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.
The structural investigation of two enantiomers of the phosphoroorganic selenium-containing compound [(+)C10H16NO2PSe ( 1 ) and (−)C10H16NO2PSe ( 1 ′)] was carried out. The absolute configuration at the P and C3 atoms was found to be R for compound 1 and, respectively, S for compound 1 ′.  相似文献   

2.
The reactions of IO radicals with CH3SCH3, CH3SH, C2H4, and C3H6 have been studied using the discharge flow method with direct detection of IO radicals by mass spectrometry. The absolute rate constants obtained at 298 K are the following: IO + CH3SCH3 → products (1): k1 = (1.5 ± 0.2) × 10?14; IO + CH3SH → products (2): k2 = (6.6 ± 1.3) × 10?16; IO + C2H4 →products (3): k3 < 2 × 10?16; IO + C3H6 → products (4): k4 < 2 × 10?16 (units are cm3 molecule?1 s?1). CH3S(O)CH3 and HOI were found as products of reactions (1) and (2), respectively. The present lower value of k1 compared to our previous determination is discussed.  相似文献   

3.
Rate constants for the gas-phase reactions of the four oxygenated biogenic organic compounds cis-3-hexen-1-ol, cis-3-hexenylacetate, trans-2-hexenal, and linalool with OH radicals, NO3 radicals, and O3 have been determined at 296 ± 2 K and atmospheric pressure of air using relative rate methods. The rate constants obtained were (in cm3 molecule?1 s?1 units): cis-3-hexen-1-ol: (1.08 ± 0.22) × 10?10 for reaction with the OH radical; (2.72 ± 0.83) × 10?13 for reaction with the NO3 radical; and (6.4 ± 1.7) × 10?17 for reaction with O3; cis-3-hexenylacetate: (7.84 ± 1.64) × 10?11 for reaction with the OH radical; (2.46 ± 0.75) × 10?13 for reaction with the NO3 radical; and (5.4 ± 1.4) × 10?17 for reaction with O3; trans-2-hexenal: (4.41 ± 0.94) × 10?11 for reaction with the OH radical; (1.21 ± 0.44) × 10?14 for reaction with the NO3 radical; and (2.0 ± 1.0) × 10?18 for reaction with O3; and linalool: (1.59 ± 0.40) × 10?10 for reaction with the OH radical; (1.12 ± 0.40) × 10?11 for reaction with the NO3 radical; and (4.3 ± 1.6) × 10?16 for reaction with O3. Combining these rate constants with estimated ambient tropospheric concentrations of OH radicals, NO3 radicals, and O3 results in calculated tropospheric lifetimes of these oxygenated organic compounds of a few hours. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
The novel cannabinoids (6aR, 10aR)-N-ethyl-Δ8-tetrahydrocannabinol-18-amide (15) and (6aR, 10aR, 17 RS)-N-ethyl-17-methyl-Δ8- tetrahydrocannabinol-18-amide (16) , designed as cannabinoid affinity ligands, were synthesized from the corresponding acids 11 and 12 via the N-hydroxysuccinimide esters. Amide 16 was tested in the rat and was generalized to Δ9-tetrahydrocannabinol, being 5 times less potent than the training drug. An improved synthesis of (6aR, 10aR)-17,18-didehydro-Δ8-tetrahydrocannabinol (23) is reported. As model reaction for the preparation of a tritiated Δ8-tetrahydrocannabinol, compound 23 was selectively deuterated at C(17) and C(18) in benzene/Et3N using [(C6H5)3P]3RuCl2 as catalyst.  相似文献   

5.
The reaction of thiobenzophenone (= diphenylmethanethione; 8a ) or 9H-fluorene-9-thione ( 8b ) and methyl fumarate ( 9 ) in excess PhN3 at 80° yields a mixture of diastereoisomeric thiiranes 10 and 11 (Scheme 1). A mechanism involving the initial formation of 1-phenyl-4, 5-dihydro-1H-1, 2, 3-triazole-4, 5-dicarboxylate 12 by 1, 3-dipolar cycloaddition of PhN3 and 9 is proposed in Scheme 2. The diazo compound 13 , which is in equilibrium with 12 , undergoes a further 1, 3-dipolar cycloaddition with thioketones 8 to give 2, 5-dihydro-1, 3, 4-thiadiazoles 14 . Elimination of N2 yields the thiocarbonyl ylide 15 which cyclizes to the corresponding thiirane. Desulfurization of the thiiranes 10 and 11 with hexamethylphosphorous triamide leads to the olefinic compounds 16 (Scheme 3). The crystal structures of 10a , 11a , and 16b were determined.  相似文献   

6.
The rate coefficients for the reactions of hydrogen atoms with n-C3H7Br, s-C3H7Br, n-C4H9Br, and s-C4H9Br were determined in a discharge flow-reactor at 298 K and a pressure of 4 mbar. Molecular-beam sampling and subsequent mass-spectrometric detection with electron-impact ionisation was used for the measurement of the bromo-hydrocarbon concentration. The rate coefficients obtained are (in 1010 cm3 mol−1 s−1): 2.3±1.2 for n-C3H7Br, 2.3±1.2 for s-C3H7Br, 2.4±1.2 for n-C4H9Br, and 2.8±1.4 for s-C4H9Br. The results are compared with predictions from bond-energy bond-order (BEBO) calculations, where a reasonable agreement is found. Furthermore, also by BEBO calculations, the relative importance of bromine abstraction as compared to hydrogen abstraction is estimated. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 721–727, 1998  相似文献   

7.
Nonisothermal kinetics of the solid‐solid phase transition in (n‐C10H21NH3)2ZnCl4(C10Zn), (n‐C16H33NH3)2ZnCl4(C16Zn) and their binary system were determined by Kissinger and Ozawa methods from DSC measurements. The activation energy Ea of the binary system shows a waving dependence on WC10Zn%, which is caused by not only an intermediate (C10H21NH3)(n‐C16H33NH3)ZnCl4 but also three solid solution ranges (α, β, γ) in the phase diagram of C10Zn‐C16Zn. The variations of the layer d‐spacing are also convenient for the above result.  相似文献   

8.
The addition of phosphite anions and of tris(trimethylsilyl) phosphite (P(OSiMe3)3) to N-glycosyl-C-arylnitrones was examined. While these nitrones proved inert towards the phosphite anions, they reacted with P(OSiMe3)3 under catalysis by Lewis acids. Thus, P(OSiMe3)3 reacted with the crystalline (Z)-N-glycosylnitrones 2 and 8 to give the optically active N-hydroxy-α-aminophosphonic acids 4 and 10 , respectively, and hence the α-aminophosphonic acids 5 and 11 in yields up to 92% and with an enantiomeric excess (e.e.) up to 97% (Scheme 1). The absolute configuration of the phosphonates depend upon the nature and – in one case – upon the quantity of the catalyst (Figure). Upon catalysis by HCIO4 or Zn(OTF)2, p(OSiMe3)3 added to 2 to give, in both cases, the (+)-(R)-phenylphosphaglycine 5 (optical purity 79–84 and 90–93%, resp.). The optical purity (o.p.) was hardly influenced by the amount of these catalysts (0.02-;1 equiv.). However, catalysis by ZnCl2 gave, with trace quantities of the catalyst, (–)-(S)- 5 (o.p. 79%), while an equimolar amount of ZnCl2 yielded (+)-(R)- 5 (o.p. 82%). The HClO4-catalyzed addition of P(OSiMe3)3 to the nitrone 14 (Scheme 2) led to (+)-(R)-N-hydroxyphosphavaline 15 (78%) and hence to (–)-(R)-phosphavaline 16 (71% from 14 e.e. 95%). Under conditions leading from the nitrones 2 , 8 , 14 , and 20 (Schemes 1 and 2) predominantly to (R)-α-aminophosphonic acids, the addition of P(OSiMe3)3 to nitrone 18 , possessing a benzyloxy substituent as an additional potential ligand for the catalyst, gave (S)-phosphaserine 19 . The addition of P(OSiMe3)3 to the nitrone 20 , catalyzed by Zn(OTf)2, led to (+)-(R)-N-hydroxyphosphamehionine 21 (71%, e.e. 77%) and hence to (–)-(R)-phosphamethionine 22 (77% from 20 , e.e. 79%). Catalysis by trace quantities of ZnCl2 gave (+)-(S)- 22 (85%, e.e. 61%). The enantiomerically pure aminophosphonic acids 5 , 11 , and 16 were obtained by recrystalliztion. The e.e. of the N-hydroxyaminosphosphonic acids 10 , 15 , and 21 and the aminophosphonic acids 5 , 11 , 16 , and 22 were determined by the HPLC analysis of the dimethyl N-naphthoyl-α-aminophosphonats 7 , 13 , 17 , and 23 , on a chiral stationary phase.  相似文献   

9.
Reactions of [Cu(NCMe)4]+ with stoichiometric amount of diphosphine R2P–(C6H4)n–PR2, (R = NC4H4, n = 1; R = Ph, n = 1, 2, 3) or tri‐phosphine 1, 3, 5‐(PPh2–C6H4–)3–C6H3 ligands give the corresponding di‐ or trinuclear copper(I) acetonitrile‐phosphine complexes 1 – 5 . Substitution of the labile acetonitrile groups with chelating aromatic diimines – 2, 2′‐bipyridine (bpy), 1, 10‐phenanthroline (phen), 5, 6‐dimethyl‐1, 10‐phenanthroline (dmp), 5, 6‐dibromo‐1, 10‐phenanthroline (phenBr2) – gives the corresponding substituted compounds 6 – 16 . In all complexes 1 – 16 each central CuI atom has tetrahedral configuration completed with two N‐ and two P‐donor groups. The compounds obtained were characterized using elemental analysis, ESI‐MS, X‐ray crystallography, and NMR spectroscopy. All phosphine‐diimine compounds 6 – 16 are photoluminescent at room temperature both in dichloromethane solution and in solid state (λex = 385 nm). In CH2Cl2 solution the maxima of emission bands are found in a range 540–640 nm, and in solid in a similar range 538–620 nm. Emission of 6 – 16 is assigned to the triplet excited state dominated by the charge transfer transitions with contribution of the MLCT character.  相似文献   

10.
The synthesis of 1- and 2-aryl-substituted (aryl = Ph, 4-NO2? C6H4, and 4-MeO? C6H4) 4, 6, 8-trimethylazulenes ( 4 and 3 , respectively) in moderate yields by direct arylation of 4, 6, 8-trimethylazulene ( 8 ) with the corresponding arylhydrazines 13 in the presence of CuIIions in pyridine (see Scheme 4) as well as with 4-MeO? C6H4Pb(OAc)3 ( 16 ) in CF3COOH (see Scheme 5) is described. With 13 , also small amounts of 1, 2- and 1, 3-diarylated azulenes (see 14 and 15 , respectively, in Scheme 4) are formed. The 4-methoxyphenylation of 8 with 16 yielded also the 1, 1′-biazulene 17 in minor amounts (see Scheme 5). 4, 6, 8-Trimethyl-2-phenylazulene ( 3a ) was also obtained as the sole product in moderate yields by the reaction of sodium phenylclopentadienide ( 1a ) with 2, 4, 6-trimethylpyrylium tetrafluoroborate ( 2 ) in THF (Scheme 1). The attempted phenylation of 8 as well as of azulene ( 9 ) itself with N-nitroso-N-phenylacetamide ( 10 ) led only to the formation of the corresponding 1-(phenylazo)-substituted azulenes 12 and 11 , respectively (Scheme 3).  相似文献   

11.
Effect of para-substituents in the ethylene (E) copolymerization with 1-decene (DC), 1-dodecene (DD), and with 2-methyl-1-pentene (2M1P) using a series of Cp*TiCl2(O-2,6-iPr2-4-R-C6H2) [R=H ( 1 ), tBu ( 2 ), Ph ( 3 ), CHPh2 ( 4 ), CPh3 ( 5 ), SiMe3 ( 6 ), SiEt3 ( 7 ), and newly prepared 4-tBuC6H4 ( 8 ) and 3,5-Me2C6H3 ( 9 )]-MAO catalyst systems has been studied. The activities in these copolymerization reactions were affected by the para-substituent, and the SiMe3 ( 6 ), SiEt3 ( 7 ) and 3,5-Me2C6H3 ( 9 ) analogues showed the higher activities at 50 °C in the E copolymerization reactions with DC (1.06–1.44×106 kg-polymer/mol-Ti⋅h), DD (1.04–1.88×106 kg-polymer/mol-Ti⋅h) than the others, whereas no significant differences were observed in the comonomer incorporations. Complexes 6 and 7 also showed the higher activities at 50 °C in the E/2M1P copolymerization, and the 2M1P incorporation was affected by the para-substituent and the polymerization temperature; complex 9 showed better 2M1P incorporation at 25 °C.  相似文献   

12.
Rate constants for the gas phase reactions of O3 and OH radicals with 1,3-cycloheptadiene, 1,3,5-cycloheptatriene, and cis- and trans-1,3,5-hexatriene and also of O3 with cis-2,trans-4-hexadiene and trans -2,trans -4-hexadiene have been determined at 294 ± 2 K. The rate constants determined for reaction with O3 were (in cm3 molecule-1s?1 units): 1,3-cycloheptadiene, (1.56 ± 0.21) × 10-16; 1,3,5-cycloheptatriene, (5.39 ± 0.78) × 10?17; 1,3,5-hexatriene, (2.62 ± 0.34) × 10?17; cis?2,trans-4-hexadiene, (3.14 ± 0.34) × 10?16; and trans ?2, trans -4-hexadiene, (3.74 ± 0.61) × 10?16; with the cis- and trans-1,3,5-hexatriene isomers reacting with essentially identical rate constants. The rate constants determined for reaction with OH radicals were (in cm3 molecule?1 s?1 units): 1,3-cycloheptadiene, (1.31 ± 0.04) × 10?10; 1,3,5-cycloheptatriene, (9.12 × 0.23) × 10?11; cis-1,3,5-hexatriene, (1.04 ± 0.07) × 10?10; and trans 1,3,5-hexatriene, (1.04 ± 0.17) × 10?10. These data, which are the first reported values for these di- and tri-alkenes, are discussed in the context of previously determined O3 and OH radical rate constants for alkenes and cycloalkenes.  相似文献   

13.
The dependency of the cross-coupling of orthogonally C-protected dialkynyl monosaccharides on the nature of the coupling partners has been studied. The required dialkyne 5 was synthesized from levoglucosan in six steps and 39% overall yield and transformed into 7 , 10 , 12 , 13 , and 14 by orthogonal C-deprotection and bromination (Scheme 1). Optimization of the conditions of their cross-coupling to 16 showed that yields were higher for the coupling of the propargylic bromoalkyne 10 than for the homopropargylic bromoalkyne 14 (Scheme 2). Deprotection of 16 gave the nano-crystalline dimer 20 . To obtain more highly crystalline products, the monomers 7 and 13 were coupled with 1-iodo-4-nitrobenzene to the arylated monomers 21 and 24 (Scheme 3). The 4-NO2C6H4 substituent lowered the yield of the dimerizations to the mono- and diarylated dimers 26 – 28 (Scheme 4) but had no effect on crystallinity.  相似文献   

14.
In the course of our investigations on the development of cardiovascular agents, 3-butyl-2-[2′-(2H-tetrazol-5-yl)bipheny]-4-yl]methyl-2H-1, 2, 4-benzothiadiazine 1, 1-dioxide ( 2 ) was considered as a potential angiotensin II antagonist on the basis of bioisosteric replacement of the quinazoline ring of compound 1 with a 1, 2, 4-benzothiadiazine 1, 1-dioxide ring system. Alkylation of 6 with 4 afforded 7 and 8 in 24% and 28% yields, respectively. An attempt to remove the trityl group of compounds 7 and 8 under acidic condition gave the ring opened products 9 and 11 in 28% and 36% yields, respectively. However, compounds 2 and 10 were obtained in 46% and 85% yields when compounds 7 and 8 were refluxed in methanol. Preliminary assays of compounds 9 and 11 against angiotensin II receptors revealed weak activity with IC50 values of 3.6 μM and 5.4 μM, respectively. Compound 10 (IC50 = 87 nM) exhibited stronger binding affinity than compound 2 (IC50 = 750 nM).  相似文献   

15.
Rate constants have been determined at 296 ± 2 K for the gas phase reaction of NO3 radicals with a series of aromatics using a relative rate technique. The rate constants obtained (in cm3 molecule?1 s?1 units) were: benzene, <2.3 × 10?17; toluene, (1.8 ± 1.0) × 10?17; o? xylene, (1.1 ± 0.5) × 10?16; m? xylene, (7.1 ± 3.4) × 10?17; p? xylene, (1.4 ± 0.6) × 10?16; 1,2,3-trimethylbenzene, (5,6 ± 2.6) × 10?16; 1,2,4-trimethylbenzene (5.4 - 2.5) × 10?16; 1,3,5-trimethylbenzene, (2.4 ± 1.1) × 10?16; phenol, (2.1 ± 0.5) × 10?12; methoxybenzene, (5.0 ± 2.8) × 10?17; o-cresol, (1.20 ± 0.34) × 10?11; m-cresol, (9.2 ± 2.4) × 10?12; p-cresol, (1.27 ± 0.36) × 10?11; and benzaldehyde, (1.13 ± 0.25) × 10?15. These kinetic data, together with, in the case of phenol, product data, suggest that these reactions proceed via H-atom abstraction from the substituent groups. The magnitude of the rate constants for the hydroxy-substituted aromatics indicates that the nighttime reaction of NO3 radicals with these aromatics can be an important loss process for both NO3 radicals and these organics, as well as being a possible source of nitric acid, a key component of acid deposition.  相似文献   

16.
This study shows the results of application liquid chromatography‐tandem mass spectrometry (LC/MS/MS) for assay of the content of α‐tocopherol and coenzyme Q10 in bee products of animal origin, i.e. royal jelly, beebread and drone homogenate. The biological matrix was removed using extraction with n‐hexane. It was found that drone homogenate is a rich source of coenzyme Q10. It contains only 8 ± 1 µg/g of α‐tocopherol and 20 ± 2 µg/g of coenzyme Q10. The contents of assayed compounds in royal jelly were 16 ± 3 and 8 ± 0.2 µg/g of α‐tocopherol and coenzyme Q10, respectively. Beebread appeared to be the richest of α‐tocopherol. Its level was 80 ± 30 µg/g, while the level of coenzyme Q10 was only 11.5 ± 0.3 µg/g. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

17.
1H-NMR studies were carried out for solution of amphiphilic betaine ester derivatives (of the general formula (CH3)3N+CH2COOC n H2n+1Cl (V-n), wheren=10, 12, 14, and 16) andn-dodecyltrimethylammonium chloride (I-12). The spectra were taken at concentrations above and below critical micelle concentrations and chemical shifts were analyzed. It was stated that micelles are hydrated at the depth of the two CH2 groups in the case ofV-n and the CH2COO group in the case ofI-12. Therefore, the CH2COO group during the micellization behaves as if it were CH2CH2 group.  相似文献   

18.
1-C-Nitroglycals. Preparation and Reaction with Some Nitrogen Nucleophiles Acetylation of the 1-deoxy-1-nitromannopyranoses 2 and 6 was accompagnied by spontanous β-elimination to give the 1-C-nitroglucals 3 and 7 , respectively, while acetylation of the gluco- and galacto-configurated 1-deoxy-1-nitropyranoses 8 and 14 gave the acetates 9 and 15 , respectively (Scheme 1). The acetylation of the ribo- and arabino-configurated 1-deoxy-1-nitrofuranoses 19 and 21 also occurred without β-elimination to give the acetates 20 and 22 , respectively (Scheme 2). Mild base treatment of the previously described O-acetylnitro-β-D -glucose 4 , the O-acetylnitro-β-D -pyranoses 9 and 15 , and the O-acetylnitro-β-D -furanoses 17 , 20 , and 22 gave the 1-C-nitroglycals 3 , 10 , 16 , 18 and 23 , respectively (Scheme 1 and 2). The previously obtained 1-C-nitroglucal 3 was deacetylated by treatment with MeOH in the presence of KCN or sodium m-nitrophenolate to give the free nitroglucal 5 . Deacetylation of the benzylidene protected 1-C-nitroglucal 10 (MeOH, NaOMe) gave the 4,6-O-benzylidene-1-C-nitroglucal 11 and traces of the 2-O-methyl-1-C-nitromannoses 12 and 13 . The UV, IR, 1H-NMR and 13C-NMR spectra of the 1-C-nitroglycals are discussed. In solution, the 1-C-nitroglycals 1 , 5 , 7 , 10 , 11 , and 16 adopt approximately a 4H5? and 3 a flattened 4H5 conformation. The structure of 5 was established by X-ray analysis. In the solid state, 5 adopts a sofa conformation, which is stabilized by an intramolecular H-bond. The β-addition of NH3 to the 1-C-nitroglucals 7 and 10 was followed by an O→ N acetyl migration to give exclusively anomeric pairs of the N-acetyl-1-nitromannosamine derivatives 24 / 25 and 26/27 , respectively (Scheme 3). The β-addition of methylamine, octadecylamine, and tryptamine to the 1-C-nitroglucal 11 also stereoelectronically controlled and gave the crystalline N-alkyl-1-nitromannosamines 28 , 29 , and 30 , respectively. The stereoelectronically controlled β-addition of NH3 to the 1-C-nitrogalactal 16 , followed by acetylation, yielded exclusively the talosamine derivative 31 , while the reversible β-addition of azide ions to 16 gave the anomeric 2-azido-1-nitrogalactoses 32 and 33 . The β-addition of azide ions to the 1-C-nitroglucal 1 led to the 2-azido-1-nitromannose 34 . In the presence of excess formaldehyde, this addition was followed by a Henry reaction. Chromatography of the crude product was accompagnied by solvolytic removal of the NO2 group to give the 3-azidomannoheptulose 35 in high yields (Scheme 4).  相似文献   

19.
The addition of dienes, diazomethane, and carbenoids to the manno- and ribo-configurated thio-γ-O-lactones 1 and 2 was investigated. Thus, 1 (Scheme 1) reacted with 2,3-dimethylbutadiene (→ 4 , 73%), cyclopentadiene (→ 5a/b 1:1, 70%), cyclohexa- 1,3-diene (→ 9a/b 2:3, 92%), and the electron-rich butadiene 6 (→ 7a/b 3:1, 82%). Wheras 5a/b was separated by flash chromatography, 7a/b was desilylated leading to the thiapyranone 8 . Selective hydrolysis of one isopropylidene group of 9a/b and flash chromatography gave 10a and 10b . The structures of the adducts were elucidated by X-ray analysis ( 4 ), by NOE experiments ( 4 , 5a , 5b , 7a/b , 10a , and 10b ), and on the basis of a homoallylic coupling ( 7a/b ). The additions occurred selectively from the ‘exo’ -side of 1 . Only a weak preference for the ‘endo’-adducts was observed. Hydrogenation of 9a/b with Raney-Ni (EtOH, room temperature) gave the thiabicyclo [2.2.2]octane 11 . Under harsher conditions (dioxane, 110°), 9a/b was reduced to the cyclohexyl ß-D C-glycoside 12 which was deprotected to 13 . X-Ray analysis of 13 proved that the desulfuration occurred with inversion of the anomeric configuration. The regioselective addition of the dihydropyridine 14 to 1 (Scheme 2) and the methanolysis of the crude adduct 15 gave the lactams 16a (32%) and 16b (38%). Desilylation of 15 with Bu4NF · 3H2O, however, gave the unsaturated piperidinedione 17 (92%) which was deprotected to the tetrol 18 (65%). Similarly, 2 was transformed via 19 (62%) into the triol 20 (74%). The cycloaddition of 1 with CH2N2 (Scheme 3) gave a 35:65 mixture of the 2,5-dihydro- 1,3,4-triazole 21 and the crystalline 4,5-dihydro 1,2,3-triazole 22 . Treatment of 21 and 22 with base gave the hydroxytriazoles 23 and 24 , respectively. The structure of 24 was established by X-ray analysis. The triazole mixture 21/22 was separated by prep. HPLC at 5°. At room temperature, 21 already decomposed (half-life 21.6 h) leading in CDCI3 solution to a complex mixture (containing ca. 20–25% of the spirothiirane 27 and ca. 7–10% of its anomer) and in MeOH solution exclusively to the O,O,S-ortholactone 26 . Crystals of 22 proved be stable at 105°. Upon heating in petroleum ether at 100°, 22 was transformed into a ca. 1:1 mixture of 27 and the enol ether 28 . The reaction of 1 with ethyl diazoacetate (Scheme 4) in the presence of Rh2(OAc)4. 2H2O gave the unsaturated esters 29 (33%) and 30 (26%), whereas the analogous reaction with diethyl diazomalonate afforded the spirothiirane 31 (68%) and the enol ether 32 (29%). Complete transformation of 31 into 32 was achieved by the treatment with P(NEt2)3. Similary, 33 (69%) was prepared from 2 .  相似文献   

20.
After prolonged refluxing of 19-tosyloxy-16α,17α-cyclohexanopregn-5-en-3β-ol-20-one (3) with NaI in 2-propanol, the initially formed 19-iodo derivative (4) undergoes supraface migration of the CH2I group from the C(10) atom to the C(6) atom, probably through involvement of a homoallyl cation. The resulting 6β-iodomethyl-16α,17α-cyclohexano-19-norpregn-5(10)-en-3β-ol (5) was transformed in three steps into 6α-methyl-16α,17α-cyclohexano-19-norprogesterone (6α-methyl-19-nor-D′ 6-pentarane,8). The transformation of compound5 into the target product8 also gave a side product, a pentarane with aromatic ringA (10), which was isolated and characterized by spectroscopic methods. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1688–1691, September, 1997.  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