.g., DMSO) provided quantitative conversion of 2 into 6 in mild conditions, while in less polar solvents such as CCl4, the reaction rate was lowered and the yield of 5 enhanced. C-2 unsubstituted ylides 5 were treated with Ag2O or AgNO3 in the presence of Et3N at room temperature to give C-2 metallated derivatives 9 in excellent yields.  相似文献   

11.
Reactions of [Ru{C=C(H)-1,4-C6H4C≡CH}(PPh3)2Cp]BF4 ([ 1 a ]BF4) with hydrohalic acids, HX, results in the formation of [Ru{C≡C-1,4-C6H4-C(X)=CH2}(PPh3)2Cp] [X=Cl ( 2 a-Cl ), Br ( 2 a-Br )], arising from facile Markovnikov addition of halide anions to the putative quinoidal cumulene cation [Ru(=C=C=C6H4=C=CH2)(PPh3)2Cp]+. Similarly, [M{C=C(H)-1,4-C6H4-C≡CH}(LL)Cp ]BF4 [M(LL)Cp’=Ru(PPh3)2Cp ([ 1 a ]BF4); Ru(dppe)Cp* ([ 1 b ]BF4); Fe(dppe)Cp ([ 1 c ]BF4); Fe(dppe)Cp* ([ 1 d ]BF4)] react with H+/H2O to give the acyl-functionalised phenylacetylide complexes [M{C≡C-1,4-C6H4-C(=O)CH3}(LL)Cp’] ( 3 a – d ) after workup. The Markovnikov addition of the nucleophile to the remote alkyne in the cations [ 1 a–d ]+ is difficult to rationalise from the vinylidene form of the precursor and is much more satisfactorily explained from initial isomerisation to the quinoidal cumulene complexes [M(=C=C=C6H4=C=CH2)(LL)Cp’]+ prior to attack at the more exposed, remote quaternary carbon. Thus, whilst representative acetylide complexes [Ru(C≡C-1,4-C6H4-C≡CH)(PPh3)2Cp] ( 4 a ) and [Ru(C≡C-1,4-C6H4-C≡CH)(dppe)Cp*] ( 4 b ) reacted with the relatively small electrophiles [CN]+ and [C7H7]+ at the β-carbon to give the expected vinylidene complexes, the bulky trityl ([CPh3]+) electrophile reacted with [M(C≡C-1,4-C6H4-C≡CH)(LL)Cp’] [M(LL)Cp’=Ru(PPh3)2Cp ( 4 a ); Ru(dppe)Cp* ( 4 b ); Fe(dppe)Cp ( 4 c ); Fe(dppe)Cp* ( 4 d )] at the more exposed remote end of the carbon-rich ligand to give the putative quinoidal cumulene complexes [M{C=C=C6H4=C=C(H)CPh3}(LL)Cp’]+, which were isolated as the water adducts [M{C≡C-1,4-C6H4-C(=O)CH2CPh3}(LL)Cp’] ( 6 a–d ). Evincing the scope of the formation of such extended cumulenes from ethynyl-substituted arylvinylene precursors, the rather reactive half-sandwich (5-ethynyl-2-thienyl)vinylidene complexes [M{C=C(H)-2,5-cC4H2S-C≡CH}(LL)Cp’]BF4 ([ 7 a – d ]BF4 add water readily to give [M{C≡C-2,5-cC4H2S-C(=O)CH3}(LL)Cp’] ( 8 a – d )].  相似文献   

12.
Three different routes have been investigated for the preparation of 6-aryl-N-(1-arylethyl)thienopyrimidin-4-amines. First the possibilities of selective Suzuki reactions on 6-bromo-4-chlorothienopyrimidine were investigated. The preference for mono arylation at C-6 could be increased, in the case of Pd(PPh3)4 catalysis, by reducing the water content of the reaction, or by using less electron rich Pd-ligands. The highest selectivity was obtained with Pd(OAc)2 or Pd2(dba)3, while reactions with the more electron rich Pd(PPh3)4 and especially XPhos gave a lower mono- to dicoupled product ratio. Secondly, two alternative strategies avoiding this selectivity issue were tested. Suzuki reaction on C-6 of 6-bromothienopyrimidin-4(3H)-one (three examples) proceeded in 70-89% yield using Pd(PPh3)4 in dioxane/water. Similar conditions on 4-amino-6-bromo-thienopyrimidine (eight examples) gave 67-95% yield. The reaction could be performed with boronic acids containing nonprotected phenolic groups in the ortho, meta and para positions. By prolonging the reaction time, coupling with sterically crowded arylboronic acids was also efficient. Diarylation of 6-bromo-4-chlorothienopyrimidine gave the corresponding 4,6-diarylated derivatives in 71-80% yield depending on the nature of the arylboronic acid.  相似文献   

13.
Thirteen new 2,3,5-trisubstituted furans were prepared by the acid-catalyzed decomposition of 6,6-di-substituted 1,2-dioxan-3-ols in 57-93% yields. The reaction could be accounted for as the consequence of an oxygen-oxygen bond cleavage by acid and the migration of a phenyl group at the C-6 position followed by cyclization and elimination of a phenol. The migratory aptitude was in the order of 4-MeOC6H4- > 4-MeC6H4- > Ph- = 4-FC6H4- > 4-ClC6H4- = 4-BrC6H4- that was found from the competitive phenyl migration in the reaction of 1,2-dioxan-3-ols bearing two different substituents at the C-6 position.  相似文献   

14.
Organotin compounds R3Sn(CH2)n+2OC6H4C6H4Y (R3=Ph3, Ph2Bu; Y=H, CN; n=1-3) and RX2Sn(CH2)n+2OC6H4C6H4Y (R=Ph, Bu; Y=H, CN; X=Br, I; n=1-3) have been synthesised and characterised by 1H-, 13C-, 119Sn-NMR and Mössbauer spectroscopies. X-ray crystallography reveals tetrahedral geometries for Ph3Sn(CH2)4OC6H4C6H5 and Ph3Sn(CH2)3OC6H4C6H4CN, a six-coordinated, bromine-bridged dimeric structure for PhBr2Sn(CH2)3OC6H4C6H5 containing a mer-Br3C2OSn coordination sphere about tin and a five-coordinated monomeric structure for PhBr2Sn(CH2)3OC6H4C6H4CN. In all cases there is strong alignment of mesogenic groups in the solid-state but only PhBr2Sn(CH2)3OC6H4C6H4CN shows any indication of liquid-crystal behaviour. Wurtz polymerisation of RBr2Sn(CH2)5OC6H4C6H5 (R=Ph, Bu), both of which contain non-chelating ether functions, generated polystannanes (RR′Sn)n with Mn 2.3×105; Mw 3.0×105; Mw/Mn 1.30 and Mn 1.3×105; Mw 2.5×105; Mw/Mn 1.96, respectively, while no polymer was obtained from chelated PhBr2Sn(CH2)3OC6H4C6H5  相似文献   

15.
A catalyst derived from 2,4-pentanedionatobis(ethylene)rhodium(I), I, promoted the addition of 4-pentenal to ethylene. The reaction was accompanied by the formation of double bond migration products derived from the 4-pentenal reactant and from the 6-hepten-3-one primary product. Compound I accomplished the addition of 4-hexenal to ethylene to afford high yields of 6-octen-3-one. The fate of the aldehyde hydrogen in this transformation has been determined in experiments employing 4-hexenal-1-d as reactant. Treatment of 4-hexenal-1-d with I in CHCl3 and CDCl3 afforded 6-octen-3-one possessing >50% do molecules while the isotopic composition of recovered unexpended 4-hexenal remained >96% d1. 6-Octen-3-one products with isotopic compositions of >66% do were afforded when ethylene was introduced to reaction mixtures. The location of deuterium in 6-octen-3-one, derived from treatment of 4-hexenal-1-d with I in the absence of added C2H4, was determined to be distributed at C-1 and C-2 and at the CC bond by analysis of the 1H and 2H NMR spectra. Unexpended ethylene was recovered and was found to contain a substantial amount of deuterium. Mechanistic implications of these results are discussed.  相似文献   

16.
The sources of the migrant hydrogen atom(s) in reactions (a) and (b) in the electron impact mass spectrum of n-propyl benzoate have been investigated: (a) [C6H5CO2C3H7]+ →[C6H5CO2H]+ + C3H6; (b) [C6H5CO2C3H7]+ → [C6H5CO2H2]+ + C3H5sdot;. Deuterium labelling of the propyl group showed that, for reaction (a) at 70 eV ionizing energy 3 ± 1% of the hydrogen originates from C-1 of the propyl group, 86 ± 4% from C-2 and 11 ± 3% from C-3. The specificity of the transfer from C-2 increases as the internal energy of the fragmenting ions decreases, indicating that the results cannot be rationalized in terms of H/D interchanges between positions in the propyl group, but rather that the reaction involves specific, competing, H transfer reactions from each propyl position, in contrast to the high site specificity characteristic of the McLafferty rearrangement. Reaction (b) involves, almost exclusively, transfer of one hydrogen from C-2 and one from C-3 with only very minor participation of C-1 hydrogens. The [C6H5COOH]+ ion produced in reaction (a) fragments further to [C6H5CO]+ + OH. and the labelling results indicate some interchange of the carboxylic hydrogen with (ortho) ring hydrogens for those ions fragmenting in the first drift region. The extent of interchange is less than that observed for fragmentation of the same ion produced by direct ionization of benzoic acid or by reaction (a) in ethyl benzoate.  相似文献   

17.
Cationic dialkylaluminum and m-terphenylalkylaluminum compounds catalyze the intramolecular hydroamination of primary and secondary aminopentenes. The reaction rates are strongly dependent on the substrate and the catalyst substituents. The bulky species [Dipp1AlEt][CHB11H5I6] (Dipp1 = 2,6-Dipp2C6H3–, Dipp = 2,6-iPr2C6H3–), 4, was the most active catalyst. Although the neutral species DcpAlEt2 (Dcp = 2,6-(2,6-Cl2C6H3)2C6H3–), 7, and Dipp1AlEt2, 8, showed some catalytic activity, they were more than 25 times less reactive than their cationic counterparts [DcpAlEt][CHB11H5Cl6], 3, and 4. The cyclization of secondary benzylaminopentenes with [Et2Al][CHB11H5I6], 1, was strongly dependent on the substitution of the C-2 olefinic carbon.  相似文献   

18.
Cationic dialkylaluminum and m-terphenylalkylaluminum compounds catalyze the intramolecular hydroamination of primary and secondary aminopentenes. The reaction rates are strongly dependent on the substrate and the catalyst substituents. The bulky species [Dipp1AlEt][CHB11H5I6] (Dipp1 = 2,6-Dipp2C6H3–, Dipp = 2,6-iPr2C6H3–), 4, was the most active catalyst. Although the neutral species DcpAlEt2 (Dcp = 2,6-(2,6-Cl2C6H3)2C6H3–), 7, and Dipp1AlEt2, 8, showed some catalytic activity, they were more than 25 times less reactive than their cationic counterparts [DcpAlEt][CHB11H5Cl6], 3, and 4. The cyclization of secondary benzylaminopentenes with [Et2Al][CHB11H5I6], 1, was strongly dependent on the substitution of the C-2 olefinic carbon.  相似文献   

19.
Loss of CH, CH4, C2H4, C3H, C3H6 and C3H7 from the molecular ions of a number of 13C-labeled analogs of 4,4-dimethyl-1-pentene was studied both in normal (source) 70-eV electron impact (EI) spectra dn in metastable spectra. For loss of CH in the source, 96% of the methyl comes frm positions of 5, 5′ and 5″, while the remainder comes from position 1. In the metastable spectra, loss of C-1 (16%) and C-3 (9%) is increasing in importance. The loss of ethylene is a particular case: either C-1 or C-3 are lost with any other C-atom from positions 2,5,5′, and 5″ (8 × 10%) in the metastable spectra, the probability for simultaneous loss of C-1 and C-3 being 6%. If C-1 seems to these two positions become completely equivalent in the metastable time range. The T-values (kinetic energy release) for the different positions show small, but statisticaly different values and a small isotope effect. Loss of C3H5 (allylic cleavage) is 100% C-1, C-2 and C-3, i.e., no evidence for skeletal rearrangement is seen. This is also true for loss of C3C6 (McLafferty rearrangement) within the source, but in metastable decay the other positions gain in importance. The neutral fragment C3H appears to be the the result of consecutive loss of CH and C3H4, rather than a one-step loss of propyl radical or the inverse reactions sequence. No metastable reaction can be seen for this reaction. Decomposition of labeled C6H and C5H secondary ions occurs in an essentially random fashion.  相似文献   

20.
The thermolytic decomposition of Ph4SbSAr in organic solvents are reported. Solvent-derived products from decompositions in CCl4 and cyclohexane confirm the free radical nature of the reactions.Thermal decompositions of Ph3 (p-MeC6H4)SbSC6H4OMe-p, or Ph(p-MeC6H4)3SbSC6H4X(Xp-MeO or H) provide mixed triarylstibines,[Ph3nSb(C6H4Me-p)n (n0, 1, 2 and 3), disulphide (XC6H4SSC6H4X), monosulphides (PhSC6H4X and p-MeC6H4SC6H4X), arenes, (PhH and PhMe) and biaryls (Ph2,PhC6H4Me-p and p-MeC6H4C6H4Me-p).  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