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.
Studies on the syntheses of 18′-epi-4′-deoxo-4′-epivinblastine (IX, R = CO2CH3; R1 = H), 18′-decarbomethoxy-18′-epi-4′-deoxo-4′-epivinblastine (IX, R = R1 = H) and related analogues are described. The synthetic method employs a coupling reaction involving chloroindolenine derivatives of the cleavamine series (for example, III) with vindoline (V) under acidic conditions. The complete structures, including absolute configuration, of the resulting dimers are established by a combination of chemical and spectroscopic techniques, including X-ray analysis.  相似文献   

2.
Some new oxygen–sulfur, multibenzo macrocyclic ligands containing amide groups have been prepared using the macrocyclization process with the reaction of 2,2′-thiobis-[4-methyl(2-aminophenoxy)phenyl ether] as a symmetrical diamine with appropriate dicarboxylicacid dichlorides in moderate yields. This macrocyclization led to the formation of di- and tetramide macrocycles. These reactions were routinely carried out at ambient temperature in CH2Cl2 as solvent in high dilution without template effect conditions. It is found that sulfur the atom affects the rigidity of the macrocycles and diastereotopicity of nuclei in the ring of these series of macrocyclic compounds.  相似文献   

3.
A detailed study of the reaction of catharanthine N-oxide and vindoline has been carried out employing various conditions. Under optimum conditions, which involve low temperatures and trifluoroacetic anhydride as reagent, 3′, 4′-dehydrovinblastine (XIII, R = COOCH3), in reasonable yields is essentially the exclusive product. However two additional products, 18′ (epi)- 3′, 4′-dehydrovinblastine (XIV, R = COOCH3) and 1′-hydroxy- 3′, 4′-dehydrovinblastine (XVI, R = COOCH3) are also often isolated. The reaction, which follows the course of a Polonovski-type fragmentation process, has been extended to the N-oxide derivatives of dihydrocatharanthine and decarbomethoxycatharanthine to provide again a series of bisindole alkaloid derivatives, also vinblastines. A mechanistic rationale is provided to explain the various results obtained.  相似文献   

4.
Reaction of Metal and Metalloid Compounds with Polyfunctional Molecules. XLI. Synthesis of 9-Bora-10-oxa-phenanthrenes, Bis(2′-oxy-terphenylyl)halogenoboranes, and Bis(9-bora-10-oxa-phenanthrenyl)oxides The reaction of 2′-hydroxy-m-terphenyl and 2-hydroxy-biphenyl resp. with n-C4H9Li and Na resp. and subsequent addition of trihalogenoboranes yields 9-halogeno-9-bora-10-oxa-phenanthrenes 1, 2, 5 and A besides 9-2′-oxy-m-terphenylyl-9-bora-10-oxa-phenanthrene 3 and bis-(9-bora-10-oxa-phenanthrenyl)oxides 4 and 6 . Under somewhat different reaction conditions one obtains also the bis(2′-oxy-terphenylyl)halogenoboranes 7 and 8 , tris(2-oxy-biphenylyl)-borane 9 and t-butyl-chloro-(2-oxy-biphenylyl)borane 10. 11 results from reaction of 2 with lithated hexamethyldisilazane. The compounds are characterized analytically and spectroscopically (1H, 11B-nmr spectra; mass spectra). For 4 and 6 X-ray analyses were performed.  相似文献   

5.
4-Vinylpyridine ( 1a ) combines with 3 moles of dienophilic N-alkyl-maleinimides ( 2 ) in the presence of polymerization inhibitors. The first step of the reaction probably consists of 1:1-addition with participation of an aromatic double bond, comparable to the analogous behavior of styrene and its derivatives under similar conditions. The unstable intermediates 3 , like other Schiff bases (imines), add 2 further moles of the N-alkyl-maleinimides forming the spiro compounds 4 . These are split in an acidic medium into the N-alkyl-5,6,7,8-tetrahydroisoquinoline-7,8-dicarboximides ( 5 ), and N,N′-dialkyl-2-butene-1,2,3,4-tetracarboxylic 1,2,:3,4-diimides ( 6 ). LiAlH4 reduction of these two types of compounds leads to N-alkyl-1 H-(3,4-d)-pyrrolo-2,3,3a,4,5,9b-hexahydroisoquinolines ( 7 ) and to N,N′dialkyl-3,3′-bipyrrolidyls ( 8A ) and their dehydro-products 8B , respectively. From the reaction of 2-vinylpyridine ( 1b ) with N-alkyl-maleinimides ( 2 ) the 1:2-addition products 9 can be isolated in the presence of polymerization inhibitors, which are derivatives of N-alkyl-5,6,7,8-tetrahydroquinoline-5,6-dicarboximides ( 9 ). This again corresponds to the reaction type of cycloadditions with styrene. Furthermore 1:3 adducts are formed which according to 1H- and 13C-NMR.-data most likely have the structure 10 , representing a new type of cycloaddition involving the pyridine nitrogen.  相似文献   

6.
Diaminomethylene- and aminomethylthiomethylenehydrazones [2] of cyclic ketones 1–8 readily reacted with ethoxymethylenemalononitrile to give spiro[cycloalkane-1,2′-[1,2′,4′]triazolo[1,5′-c]pyrimidine-8′-carbonitrile] derivatives 12–19 through the electrocyclic reaction of the initially formed condensation products 26 in moderate to high yields. The spiro[cyclopentanetriazolopyrimidine] derivatives underwent ring-opening at the cycloalkane moiety upon heating in solution to give 2-alkyl-5-substituted-[1,2,4]triazolo[1,5-c]pyrimidine-8′-carbonitriles 20–23 . When an alkyl substituent was introduced into the cyclopentane ring, cleavage of the spiro compounds occurred preferentially at the cyclopentane moiety between the spiro carbon and the more branched one. In contrast, the cyclohexane ring, especially of spiro-5-amino-triazolopyrimidines 17 and 18 strongly resisted to ring-opening under similar conditions, but those of 5-methylthiotriazolopyrimidines 14 gave up to 17 percent of cleavage after prolonged heating in hot ethanol. 2-t-Butyl-5-methylthio-2,3-dihydro[1,2,4]triazolo[1,5-c]pyrimidine-8-carbonitrile 25 [R3 = C(CH3)3] was highly susceptible to the cleavage even at room temperature and produced the corresponding 2-unsubstituted triazolopyrimidine 24 with loss of the t-butyl group.  相似文献   

7.
β-D-Arabinofurano[1′,2′:4,5]oxazolo-s-triazin-4-one-6-thione ( 7b ) and its t-butyldimethylsilyl protected counterpart 7a were synthesized by treating the appropriate 2-amino-β-D-arabinofurano[1′,2′:4,5]-2-oxazoline with ethoxycarbonyl isothiocyanate. These 2,2′-anhydro-s-triazine nucleosides were then subjected to alkylation under similar reaction conditions. Alkylation of 3′,5′-bis(O-t-butyldimethylsilyl)-β-D-arabinofurano[1′,2′:-4,5]oxazolo-s-triazin-4-one-6-thione ( 7a ) provided the targeted S-alkylated nucleosides, i.e., the C6-SCH3 ( 9a ), C6-SCH2-CH = CH2 ( 10a ), and C6-S-CH2-C = CH ( 11a ), in reasonable yields. Attempted deprotection of these nucleosides failed. In order to circumvent this problem, 7b was alkylated with the same reagents. In each case, instead of the expected S-alkylated anhydronucleosides, a mixture of the 5-N-alkylanhydro-s-triazine-4,6-dione and 5-N-alkylanhydro-s-triazin-4-one-6-thione derivatives were obtained. The 2,2′-anhydro linkage of 7a was also found to be more stable than the s-triazine ring to mild base. Basic conditions displaced the C6-sulfur substituent and eventually caused ring opening of the s-triazine aglycone.  相似文献   

8.
9.
Base-catalyzed rearrangements of both individual 4-(acylmethylidene)butenolides and their mixtures prepared by condensation of citraconic anhydride with various phosphoranes occur successfully only in the presence of 5.2% MeONa in MeOH (molar ratio MeONa: substrate ≤ 10: 1, room temperature, 1–2 h). Under these conditions, the yields of 2-cinnamoyl-4-methylcyclopent-4-ene-1,3-dione (coruscanone B) and 2-acetyl-4-methylcyclopent-4-ene-1,3-dione are 56 and 65%, respectively. With a considerable increase in the reaction temperature or the molar ratio MeONa: substrate, formal addition of MeOH to the C(4)=C(5) double bond of these triketones becomes an appreciable (or predominant) process. A reaction of coruscanone B with CH2N2 in ether gives coruscanone A as a ~3: 2 mixture of (Z)- and (E)-methyl enolates (43%); other products (10%) result from the expansion and aromatization of the five-membered ring of the triketone. The simplest analog of coruscanone B, 2-acetyl-4-methylcyclopent-4-ene-1,3-dione, reacts with CH2N2 in a similar way.  相似文献   

10.
The reaction of triethyl 3-methyl-4-phosphonobut-2-enoate (1) with three alkyl halides AlkX (Alk=Pri, Me2CHCH2CH2, andc-C5H9; X=Br, I) in the system KOH(solid)∩DMF∩Bun 4NBr at ≈20°C gives exclusively products of alkylation at C(2) with Δ2 and/or Δ3 position of the double bond. Under the same conditions, the reaction of 1 with MeI gives a mixture of products with different substitution patterns. Only the use of an ion pair extraction technique affords 2-methyl-Δ2-products selectively, albeit in rather moderate yields. The Horner—Emmons olefination of PhCHO with the resulting phosphonates gives ethyl 2-alkyl-3-methyl-5-phenylpenta-2,4-dienoates in high yields. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1839–1844, October, 1997.  相似文献   

11.
The reaction of allyl palladium(II) chloride dimer and 4,4′‐bis(RfCH2OCH2)‐2,2′‐bpy, 1a–b , in the presence of AgOTf resulted in the synthesis of cationic palladium complex, [Pd(η3‐allyl)(4,4′‐bis‐(RfCH2OCH2)‐2,2′‐bpy)](OTf), 2a–b where Rf = C9F19 ( a ), C10F21 ( b ), respectively. The reaction of [PdCl2(CH3CN)] or K2PdCl4 with 1b , gave rise to the synthesis of [PdCl2(4,4′‐bis‐(C10F21CH2OCH2)‐2,2′‐bpy)], 3b . The quantitatively determined solubility curves of 2a–b and 3b in DMF indicated dramatic increase of solubility for 2a – b and 3b from ?40 to 90 °C. The catalyst‐recoverable Pd‐catalyzed Heck/Sonogashira reactions were successfully achieved in DMF with microwave‐assistance. The cationic Pd‐catalyzed Heck arylation of methyl acrylate was selected to demonstrate the feasibility of recycling 2a–b using DMF as a solvent under microwave‐assisted thermomorphic conditions. At the end of each cycle, the product mixtures were cooled, and then the catalysts were recovered by decantation. The Heck arylation catalyzed by 2b under microwave‐assisted mode exhibited good recycling results favoring the trans product. To our knowledge, this is the first example of cationic Pd‐catalyzed Heck arylation under microwave‐assisted thermomorphic conditions. Additionally, recoverable 3b ‐catalyzed Sonogashira reactions were also achieved successfully in DMF. The reactions under microwave‐assistance gave much better results in yield and in efficiency than that under conventional thermal heating.  相似文献   

12.
The photo-oxygenation of 2-(methoxymethylidene)adamantane ( 3 ) creates a zwitterionic peroxide which may be captured by acetaldehyde to give the corresponding pair of diastereoisomeric tricyclo[3.3.1.13,7]decane-2-spiro-6′ -(3′ -methyl-5′ -methoxyl′, 2′, 4′ -trioxanes) ( 4 ). Ease of capture depends strongly on solvent polarity and temperature. When these are low, yields of trioxine are high (~ 80%). Conversely, 1,2-dioxetane formation is favoured at high temperature and solvent polarity. 2-(Phenoxymethylidene)adamantane ( 5 ), on photo-oxygenation, only gives the corresponding 1,2-dioxetene, even in the presence of acetaldehyde. From a Hammett, study of the-oxygenation of the enol ether 5 and p-methoxy, p-methyl, p-chloro and m-chloro derivatives, 9, 11, 13 , and ( 15 ), a good linear relation was found between substituent constants and oxygenation rates which yielded reaction constants (ρ) of 2.59, ?2.40, ?1.09, and ?0.90 in benzene, AcOET, CH2Cl2, and MeOH respectively. This data to the formation of a zwitterionic peroxide which enjoys stabilization from its won substituents and by competing solvation and further explains the predominance of dioxetane to the detriment of trioxane formation.  相似文献   

13.
Condensations between 3-X-2,4-dimethylpyrroles (X = H, CH3, C2H5, and CO2C2H5) and acyl chlorides gave derivatives of 3,5,3′,5′-tetramethylpyrromethene (isolated as their hydrochloride salts): 6-methyl, 6-ethyl, 4,4′,6-trimethyl, 4,4′-diethyl-6-methyl, and 4,4′-dicarboethoxy-6-ethyl derivatives for conversion on treatment with boron trifluoride to 1,3,5,7-tetramethylpyrromethene–BF2 complex (TMP–BF2) and its 8-methyl (PMP–BF2), 8-ethyl, 2,6,8-trimethyl (HMP–BF2),2,6,-diethyl-8-methyl (PMDEP–BF2), and 2,6-dicarboethoxy-8-ethyl derivatives. Chlorosulfonation converted, 1,3,5,7,8-pentamethylpyrromethene–BF2 complex to its 2,6-disulfonic acid isolated as the lithium, sodium (PMPDS–BF2), potassium, rubidium, cesium, ammonium, and tetramethylammonium disulfonate salts and the methyl disulfonate ester. Sodium 1,3,5,7-tetramethyl-8-ethylpyrromethene-2,6-disulfonate–BF2 complex was obtained from the 8-ethyl derivative of TMP–BF2. Nitration and bromination converted PMP–BF2 to its 2,6-dinitro-(PMDNP–BF2) and 2,6-dibromo- derivatives. The time required for loss of fluorescence by irradiation from a sunlamp showed the following order for P–BF2 compounds (10−3 to 10−4 M) in ethanol: PMPDS–BF2, 7 weeks; PMP–BF2, 5 days; PMDNP–BF2, 72 h; HMP–BF2, 70 h; and PMDEP–BF2, 65 h. Under similar irradiation PMPDS–BF2 in water lost fluorescence after 55 h. The dibromo derivative was inactive, but each of the other pyrromethene–BF2 complexes under flashlamp excitation showed broadband laser activity in the region λ 530–580 nm. In methanol PMPDS–BF2 was six times more resistant to degradation by flashlamp pulses than was observed for Rhodamine-6G (R-6G). An improvement (up to 66%) in the laser power efficiency of PMPDS–BF2 (10−4 M in methanol) in the presence of caffeine (a filter for light <300 nm) was dependent on flashlamp pulse width (2.0 to 7.0 μsec).  相似文献   

14.
Supported Organometallic Complexes. IV. Structural Investigations on Neutral and Cationic (Ether-phosphane)palladium(II) Complexes . Reaction of the (ether-phosphane) ligands PhP(R)CH2—D ( 2a?c ) [D=CH2OCH3: R=Ph ( a ), (CH2)3Si(OCH3)3 ( b ), (CH2)3SiO3/2 ( b ′); D= R=(CH2)3Si(OCH3)3 ( c ), (CH2)3SiO3/2 ( c ′)] with Cl2Pd(COD) ( 1 ) results in the formation of Cl2Pd(P — O)2 ( 3a?c ). Cleavage of Cl? from 3 with AgSbF6 yields the cationic, monochelated complexes [ClPd(P — O)(P ∩ O)]+ ( 4 a—c ) (P — O: η1-P-coordinated; P ∩ O: η2-O ∩ P-coordinated). 4 a crystallizes in the monoclinic space group P21/c with the lattice constants a=1 062,4(2), b=1 912,2(4) und c=1 635,5(3) pm, β=101,22(3)° and Z=4 (R=0,0341; Rw=0,033). With water 3 b, c and 4 b, c are subjected to polycondensation to give the supported complexes 3 b′, c′, 4 b′, c ′. The structure 3 b′, c′, 4 b′, c ′ is investigated by comparison of their 31P CP MAS data with the 31P{1H} NMR spectra of their soluble precursors 3 b, c, 4 b, c . 13C CP MAS NMR spectra of 3 b′, c ′ and 4 b′, c ′ indicate η1-P- and η2-O ∩ P-coordination of the ligands. The polysiloxane network of 4 b ′ was inspected by contact time variation of the 29Si CP MAS NMR spectra and it appeared that 77% of the Si—O units are crosslinked, corresponding to a ratio T4:T3:T1=67:100:10.  相似文献   

15.
Pentafluorosulfanylamines and Sulfanylammonium Salts . From the addition of HF to sulfurtetrafluorideimides N-alkylpentafluorosulfanylamines RNHSF5(2a, 2b: R=CH3, C2H5) are obtained in quantitative yield. N, N-dialkylpentafluorosuIfanylamines Et2NSF5(5a) and pip-SF5 (5b) are isolated from the reaction of the appropriate sulfurdifluoronitridearnides NSF2NR2 and HF/SF4. Protonation of the amines with the superacidic system HF/AsF5 gives stable pentafluorosulfanyl-ammonium salts SF5NHRR′. AsF6 (12: R = R′ = CH3; 14: R=R′=H; 10: R=CH3, R′ = H). Under the same conditions the adduct AsF5· NSF2CF(CF3)2 (15) forms a cation with hexacoordinated sulfur (trans-H3NSF4CF(CF3)2?AsF66: 16), while with Asp5 · NSF2NMe2 (17) the reaction stops at tetracoordination (HNSF2NMe2+AsF6 : 18).  相似文献   

16.
7‐Oxabenzonorbornadienes derivatives 1 a – d underwent reductive coupling with alkyl propiolates CH3C?CCO2CH3 ( 2 a ), PhC?CCO2Et ( 2 b ), CH3(CH2)3C?CCO2CH3 ( 2 c ), CH3(CH2)4C?CCO2CH3 ( 2 d ), TMSC?CCO2Et ( 2 e ), (CH3)3C?CCO2CH3 ( 2 f ) and HC?CCO2Et ( 2 g ) in the presence of [NiBr2(dppe)] (dppe=Ph2PCH2CH2PPh2), H2O and zinc powder in acetonitrile at room temperature to afford the corresponding 2alkenyl‐1,2‐dihydronapthalen‐1‐ol derivatives 3 a – n with remarkable regio‐ and diastereoselectivity in good to excellent yields. Similarly, the reaction of 7azabenzonorbornadienes derivative 1 e with propiolates 2 a, b and d proceeded smoothly to afford reductive coupling products 2alkenyl‐1,2‐dihydronapthalene carbamates 3 o – p in good yields with high regio‐ and stereoselectivity. This nickel‐catalyzed reductive coupling can be further extended to the reaction of 7oxabenzonorbornene derivatives. Thus, 5,6‐di(methoxymethyl)‐7‐oxabicyclo[2.2.1]hept‐2‐ene ( 4 ) reacted with 2 a and 2 d to furnish cyclohexenol derivatives bearing four cis substituents 5 a and b in 81 and 84 % yield, respectively. In contrast to the results of 4 with 2 , the reaction of dimethyl 7oxabicyclo[2.2.1]hept‐5‐ene‐2,3‐dicarboxylate ( 6 ) with propiolates 2 a – d afforded the corresponding reductive coupling/cyclization products, bicyclo[3.2.1]γ‐lactones 7 a – d in good yields. The reaction provides a convenient one‐pot synthesis of γ‐lactones with remarkably high regio‐ and stereoselectivity.  相似文献   

17.
A new series of (4′-hydroxy-3′,5′-dinitrophenyl) (3-aryloxiran-2-yl) methanone derivatives has been synthesized by the reaction of 4′-hydroxy-3′,5′-dinitro-substituted chalcones and alkaline H2O2. The resulted oxiranes on sulfanilic acid-catalyzed aminolysis afforded 2-hydroxy-1-(4′-hydroxy-3′,5′-dinitrophenyl)-3-aryl-3-(arylamino) propan-1-one derivatives. The advantage of this environmentally benign safe protocol offers a simple reaction set-up, mild reaction conditions, high product yields and short reaction time. The catalyst was reused several times without significant loss of catalytic activity.  相似文献   

18.
Two highly fluorinated bipyridine derivatives, (4,4′‐bis(RfCH2OCH2)‐2,2′‐bpy) {Rf = n‐C10F21 ( 1a ), n‐C10F23 ( 1b )}, have been synthesized starting from 4,4′‐bis(BrCH2)‐2,2′‐bpy and the corresponding fluorinated alkoxides. The fluorine contents of ligands 1a‐b are 62.3% and 63.3%, respectively, both being white solids, virtually insoluble in CH2Cl2 or DMF and highly fluorophilic with a partition ratio between DMF and n‐C8F18 less than 1:1000. The reaction of ligands 1a‐b with [Pd(CH3CN)2Cl2] results in novel Pd complexes [PdCl2(4,4′‐bis‐(RfCH2OCH2)‐2,2′‐bpy)] where Rf = n‐C10F21 ( 2a ), n‐C10F23 ( 2b ), respectively. The Pd complexes 2a‐b are pale yellow solids, soluble only in fluorinated solvents. The Pd complexes 2a‐b have been satisfactorily tested for Mizoroki‐Heck arylation under fluorous biphasic catalysis conditions in that the Pd complexes 2a‐b are easily recovered and maintain good catalytic activity after 8 consecutive cycles (> 90% yield). The TGA studies indicate that the Pd complexes 2a‐b are thermally stable up to 300 °C.  相似文献   

19.
Three copper(II) complexes, [Cu2(OAc)4L2] · 2CH3OH ( 1 ), [CuBr2L′2(CH3OH)] · CH3OH ( 2a ), and [CuBr2L′2(DMSO)] · 0.5CH3OH ( 2b ) {L = N‐(9‐anthracenyl)‐N′‐(3‐pyridyl)urea and L′ = N‐[10‐(10‐methoxy‐anthronyl)]‐N′‐(3‐pyridyl)urea} have been synthesized by the reaction of L with the corresponding copper(II) salts. Complex 1 shows a dinuclear structure with a conventional “paddlewheel” motif, in which four acetate units bridge the two CuII ions. In complexes 2a and 2b , the anthracenyl ligand L has been converted to an anthronyl derivative L′, and the central metal ion exhibits a distorted square pyramidal arrangement, with two pyridyl nitrogen atoms and two bromide ions defining the basal plane and the apical position is occupied by a solvent molecule (CH3OH in 2a and DMSO in 2b ).  相似文献   

20.
By allowing dimethyl peroxide (10?4M) to decompose in the presence of nitric oxide (4.5 × 10?5M), nitrogen dioxide (6.5 × 10?5M) and carbon tetrafluoride (500 Torr), it has been shown that the ratio k2/k2′ = 2.03 ± 0.47: CH3O + NO → CH3ONO (reaction 2) and CH3O + NO2 → CH3ONO2 (reaction 2′). Deviations from this value in this and previous work is ascribed to the pressure dependence of both these reactions and heterogeneity in reaction (2). In contrast no heterogeneous effects were found for reaction (2′) making it an ideal reference reaction for studying other reactions of the methoxy radical. We conclude that the ratio k2/k2′ is independent of temperature and from k1 = 1010.2±0.4M?1 sec?1 we calculate that k2′ = 109.9±0.4M?1 sec?1. Both k2 and k2′ are pressure dependent but have reached their limiting high-pressure values in the presence of 500 Torr of carbon tetrafluoride. Preliminary results show that k4 = 10.9.0±0.6 10?4.5±1.1M?1 sec?1 (Θ = 2.303RT kcal mole?1) and by k4 = 108.6±0.6 10?2.4±1.1M?1 sec?1: CH3O + O2 → CH2O + HO2 (reaction 4) and CH3O + t-BuH → CH3OH + (t-Bu) (reaction 4′).  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