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 preparation of novel electrophilic building blocks for the synthesis of enantiomerically pure compounds (EPC) is described. Thus, the 2-(tert-butyl)dioxolanones, -oxazolidinones, -imidazolidinones, and -dioxanones obtained by acetalization of pivalaldehyde with 2-hydroxy-, 3-hydroxy-, or 2-amino-carboxylic acids are treated with N-bromosuccinimide under typical radical-chain reaction conditions (azoisobuytyronitril/CCl4/reflux). Products of bromination in the α-position of the carbonyl group of the five-membered-ring acetals are isolated or identified ( 2, 5 , and 8 ; Scheme 1). The dioxanones are converted to 2H, 4H-dioxinones under these conditions ( 12 , 14 , 15 , 21 , and 22 ; Schemes 2 and 3). The products can be converted to chiral derivatives of pyruvic acid (methylidene derivatives 3 and 6 ) or of 3-oxo-butanoic and -pentanoic acid ( 16 and 23 ). The mechanism of the brominations is interpreted. The conversion of serine to enactiomcrically pure dioxanones 26–28 (Scheme 4) is also discussed.  相似文献   

2.
Using the readily accessible chiral auxiliaries 1 – 3 the sulfonamide-shielded O-silylated esters 5 underwent π-face-selective α-acetoxylation on successive treatment with Pb(OAc)4 and NEt3 HF to give after recrystallization α-acetoxy ester 6 in 55–67% yields and in 95–100% d.e. Starting from conjugated enoates addition of RCu and subsequent acetoxylation 10 → 11 → 12 yielded α,β-bifunctionalized esters 12 with >95% configurational control at both Cα and Cβ. Nondestructive removal of the auxiliary ( 6 → 7 , 6 → 8 and 12 → 13 ) gave either α-hydroxycarboxylic acids or terminal α-glycols in high enantiomeric purity. The prepared glycols 8c and 13a are key intermediates for previously reported syntheses of the natural products 16 and 17 , respectively.  相似文献   

3.
The bicyclic ketones 1–6 entered into diastereoselective (> 95% d.e.) aldol reactions with a variety of aldehydes (Scheme 1 and Table 1). A representative series of aldols was converted (E)-selectively into α,β-unsaturated ketones by (i) spontaneous base-promoted dehydration (Scheme 1 and Table 2) and also by (ii) conversion into brosylate and base-mediated elimination with lithium diisopropylamide/N,N,N′,N′-tetramethylethylenediamine (LDA/TMEDA; Scheme 2). The simple α-methylidene ketones 17a and 18a were obtained via oxidation of the phenylselenides 19 and 20 , respectively (Scheme 4). The tertiary aldol 27 was synthesized best by treatment of 1,3-diketone 26 with Me4Zr (Table 4). In this fashion, the facile retro-aldol reaction of 27 was suppressed effectively.  相似文献   

4.
Glycosylidene carbenes derived from the GlcNAc and AllNAc diazirines 1 and 3 were generated by the thermolysis or photolysis of the diazirines. The reaction of 1 with i-PrOH gave exclusively the isopropyl α-D -glycoside of 5 besides some dihydrooxazole 9 (Scheme 2). A similar reaction with (CF3)2CHOH yielded predominantly the α-D -anomer of 6 , while glycosidation of 4-nitrophenol (→ 7 ) proceeded with markedly lower diastereoselectivity. Similarly, the Allo-diazirine 3 gave the corresponding glycosides 12–14 , but with a lower preference for the α-D -anomers (Scheme 3). The reactions of the carbene derived from 1 with Ph3COH (→ 8 ) and diisopropylideneglucose 10 (→ 11 ) gave selectively the α-D -anomers (Scheme 2). The αD -selectivity increases with increasing basicity (decreasing acidity) of the alcohols. It is rationalized by an intermolecular H-bond between the acetamido group and the glycosyl acceptor. This H-bond increases the probability for the formation of a 1,2-cis-glycosidic C–O bond. The gluco-intermediates are more prone to forming a N–H…?(H)OR bond than the allo-isomers, since the acetamido group in the N-acetylallosamine derivatives forms an intramolecular H-bond to the cis-oriented benzyloxy group at C(3), as evidenced by δ/T and δ/c experiments.  相似文献   

5.
The reaction of 1,4,5‐trisubstituted 1H‐imidazole 3‐oxides 1 with Ac2O in CH2Cl2 at 0 – 5° leads to the corresponding 1,3‐dihydro‐2H‐imidazol‐2‐ones 4 in good yields. In refluxing Ac2O, the N‐oxides 1 are transformed to N‐acetylated 1,3‐dihydro‐2H‐imidazol‐2‐ones 5 . The proposed mechanisms for these reactions are analogous to those for N‐oxides of 6‐membered heterocycles (Scheme 2). A smooth synthesis of 1H‐imidazole‐2‐carbonitriles 2 starting with 1 is achieved by treatment with trimethylsilanecarbonitrile (Me3SiCN) in CH2Cl2 at 0 – 5° (Scheme 3).  相似文献   

6.
The reactions of α-diazo ketones 1a,b with 9H-fluorene-9-thione ( 2f ) in THF at room temperature yielded the symmetrical 1,3-dithiolanes 7a,b , whereas 1b and 2,2,4,4-tetramethylcyclobutane-1,3-dithione ( 2d ) in THF at 60° led to a mixture of two stereoisomeric 1,3-oxathiole derivatives cis- and trans- 9a (Scheme 2). With 2-diazo-1,2-diphenylethanone ( 1c ), thio ketones 2a–d as well as 1,3-thiazole-5(4H)-thione 2g reacted to give 1,3-oxathiole derivatives exclusively (Schemes 3 and 4). As the reactions with 1c were more sluggish than those with 1a,b , they were catalyzed either by the addition of LiClO4 or by Rh2(OAc)4. In the case of 2d in THF/LiClO4 at room temperature, a mixture of the monoadduct 4d and the stereoisomeric bis-adducts cis- and trans- 9b was formed. Monoadduct 4d could be transformed to cis- and trans- 9b by treatment with 1c in the presence of Rh2(OAc)4 (Scheme 4). Xanthione ( 2e ) and 1c in THF at room temperature reacted only when catalyzed with Rh2(OAc)4, and, in contrast to the previous reactions, the benzoyl-substituted thiirane derivative 5a was the sole product (Scheme 4). Both types of reaction were observed with α-diazo amides 1d,e (Schemes 5–7). It is worth mentioning that formation of 1,3-oxathiole or thiirane is not only dependent on the type of the carbonyl compound 2 but also on the α-diazo amide. In the case of 1d and thioxocyclobutanone 2c in THF at room temperature, the primary cycloadduct 12 was the main product. Heating the mixture to 60°, 1,3-oxathiole 10d as well as the spirocyclic thiirane-carboxamide 11b were formed. Thiirane-carboxamides 11d–g were desulfurized with (Me2N)3P in THF at 60°, yielding the corresponding acrylamide derivatives (Scheme 7). All reactions are rationalized by a mechanism via initial formation of acyl-substituted thiocarbonyl ylides which undergo either a 1,5-dipolar electrocyclization to give 1,3-oxathiole derivatives or a 1,3-dipolar electrocyclization to yield thiiranes. Only in the case of the most reactive 9H-fluorene-9-thione ( 2f ) is the thiocarbonyl ylide trapped by a second molecule of 2f to give 1,3-dithiolane derivatives by a 1,3-dipolar cycloaddition.  相似文献   

7.
Formation of cyclic ions and bicyclic transition states in the mass spectral decomposition of substituted α,ω-alkanediamines. N-Phenethyl-N(4-acetamidobutyl)-p-toluene-sulfonamide ( 4 ) and its homologues were synthesized and the mass spectral behaviour investigated. After loss of a benzyl radical from the molecular ion two different fragmentation reactions are observed. The lower homologous members – namely compounds 1 , 2 and 3 – lose ketene by formation of cyclic ions (Scheme 1). The higher homologues of this series of compounds ( 4 , 5 , 6 ) show a pronounced (to 18% ∑50) loss of p-toluene sulfonic acid. This decomposition reaction proceeds presumably through a bicyclic transition state (Scheme 3).  相似文献   

8.
It is proposed to study the influence of interresidue H-bonds on the structure and properties of polysaccharides by comparing them to a series of systematically modified oligosaccharide analogues where some or all of the glycosidic O-atoms are replaced by buta-1,3-diyne-1,4-diyl groups. This group is long enough to interrupt the interresidue H-bonds, is chemically versatile, and allows a binomial synthesis. Several approaches to the simplest monomeric unit required to make analogues of cellulose are described. In the first approach, allyl α-D -galactopyranoside ( 1 ) was transformed via 2 and the tribenzyl ether 3 into the triflate 4 (Scheme 2). Substitution by cyanide (→ 5–7 ) followed by reduction with DIBAH led in high yield to the aldehyde 9 , which was transformed into the dibromoalkene 10 and the alkyne 11 following the Corey-Fuchs procedure (Scheme 3). The alkyne was deprotected via 12 or directly to the hemiacetal 13 . Oxidation to the lactone 14 , followed by addition of lithium (trimethylsilyl)acetylide Me3SiC?CLi/CeCl3 (→ 15 ) and reductive dehydroxylation afforded the disilylated dialkyne 16 . The large excess of Pd catalyst required for the transformation 11 → 13 was avoided by deallylating the dibromoalkene 10 (→ 17 → 18 ), followed by oxidation to the lactone 19 , addition of Me3SiC?CLi to the anomeric hemiketals 20 (α-D /β-D 7:2), dehydroxylation to 21 , and elimination to the monosilylated dialkyne 22 (Scheme 3). In an alternative approach, treatment of the epoxide 24 (from 23 ) with Me3SiC?CLi/Et2AlCl according to a known procedure gave not only the alkyne 27 but also 25 , resulting from participation of the MeOCH2O group (Scheme 4). Using Me3Al instead of Et2AlCl increased the yield and selectivity. Deprotection of 27 (→ 28 ), dibenzylation (→ 29 ), and acetolysis led to the diacetate 30 which was partially deacetylated (→ 31 ) and oxidized to the lactone 32 . Addition of Me3SiC?CLi/TiCl4 afforded the anomeric hemiketals 33 (α-D /β-D 3:2) which were deoxygenated to the dialkyne 34 . This synthesis of target monomers was shortened by treating the hydroxy acetal 36 (from 27 ) with (Me3SiC?C)3Al (Scheme 5): formation of the alkyne 37 (70%) by fully retentive alkynylating acetal cleavage is rationalised by postulating a participation of HOC(3). The sequence was further improved by substituting the MeOCH2O by the (i-Pr)3SiO group (Scheme 6); the epoxide 38 (from 23 ); yielded 85% of the alkyne 39 which was transformed, on the one hand, via 40 into the dibenzyl ether 29 , and, on the other hand, after C-desilylation (→ 41 ) into the dialkyne 42 . Finally, combined alkynylating opening of the oxirane and the 1,3-dioxolane rings of 38 with excess Et2Al C?CSiMe3 led directly to the monomer 43 which is thus available in two steps and 77% yield from 23 (Scheme 6).  相似文献   

9.
The conjugate addition of the chiral, non-racemic alkoxy-enolates 5 and 6 to nitroolefins furnishes the hydroxynitroesters 7–13 , which are catalytically hydrogenated to give the lactams 14–18 . The configuration of adduct 7 from nitroethylene was elucidated by NMR. analysis of the acetal 20 derived from 7 . The assignment establishes that the reaction follows the stereochemical rule of attack depicted in 21 and previously deduced for other electrophiles, i.e. formation of erythro-products of type 3b and 4b . No stereocontrol was found at the newly formed chiral centers in α- and β-position to the NO2 group of 8–12 .  相似文献   

10.
The crystal structures of 32 derivatives of 4‐isopropyl‐5,5‐diphenyl‐1,3‐oxazolidin‐2‐one ( A and 1 – 31 ) are presented (Fig. 2 and Tables 1–3). In all but four structures, the Me2CH group is in a disposition that mimick a Me3C group (Figs. 3–5). The five‐membered ring shows conformations from an envelope form with the Ph2C group out of the plane containing the other four atoms to the twist form with the twofold axis through the CO group (Fig. 6, and Table 2). In the entire series, the Me2CH and the neighboring trans Ph group are approximately antiperiplanar (average torsion angle 155°). The structural features are used to interpret the previously observed reactivity behavior of the diphenyl‐oxazolidinone derivatives. The practical advantages of the title compound over classical Evans auxiliaries are outlined (Figs. 1 and 7, and Scheme 2): high crystallinity of all derivatives, steric protection of the CO group in the ring, excellent stereoselectivities in reactions of its derivatives, and safe preparation and easy recovery of the auxiliary.  相似文献   

11.
Thermolysis of the 1,2-bis(glucosylalkynyl)benzenes 6 and 16 was studied to evaluate the effects of intramolecular H-bonding on the activation energy of the Bergman-Masamune-Sondheimer cycloaromatization, and to evaluate the use of the cycloaromatization for the synthesis of di-glycosylated naphthalenes. The dialkynes were prepared by cross-coupling of the O-benzylated or O-silylated glucosylalkynes 1 and 4 (Scheme 1). Thiolysis of the known 1 , or acetolysis of 1 , followed by deacetylation ( →2→3 ) and silylation gave 4 . Cross-coupling of 1 or 4 with iodo- or 1,2-diiodobezene depended upon the nature of the added amine and on the protecting group, and led to the mono- and dialkynylbenzenes 5 and 6 , or 12, 13 , and 15 , respectively. The benzyl ethers 5 and 6 gave poor yields upon acetolysis catalyzed by BF3 · OEt2, while Ac2O/CoCl2 · 6 H2O transformed 6 in good yields into the regioselectively debenzylated 10 . Desilylation of 7 and 13 gave the alcohols 8 and 14 , respectively. Thermolysis of 6 in PhCl gave 22 and 23 , independently of the presence or absence of 1,4-cyclohexadiene; 23 was formed from 22 (Scheme 2). Acetolysis of 22 gave the hexaacetate 24 that was completely debenzylated by thiolysis, yielding the diol 26 and trans-stilbene, evidencing the nature and position of the bridge between the glucosyl moieties (Scheme 3). Thiolysis of 22 yielded the unprotected 2,3-diglucosylnaphthalene 28 , a new type of C-glycosides. Depending upon conditions, hydrogenation of 22 led to 29 (after acetylation), 30 , or 32 . NMR and particularly NOE data evidence the threo-configuration of the bridge. The structure of 23 was confirmed by hydrolysis to the diol 34 and diphenylacetaldehyde, and by correlation of 23 with 22 via the common product 31 . Formation of 22 is rationalized by a Bergman cyclization to a diradical, followed by regioselective abstraction of a H-atom from the BnO? C(2) group, and diastereoselective combination of the doubly benzylic diradical (Scheme 4). While thermolysis of 3 in EtOH sets in around 140°, 16 did not react at 160° and decomposed at 180–220°. No evidence for intramolecular H-bonds of 16 , as compared to 14 , were found.  相似文献   

12.
Condensation of 3,4‐dimethoxybenzeneethanamine ( 3d ) and various benzeneacetic acids, i.e., 4a – e , via a practical and efficient one‐pot Bischler–Napieralski reaction, followed by NaBH4 reduction, produced a series of 1‐benzyl‐1,2,3,4‐tetrahydroisoquinolines, i.e., 5a – e , in satisfactory yields (Scheme 3). Oxidative coupling of the N‐acyl and N‐methyl derivatives 6a – e of the latter with hypervalent iodine ([IPh(CF3COO)2]) yielded products with two different skeletons (Scheme 4). The major products from N‐acyl derivatives 6a – c were (±)‐N‐acylneospirodienones 2a – c , while the minor was the 3,4‐dihydroisoquinoline 7 . (±)‐Glaucine ( 1 ), however, was the major product starting from N‐methyl derivative 6e . Possible reaction mechanisms for the formation of these two types of skeleton are proposed (Scheme 5).  相似文献   

13.
Treatment of the (butadiene)ML2 complexes 1 [ML2 = Cp2Zr ( a ), Cp2Hf ( b ), and (.-C5H4CH3)2Zr ( c )] with B(C6F5)3 gives the 1:1 addition products (CH2CHCHCH2–B(C6F5_3)ML2 ( 3a – c ). At –40°C the betaine complex 3a inserts one equivalent of methylenecyclopropane to give the regioisomeric insertion products 5a and 6a in a 60:40 ratio. These products exhibit the cyclopropylidene moiety in the α- and β-positions, respectively, relative to zirconium. The corresponding hafnocene complexes 5b and 6b are obtained in a 70:30 ratio starting from 3b . The reaction of 3 ( a – c ) with allene gives a single insertion product ( 7a – c ) in each case where the exo-methylene group is in the α-position to the metal center ([2,1]-insertion). The complexes 5 – 7 are chiral. They all exhibit a pronounced ·-interaction of the internal –C4H=C5H double bond of the s̀-ligand chain with the metal center in addition to a metallocene/–C6H2–[B] ion pair interaction. The relative contributions of the cationic metallocene end of the dipolar complexes 5 – 7 are quite dependent on the steric and electronic properties of the respective metallocene units involved. This is revealed by a comparison to typical 13C-NMR parameters of the complexes 5 – 7 with a pair of suitable model complexes, namely the ethylene insertion product 4 into the betaine system 3a and its THF adduct 4 .THF.  相似文献   

14.
1,5-Dipolar Electrocyclization of Acyl-Substituted ‘Thiocarbonyl-ylides’ to 1,3-Oxathioles The reaction of α-diazoketones 15a, b with 4,4-disubstituted 1,3-thiazole-5(4H)-thiones 6 (Scheme 3), adamantanethione ( 17 ), 2,2,4,4-tetramethyl-3-thioxocyclobutanone ( 19 ; Scheme 4), and thiobenzophenone ( 22 ; Scheme 5), respectively, at 50–90° gave the corresponding 1,3-oxathiole derivatives as the sole products in high yields. This reaction opens a convenient access to this type of five-membered heterocycles. The structures of three of the products, namely 16c, 16f , and 20b , were established by X-ray crystallography. The key-step of the proposed reaction mechanism is a 1,5-dipolar electrocyclization of an acyl-substituted ‘thiocarbonyl-ylide’ (cf. Scheme 6). The analogous reaction of 15a, b with 9H-xanthen-9-thione ( 24a ) and 9H-thioxanthen-9-thione ( 24b ) yielded α,β-unsaturated ketones of type 25 (Scheme 5). The structures of 25a and 25c were also established by X-ray crystallography. The formation of 25 proceeds via a 1,3-dipolar electrocyclization to a thiirane intermediate (Scheme 6) and desulfurization. From the reaction of 15a with 24b in THF at 50°, the intermediate 26 (Scheme 5) was isolated. In the crude mixtures of the reactions of 15a with 17 and 19 , a minor product containing a CHO group was observed by IR and NMR spectroscopy. In the case of 19 , this side product could be isolated and was characterized by X-ray crystallography to be 21 (Scheme 4). It was shown that 21 is formed – in relatively low yield – from 20a . Formally, the transformation is an oxidative cleavage of the C?C bond, but the reaction mechanism is still not known.  相似文献   

15.
The synthesis of the trisaccharide α-L -Fucp-(1 → 2)-β-D -Galp-(1 → 3)-β-D -GalpNAc-1-OPr ( 2 ) is described. The N-acetylgalactosamine 6 was obtained from 4 by an intramolecular displacement of a (trifluoromethyl)sulfonyloxy by a pivaloyloxy group with its concomitant migration from position 3 to position 4 (Scheme 1). The galactosyl donor 9 was obtained from 7 via 8 by regioselective opening of the orthoester function with AcOH/pyridine followed by treatment with CCl3CN and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) (Scheme 2). Glycosylation of 6 with 9 in the presence of BF3 · OEt2 gave the disaccharide 10 . Selective deprotection of 10 at O? C(2′) followed by glycosylation with 12 and by standard deprotection afforded the title trisaccharide 2 (Scheme 3). Preliminary biological testing showed that 2 is able to inhibit the binding of the monoclonal antibody MBrl to the target tumor cells MCF7 in a dose-dependent manner.  相似文献   

16.
The application of the allyl-ester moiety as protecting principle for the carboxy group of N-acetylneuraminic acid is described. Peracetylated allyl neuraminate 2 is synthesized by reacting the caesium salt of the acid 1 with allyl bromide. Treatment of 2 with HCl in AcCl or with HF/pyridine gives the corresponding 2-chloro or 2-fluoro derivatives 3 and 4 , respectively (Scheme 1). In the presence of Ag2CO3, the 2-chloro carbohydrate 3 reacts with di-O-isopropylidene-protected galactose 5 to give the 2–6 linked disaccharide with the α-D -anomer 6a predominating (α-D /β-D = 6:1; Scheme 2). Upon activation of the 2-fluoro derivative 4 with BF3 · Et2O, the β-D -anomer 6b is formed preferentially (α-D /β-D = 1:5). In further glycosylations of 4 with long-chain alcohols, the β-D -anomers are formed exclusively (see 10 and 11 ; Scheme 4). The allyl-ester moiety can be removed selectively and quantitatively from the neuraminyl derivatives and the neuraminyl disaccharides by Pd(0)-catalyzed allyl transfer to morpholine as the accepting nucleophile (see Scheme 5).  相似文献   

17.
Alkylation of the chiral glycine derivative 2 with “activated” organohalides under ultrasound-assisted phasetransfer catalysis or with activated and nonactivated organohalides in anhydrous medium provides (mostly crystalline) alkylation products 3 . Acidic hydrolysis of the pure products 3 gives (aminoacyl)sultams 4 which by mild saponification furnish pure α-amino acids 5 in good overall yields from 2 , along with recovered auxiliary 1 (Scheme 1). Pure ω-protected α,ω-diamino acids and α-amino-ω-(hydroxyamino)acids 12–16 are readily accessible from (ω-haloacyl)sultams 3 via reaction with N-nucleophiles followed by acidic and basic hydrolyses (Scheme 2). A reliable determination of the enantiomeric purity of α-amino acids using HPLC analysis of their N-(3,5-dinitrobenzoyl)prolyl derivatives 17 is presented.  相似文献   

18.
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   

19.
The Me3Si? C(1) bond of the bis-(trimethylsilyl)ethynylated anhydroalditol 2 is selectively cleaved with BuLi to yield 3 / 4 , while AgNO2/KCN in MeOH cleaves the Me3Si? C(2′) bond, leading to 5 (Scheme 1). Both Me3Si groups are removed with NaOH in MeOH (→ 7 ), the (i-Pr)3Si group is selectively cleaved with HCl in aq. MeOH ( → 6 ); all silyl substituents are removed with Bu4NF ( → 8 ). Acetolysis transformed 9 into 13 , which was desilylated to 14 , while thiolysis of 9 led to a mixture 11 / 12 . The tetraacetate 14 has also been obtained from 9 via 10 . Oxidative dimerisation of either 3 or 5 , or of a mixture 3 / 5 yields only the homodimers 15 and 16 (Scheme 2); treatment of 16 with AgNO2/KCN yielded 17 , deprotection proceeding much more slowly than the cleavage of the Me3Si? C(2′) group of 2 . The iodoalkyne 20 , required for the cross-coupling with 5 according to Cadiot-Chodkiewicz, was prepared by deprotection of 3 / 4 to 18 , methoxymethylation (→ 19 ), and iodination. Cross-coupling yielded mostly 21 , besides the homodimer 22 . Similarly, cross-coupling of 20 and 23 (obtained from 5 ) led to 24 and 22 . The structure of 24 was established by X-ray analysis (Fig.), showing a C(6)–C(5′) distance of 5.2 Å. The conditions for deprotecting 2 were applied to 21 , and led to 25 (AgNO2/KCN), 26 (aq. NaOH), 27 (Bu4NF), and 29 (HCl/MeOH; Scheme 3). Attempted deprotection of the propargylic-ether moiety with BuLi, however, failed. The dimer 27 was further deprotected to 28 . Acetolytic (Ac2O/Me3SiOTf) debenzylation of the dimer 30 , obtained from 10 , gave 31 (83%) which was deacetylated to 32 (Scheme 4). Cross-coupling of 5 and the bromoalkyne 33 , obtained from 10 , yielded 34 ; again, acetolysis proceeded well, leading to 35 . The cellobiose derivative 38 was prepared from the lactone 36 via 37 . The glycosidic linkage of 38 proved resistant to the conditions of acetolysis, leading to 39 . Acetolysis of the benzylated thiophene 40 (from 30 with Na2S) yielded the octaacetate 41 , but proceeded in substantially lower yields (50%).  相似文献   

20.
A route for synthesizing C-nucleosides with 2,6-substituted pyridines as heterocyclic aglycones is described. Condensation of appropriately substituted lithiated pyridines with ribono-1,4-lactone derivatives yields hemiacetal 4a – g (Table 1), which can be reduced by Et3SiH and BF3·Et2O to the corresponding C-nucleoside (see Scheme 1 for 4d → β-D - 5 ). Conditions are presented that optimize the amount of the 2,6-dichloropyridine-derived β-D -anomer β-D - 5 formed (Table 3). Aminolysis of β-D - 5 yields the diaminonucleoside 14 (Scheme 3).  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