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 15 毫秒
1.
The desulfurization of several N,2-diaryl-5-(arylimino)-2,5-dihydro-4-nitroisothiazol-3-amines 5 with Ph3P led to complex mixtures of products in low yields. For instance, quinoxaline-2-carboxamide 1-oxides of type 6 (Scheme 2) and, in some cases, also 3-nitroquinolines of type 7 (Scheme 5) were isolated. By the desulfurization of the substituted derivatives 5b – e , a rearrangement of the intermediates yielded 6 and 7 with a different substitution pattern from that expected from the starting materials (Scheme 3). The additional formation of two isomeric 1,2,5-oxadiazole-3-carboxamides 8 was observed only in the case of 5d (R1=R2=F) (Scheme 6). Under the same reaction conditions, the major product of the desulfurization of 5c was the quinoxaline-2-carboxamide 1-oxide 9 (Scheme 7). Reaction mechanisms involving intermediate ketene imines and O transfer from the NO2 group to the neighboring ketene imine are proposed. The structures of 6a , 6e , 6k , 7b , and 8d were established by X-ray crystallography, while the structure of 9 was elucidated by 2D-NMR spectroscopy and corroborated by X-ray crystallography.  相似文献   

2.
α-Chamigren-3-one (+) -8 bearing an axial CI-atom at C(8) exists as a largely dominant conformer with Me—C(5) at the envelope-shaped enone ring pointing away from CIax?C(8) at the cyclohexane ring (= B) in the ‘normal’ chair conformation, as shown by 1H-NMR. In contrast, the α-chamigren-3-ols (+) -9 and (+) -10 , obtained from hydride reduction of (+) -8 , show a temperature-dependent equilibrium of conformers where the major conformers have ring B in the inverted-chair (and twist-boat for (+) -9 ) conformation to avoid repulsions between Me?C(5) and CIax–C(8) (Scheme 1). This is in agreement with the conformation of the epoxidation product (+) -12 of (+) -9 where Me–C(5) is pushed away from CIax–C(8) in a ring-B chair similar to that of (+) -8 (Scheme 2). Introduction of a pseudoequatorial Br-atom at C(2) of (+) -8 , as in enone (+) -15 (Scheme 3), does not affect the conformation; but a pseudoaxial Br? C(2) experiences repulsive interactions with Heq–C(7), as shown by the 1H-NMR data of the isomeric enone (+) -16 where the ‘normal’-chair conformer Cβ -16 is in an equilibrium with the inverted chair conformer ICβ -16 (Scheme 3). These results and the accompanying paper allow a unifying view on the conformational behavior of marine polyhalogenated α-chamigrenes. This view is supported by the acid-induced isomerization of α-chamigrene (+) -9 (inverted chair) to β-chamigrene (+) -17 (‘normal’ chair; Scheme 4), the driving force being the lesser space requirement of CH2?C(5) than of Me–C(5). This explains why β-chamigrenes are so common in nature.  相似文献   

3.
Enantiomeric oligoribonucleotides (= ent-RNA) up to a sequence length of thirty-five and consisting of the (L -configurated) nucleosides ent-adenosine, ent-guanosine, ent-cytidine, ent-uridine, and 1-(β-L -ribofuranosyl)thymine were prepared by automated synthesis from appropriate building blocks, carrying a known photo-labile 2′-O-protecting group. A simple large-scale synthesis of the new, prefunctionalized L -ribose derivative 5 from D -glucose (Scheme 1) and its straightforward conversion into the five phosphoramidites 28 – 32 and five solid supports 38 – 42 , respectively, were elaborated (Scheme 4). Within this project, a novel, superior strategy for the synthesis of the 2′-O-{[(2-nitrobenzyl)oxy]methyl}-substituted key intermediates 18 – 22 by regioselective alkylation of their 5′-O-dimethoxytritylated precursors 13 – 17 was developed. Furthermore, an improved set-up for the final light-induced cleavage of the 2′-O-protecting groups from the oligonucleotide sequences was designed (Scheme 5 and Fig. 1). The correct composition of all ent-oligoribonucleotides prepared was established by their MALDI-TOF mass spectra. The 1H-NMR-spectroscopic data of a dodecameric ent-RNA sequence was in excellent agreement with the published data of its natural counterpart, synthesized by conventional methods. The known specific cleavage of a tetradecamer sequence by a 35mer ribozyme structure could be reproduced by ent-oligoribonucleotides, synthesized by the presented methods (Fig. 4).  相似文献   

4.
The Photochemistry of Conjugated Epoxy-Inones: Photolysis of 5,6-Epoxy-5-isopropyl-6-methyl-hept-3-in-2-on This paper continues the series of investigations of the photochemistry of α,β-unsaturated γ,δ-epoxyketones by examining the hitherto unknown photochemical behaviour of α,β-acetylenic-γ,δ-epoxy-ketones. As model compound, the aliphatic epoxy-ynone 7 (thermally stable at 180°) was synthesized (Scheme 1). It can be converted with BF3O (C2H5)2 in good yields to the 1,5-diketone 8 , the yne-1,4-diketone 49 and in small amounts to the fluorhydrine 50 (Scheme 1). On n,π*- or π, π*-excitation, 7 shows mainly cleavage of the C (γ)-O-bond to give a diradical a (Scheme 11), whose ultimate fate is strongly solvent dependent. In acetonitrile a mainly rearranges to the 1,5-diketone 8 and, to a smaller extent, shows fragmentation to acetone and formation of polymers. Except for small amounts of the dimeric products 9A,9B and biphenyl, the same compounds are obtained in benzene. In cyclopentane, however, a gives only little of 8 , and mainly a plethora of compounds formed by a radical process like H-abstraction from solvent, incorporation of cyclopentylradicals, dimerization and fragmentation reactions (9A, 9B, 11–20) (Scheme 3). Irradiation of 7 in propan-2-ol or in dioxane yields products of analogous radical processes as well of photoreduction (Scheme 4). However, the analogous epoxyenone 32 gives mainly products of photoisomerizations without interference by the solvent [6]. On photochemical excitation in acetonitrile, the 1,5-diketone 8 shows unspecific decomposition, but in cyclopentane it yields the reduction products 12, 26A, 26B, 27, 28 plus cyclopentylcyclopentane (15) (Scheme 6).  相似文献   

5.
Two new chiral bidentate (phosphinophenyl)benzoxazine P,N-ligands 2a and 2b were synthesized from highly enantiomer-enriched 2-(1-aminoalkyl)phenols 4 . Ligand rac- 2a was obtained on refluxing the t-Bu-substituted (aminomethyl)phenol 4a with 2-(diphenylphosphino)benzonitrile in chlorobenzene in the presence of anhydrous ZnCl2 followed by decomplexation (Scheme 2). This reaction, when carried out with (+)-(S)- 4a , was accompanied by racemization at the stereogenic center of the alkyl side chain. The enantiomerically pure ligands (+)-(R)- 2a and (−)-(S)- 2a were obtained using a stepwise procedure via the amides (−)-(R)- and (+)-(S)- 5b , respectively, followed by cyclization to benzoxazines (+)-(R)- and (−)-(S)- 7b , respectively, with triflic anhydride and by F-atom substitution by diphenylphosphide (Schemes 3 and 5). In the case of the i-Pr analogue 2b , this last step resulted in racemization (Scheme 6). This was overcome by preparing the bromo derivative and introducing the diphenylphosphine group via Br/Li exchange and reaction with chlorodiphenylphosphine (Scheme 7). The first application of (+)-(R)- 2a in an asymmetric Heck reaction showed high enantioselectivity (91%) (Scheme 8).  相似文献   

6.
《Tetrahedron: Asymmetry》1999,10(12):2399-2410
The first synthesis of enantiomerically pure 2-azabicyclo[3.3.1]nonanes by an intramolecular radical reaction of the trichloroacetamido group bearing an (S)-N-1-phenylethyl substituent with the silyl enol ether moiety in compounds 7 is described. The procedure allows the two enantiomers of the 2-azabicyclo[3.3.1]nonane-3,6-dione, 3 and ent-3, to be prepared separately. β-Lactam 8 and normorphan 9 are also formed from 7 through an initial radical translocation process in the cyclization step.  相似文献   

7.
Three spermidine alkaloids – oncinotine ( 1 ), neooncinotine ( 3 ), and isooncinotine ( 2 ) – have been isolated from the stem bark of Oncinotis nitida BENTH . (Scheme 1); 1 and 3 are so far an unseparable mixture. However, by treatment of this mixture with K-t-butoxide, neooncinotine is completely converted into isooncinotine, and oncinotine, the main alkaloid, is obtained in pure form. The structural assigment of these alkaloids is based on chemical and spectral evidence. Thus oncinotine ( 1 ) has been degraded via 24 (Scheme 4) and 32 to the putrescine derivative 35 and the piperidine derivative 34 (Scheme 5). Similarly neooncinotine ( 3 ) and isooncinotine ( 2 ), have given 34 along with the 1, 3-diaminopropane derivative 36 (Scheme 5). The major decomposition pathways of 24 , 35 and 36 in the mass spectra are described in Schemes 8, 6 and 7 respectively. The absolute configuration of 1 , 2 and 3 is derived by chiroptical correlations with (R)-(?)-N-methylconiine ( 38 ).  相似文献   

8.
The theoretical structure of a cyclic phosphoric triamide 3 and of its monolithiated isomers 4 – 6 was calculated by ab initio methods (Fig. 1, Tables 1 and 2). The global minimum in 4 consists of a five-membered Li−C−N−P−O chelate. The intermediates 5 and 6 are, relative to 4 , energetically unfavorable by 15 and 18 kcal mol−1, respectively, due to distortion in order to accommodate the N-complexation of the Li+ ions. NMR Investigations (1H, 13C, 31P, and 7Li) of the lithiated bicyclic phosphoric triamide 1 were performed (Tables 3 – 5). The lithium aminomethanide 2 is characterized by a sp3-hybridized anion supporting Li−C contacts. The anions exist in a monomer-dimer equilibrium in solution (Scheme 2). Trapping reactions of rac- 2 with carbonyl compounds generated the corresponding amino-alcohol derivatives with high diastereoselectivities (Scheme 3, Table 6). A rational for the stereochemical outcome is given (Fig. 4). In the presence of LiBr, a P−N bond cleavage occurred on reaction of rac- 2 with aldehydes, which allowed the synthesis of (1-hydroxylalkyl)phosphonic diamides (Scheme 5, Table 7).  相似文献   

9.
Rearrangements of (2′-Propinyl)cyclohexadienols and -semibenzenes The acid-catalyzed dienol-benzene rearrangement of 3- and 5-methyl-substituted (2′-propinyl)cyclohexadienols has been investigated. Treatment of the dienols with CF3COOH in CCl4 yields allenyl- and (2′-propinyl)benzenes via [3,4]- and [1,2]-sigmatropic rearrangements, respectively. The reaction with H2SO4 in Et2O leeds to a mixture of allenyl-, 2′-propinyl-, 3′-butinyl- and (2′,3′-butadienyl)benzenes (Scheme 3). The latter are products of a thermal semibenzene-benzene rearrangement (cf. Scheme 9). The corresponding semibenzenes have been prepared by dehydration of the cyclohexadienols with H2SO4 or POCl3 (Schemes 6 and 7). Under acidic conditions, the p-(2′-propinyl)semibenzenes 33–35 (Scheme 8) undergo [3,4]- and [1,2]-sigmatropic rearrangements to give again allenyl- and (2′-propinyl)benzenes, whereas the thermal rearrangements to the 3′-butinyl- and (2′,3′-butadienyl)benzenes (Scheme 9) involves a radical mechanism. In contrast, the o-(2′-propinyl)semibenzene b (Scheme 7) leads to (2′,3′-butadienyl)benzene 32 via a thermal [3,3]-sigmatropic rearrangement.  相似文献   

10.
The 3‐methyl‐4‐(tricyclo[5.2.1.02,6]dec‐4‐en‐8‐ylidene)butan‐2‐ols (=Fleursandol®; rac‐ 10 ), a new class of sandalwood odorants, were synthesized in their enantiomerically pure forms by use of tricyclo[5.2.1.02,6]dec‐4‐en‐8‐ones 17 and ent‐ 17 and (tetrahydro‐2H‐pyran‐2‐yl)‐protected 4‐bromo‐3‐methylbutan‐2‐ols 22 and ent‐ 22 as starting materials (Schemes 2–4). Only four of 16 possible stereoisomers of rac‐ 10 possess the typical, very pleasant, long‐lasting sandalwood odor (Table 1). The (2S,3R,4E,1′R,2′R,6′R,7′R)‐isomer ent‐ 10a is by far the most important representative, with an odor threshold of 5 μg/l in H2O.  相似文献   

11.
Earlier phytochemical work on Plectranthus ambiguus (Lamiaceae) afforded a series of tetracyclic phyllocladane‐type (=13β‐kaurane) diterpenoids (see 1a – f ). In the course of investigations concerning the reaction behavior of this rare natural‐products, a new constituent of P. ambiguus was isolated, (2S,3R,16R)‐phyllocladane‐2,3,16,17‐tetrol 2,3‐diacetate ( 1g ), and another eighteen new phyllocladanes were prepared by chemical transformations and characterized. The main constituent 1b of P. ambiguus was chemically transformed to the known natural diterpenoid calliterpenone (=(16R)‐16,17‐dihydroxyphyllocladan‐3‐one; 2 ) thus unambiguously establishing its structure (Scheme 1). Epimerization at C(16) via the epoxy derivative 20 yielded 16‐epicalliterpenone ( 21 ), 17‐hydroxyphylloclad‐15‐ene‐3‐one ( 22 ), and (16R)‐3‐oxophyllocladan‐17‐al ( 23 ) (Scheme 6). Comparing this reaction sequence with the corresponding one starting from diastereoisomeric (16R)‐16,17‐dihydroxy‐ent‐kauran‐3‐one (=abbeokutone; 27 ) showed basically the same outcome (Scheme 7). Furthermore, three new C(16)‐substituted ent‐kauran‐3‐ones were characterized. Reliable spectroscopic arguments for the determination of the configuration at C(16) in phyllocladanes and kauranes as well as for the differentiation of the diastereoisomeric skeletons are presented.  相似文献   

12.
Reaction of 3-(Dimethylamino)-2H-azirines with 1,3-Thiazolidine-2-thione Reaction of 3-(dimethylamino)-2H-azirines 1 and 1,3-thiazolidine-2-thione ( 6 ) in MeCN at room temperature leads to a mixture of perhydroimidazo[4,3-b]thiazole-5-thiones 7 and N-[1-(4,5-dihydro-1,3-thiazol-2-yl)alkyl]-N′,N′-dimethylthioureas 8 (Scheme 2), whereas, in i-PrOH at ca. 60°, 8 is the only product (Scheme 4). It has been shown that, in polar solvents or under Me2NH catalysis, the primarily formed 7 isomerizes to 8 (Scheme 4). The hydrolysis of 7 and 8 leads to the same 2-thiohydantoine 9 (Scheme 3 and 5). The structure of 7a, 8c , and 9b has been established by X-ray crystallography (Chapt. 4). Reaction mechanisms for the formation and the hydrolysis of 7 and 8 are suggested.  相似文献   

13.
Photocyclization of 1, 1′-Polymethylene-di-2-pyridones . Benzophenone sensitized irradiation of the four dipyridones 1-4 gave the internal photocyclization products 6 (64%, Scheme 4), 7 (60%, Scheme 5), 8 (Scheme 6), and 11 (26%, Scheme 7), respectively. The decamethylene compound 5 yielded only polymeric material. The primary [2+2] photoproduct 8 from dipyridone 3 (Scheme 6) is relatively unstable. Further irradiation or heating to 65° induced a Cope rearrangement to give compound 9 which, on heating to 137°, was converted into the isomeric compound 10 . This product, as well as the other photoproducts mentioned, are rearranged back to their respective starting materials upon direct irradiation with 254 nm light or by heating to higher temperatures. The various possibilities for cycloadditions of pyridones are discussed as well as the possible factors which are responsible for the highly regioselective photoreactions of the dipyridones 1–4 .  相似文献   

14.
The Stereoselectivity of the α-Alkylation of (+)-(1R, 2S)-cis-Ethyl-2-hydroxy-cyclohexanecarboxylate In continuation of our work on the stereoselectivity of the α-alkylation of β-hydroxyesters [1] [2], we studied this reaction with the title compound (+)- 2 . The latter was prepared through reduction of 1 with baker's yeast. Alkylation of the dianion of (+)- 2 furnished (?)- 4 in 72% chemical yield (Scheme 1) and with a stereoselectivity of 95%. Analogously, (?)- 7 was prepared with similar yields. Oxidation of (?)- 4 and (?)- 7 respectively furnished the ketones (?)- 6 (Scheme 3) and (?)- 8 (Scheme 4) respectively, each with about 76% enantiomeric excess (NMR.). It is noteworthy that yeast reduction of rac- 6 (Scheme 3) is completely enantioselective with respect to substrate and product and gives optically pure (?)- 4 in 10% yield, which was converted into optically pure (?)- 6 (Scheme 3). The alkylation of the dianionic intermediate shows a higher stereoselectivity (95%) from the pseudoequatorial side than that of 1-acetyl- or 1-cyano-4-t-butyl-cyclohexane (71% and 85%) [9] or that of ethyl 2-methyl-cyclohexanecarboxylate (82%). The stereochemical outcome of the above alkylation is comparable with that found in open chain examples [1] [2]. Finally (+)-(1R, 2S)- 2 was also alkylated with Wichterle's reagent to give (?)-(1S, 2S)- 9 in 64% yield. The latter was transformed into (?)-(S)- 10 and further into (?)-(S)- 11 (Scheme 5). (?)-(S)- 10 and (?)-(S)- 11 showed an e.e. of 76–78% (see also [11]). Comparison of these results with those in [11] confirmed our former stereochemical assignment concerning the alkylation step.  相似文献   

15.
Reaction of Ethyl Diazoacetate with 1,3-Thiazole-5(4H)-thiones Reaction of ethyl diazoacetate ( 2a ) and 1,3-thiazole-5(4H)-thiones 1a,b in Et2O at room temperature leads to a complex mixture of the products 5–9 (Scheme 2). Without solvent, 1a and 2a react to give 10a in addition to 5a–9a . In Et2O in the presence of aniline, reaction of 1a,b with 2a affords the ethyl 1,3,4-thiadiazole-2-carboxylate 10a and 10b , respectively, as major products. The structures of the unexpected products 6a, 7a , and 10a have been established by X-ray crystallography. Ethyl 4H-1,3-thiazine-carboxylate 8b was transformed into ethyl 7H-thieno[2,3-e][1,3]thiazine-carboxylate 11 (Scheme 3) by treatment with aqueous NaOH or during chromatography. The structure of the latter has also been established by X-ray crystallography. In the presence of thiols and alcohols, the reaction of 1a and 2a yields mainly adducts of type 12 (Scheme 4), compounds 5a,7a , and 9a being by-products (Table 1). Reaction mechanisms for the formation of the isolated products are delineated in Schemes 4–7: the primary cycloadduct 3 of the diazo compound and the C?S bond of 1 undergoes a base-catalyzed ring opening of the 1,3-thiazole-ring to give 10 . In the absence of a base, elimination of N2 yields the thiocarbonyl ylide A ′, which is trapped by nucleophiles to give 12 . Trapping of A ′, by H2O yields 1,3-thiazole-5(4H)-one 9 and ethyl mercaptoacetate, which is also a trapping agent for A ′, yielding the diester 7 . The formation of products 6 and 8 can be explained again via trapping of thiocarbonyl ylide A ′, either by thiirane C (Scheme 6) or by 2a (Scheme 7). The latter adduct F yields 8 via a Demjanoff-Tiffeneau-type ring expansion of a 1,3-thiazole to give the 1,3-thiazine.  相似文献   

16.
The reaction of 3-(dimethylamino)-2H-azirines 1a–c and 2-amino-4,6-dinitrophenol (picramic acid, 2 ) in MeCN at 0° to room temperature leads to a mixture of the corresponding 1,2,3,4-tetrahydroquinazoline-2-one 5 , 3-(dimethylamino)-1,2-dihydroquinazoline 6 , 2-(1-aminoalkyl)-1,3-benzoxazole 7 , and N-[2-(dimethylamino)phenyl]-α-aminocarboxamide 8 (Scheme 3). Under the same conditions, 3-(N-methyl-N-phenyl-amino)-2H-azirines 1d and 1e react with 2 to give exclusively the 1,3-benzoxazole derivative 7 . The structure of the products has been established by X-ray crystallography. Two different reaction mechanisms for the formation of 7 are discussed in Scheme 6. Treatment of 7 with phenyl isocyanate, 4-nitrobenzoyl chloride, tosyl chloride, and HCl leads to a derivatization of the NH2-group of 7 (Scheme 4). With NaOH or NaOMe as well as with morpholine, 7 is transformed into quinazoline derivatives 5 , 14 , and 15 , respectively, via ring expansion (Scheme 5). In case of the reaction with morpholine, a second product 16 , corresponding to structure 8 , is isolated. With these results, the reaction of 1 and 2 is interpreted as the primary formation of 7 , which, under the reaction conditions, reacts with Me2NH to yield the secondary products 5 , 6 , and 8 (Scheme 7).  相似文献   

17.
Aromatic Sigmatropic Hydrogen-Shifts in 2-Vinyl- and 2-Allyl-phenols It is shown by deuterium labeling experiments that 2-vinylphenols, on heating at 142,5°, undergo aromatic [1,5]-H-shifts whereby o-quinone methides are formed as intermediates (Scheme 7). Thus, heating of 2-isopropenylphenol ( 6 ) in a D2O/dioxane mixture leads to a rapid deuterium incorporation into the methylidene group of the isopropenyl moiety (Table 1) whereas its methyl group shows only a slow uptake of deuterium. The latter exchange process can be attributed to intermolecular reactions (Scheme 8). The quinone methide intermediates (e.g. 26 , Scheme 8) can be regarded as vinyl homologues of alkyl ketones. Therefore, 26 can exchange hydrogen in both methyl groups by an acid- and base-catalysed mechanism. Indeed, when 6 is heated in D2O/pyridine or D2O/CH3COOD/dioxane, an almost statistical incorporation of deuterium into the methylidene and the methyl group of the isopropenyl moiety is observed (Table 3). As a consequence of thermally induced [1,5]-H-shifts, 2-(1′-propenyl)-phenols undergo rapid (E,Z) isomerization with first order kinetics on heating above 140° in decane solution. Activation parameters are given in Table 4. The observed primary +++++ H/D isotope effect of 3.3 in the (E,Z) isomerization of phenol 8 is in +++ment with intramolecular H/D-shifts in the rate determing step (Scheme 9 +++ Table 5). As expected aromatic sigmatropic [1,5]-H-shifts in 2-(1′-propenyl)-+++ are much faster than aromatic homosigmatropic [1,5]-H-shifts in 2-(2′-+++++)phenols (Scheme 1 and Table 6). The structurally comparable phenols +++ (Z)- 10 and (E)/(Z)- 14 (Scheme 3) show k([1,5])/k(homo-[1,5]) ≈ 2300 at ++++
  • 1 A more detailed discussion in English is given in [1].
  • .  相似文献   

    18.
    Partial Syntheses and Reactions of Abietanoid Derivatives (Lanugones) from Plectranthus lanuginosus and of Related Compounds Interconversions by partial syntheses of several lanugones establish their absolute configuration at C(15). Unexpected reactions exemplify the unique reactivity of these abietanoic diterpenes, - Lanugone O ( 4 ) was prepared in several steps from (15S)-coleon C ( 8a ; Scheme 2) thus establishing its (15S)-configuration. One of the intermediates, the 12-O-acetyl-6-oxoroyleanone 12 , through acetyl-migration sets up an equilibrium with the vinylogous quinone 13 (Scheme 3). - The chirality at C(15) in the dihydrofuran moiety of lanugone Q ( 16 ) was proven by acid-catalyzed conversion of lanugone O ( 4 ) to 16 . - Instead of the usual nucleophilic attack shown by quinomethanes, lanugone L (1 ) is electrophilically substituted at C(7) by acetic anhydride/pyridine (Scheme 1). - In a homosigmatropic [1,5]-H-shift, lanugone G ( 17 ) in solution is converted to the corresponding allyl substituted royleanone 18 (Scheme 4). - Methanolysis of lanugone J ( 19 ) leads to the expected royleanone 20 having the 2-methoxypropyl side chain ( Scheme 5 ). Similar reactions were found in acetolytic reactions. However, treatment-of spirocoleons with SOCl2/DMF produces mainly 12-deoxyroyleanones with allyl- and 2-chloropropyl groups, i. e. 19 → 26 and 27 ; 28 → 29 . The possible natural occurrence of these compounds is emphasized.  相似文献   

    19.
    Deprotection of the tetramer 24 , obtained by coupling the iodinated dimer 18 with the alkyne 23 gave the 8′,5‐ethynediyl‐linked adenosine‐derived tetramer 27 (Scheme 3). As direct iodination of C(5′)‐ethynylated adenosine derivatives failed, we prepared 18 via the 8‐amino derivative 17 that was available by coupling the imine 15 with the iodide 7 ; 15 , in its turn, was obtained from the 8‐chloro derivative 12 via the 4‐methoxybenzylamine 14 (Scheme 2). This method for the introduction of the 8‐iodo substituent was worked out with the N‐benzoyladenosine 1 that was transformed into the azide 2 by lithiation and treatment with tosyl azide (Scheme 1). Reduction of 2 led to the amine 3 that was transformed into 7 . 1,3‐Dipolar cycloaddition of 3 and (trimethylsilyl)acetylene gave 6 . The 8‐substituted derivatives 4a – d were prepared similarly to 2 , but could not be transformed into 7 . The known chloride 8 was transformed into the iodide 11 via the amines 9 and 10 . The amines 3 , 10 , and 16 form more or less completely persistent intramolecular C(8)N−H⋅⋅⋅O(5′) H‐bonds, while the dimeric amine 17 forms a ca. 50% persistent H‐bond. There is no UV evidence for a base‐base interaction in the protected and deprotected dimers and tetramers.  相似文献   

    20.
    Abstract

    Diastereoisomeric 2-t-butylamino-2-seleno-4,4,6-trimethyl-1,3,2-dioxaphosphorinans have been prepared and separated into pure cis- and trans-isomers. The minor isomer (cis) was investigated by means of x-ray crystallography and its has been demonstrated, that in solid state is possesses twist-boat conformation.  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