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 62 毫秒
1.
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   

2.
A new approach to ketose-derived nucteosides is described. It is based upon a chain elongation of 1-deoxy-1-nitroaldoses, followed by activation of the nitro group as a leaving group, and introduction of a pyrimidine or purine base. Thus, the nitroaldose 7 was prepared from 3 by pivaloylation (→ 4 ), synthesis of the anomeric nitrones 5/6 , and ozonolysis of 6 (Scheme 1). Partial hydrolysis of 4 yielded 8/9 , which were characterized as the acetates 10/11 and transformed into the nitrones 12/13 . Ozonolysis of 12/13 gave 14/15 , which were acetylated to 16/17 . Henry reaction of 7 lead to 19 and 20 , which were acetylated to 21 and 22 (Scheme 2). Michael addition of 7 to acrylonitrile and to methyl propynoate yielded the anomers 23/24 and 25/26 , respectively. Similar reactions of 16/17 were prevented by a facile β-elimination. Therefore, the nitrodiol 15 was transformed into the orthoesters 27 and then, by Henry reaction, partial hydrolysis, and acetylation, into 28 and 29 (Scheme 2). The structure of 19 was established by X-ray analysis. It was the major product of the kinetically controlled Henry reaction of 7 . Similarly, the β-D-configurated nitroaldoses 23 and 25 were the major products of the Michael addition. This indicates a preferred ‘endo’-attack on the nitronate anion derived from 7 . AMI calculations for this anion indicate a strong pyramidalization at C(1), in agreement with an ‘endo’-attack. Nucleosidation of 21 by 31 afforded 32 and 33 . Yields depended strongly upon the nature and the amount of the promoter and reached 77% for 33 , which was transformed into 34 , 35 , and the known ‘psicouridine’ ( 36 ; Scheme 3). To probe the mechanism, the trityl-protected 30 was nucleosidated yielding 37 , or 37 and 38 , depending upon the amount of FeCl3. Nucleosidation of the nitroacetate 28 was more difficult, required SnCl2 as a promoter, and yielded 39 and 40 . The β-D-anomer 40 was transformed into 36 . Nucleosidation of 23 (SnCl4) yielded the anomers 41 and 42 , which were transformed into 43 and 44 , and hence into 45 and 46 (Scheme 4). Similarly, nucleosidation of 25 yielded 47 and 48 , which were deprotected to 49 and 50 , respectively. The nucleoside 49 was saponified to 51 . Nucleosidation of 21 by 52 (SnCl2) afforded the adenine nucleosides 53 and 54 (Scheme 5). The adenine nucleoside 53 was deprotected (→ 55 → 56 ) to ‘psicofuranine’ (1), which was also obtained from 58 , formed along with 57 by nucleosidation of 28 . The structure and particularly the conformation of the nitroaldoses, nitroketoses, and nucleosides are examined.  相似文献   

3.
The synthesis of the phospha analogue 10 of DANA ( 2 ) is described. Bromo-hydroxylation of the known 11 (→ 12 and 13 ) followed by treatment of the major bromohydrin 13 with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) gave the oxirane 14 (Scheme 1). Depending on the solvent, TiBr4 transformed 14 into 16 or into a 15 / 16 mixture. Reductive debromination of 16 (→ 17 ), followed by benzylation provided 18 . Oxidattve decarboxylation (Pb(OAc)4) of the acid, obtained by saponification of 18 , yielded the anomeric acetates 19 and 20 . While 19 was inert under the conditions of phosphonoylation, the more reactive imidate 22 , obtained together with 23 from 19 / 20 via 21 (Scheme 2), gave a mixture of the phosphonates 24 / 25 and the bicyclic acetal 26 . Debenzylation of 24 / 25 and acetylation led to the acetoxyphosphonates 27 / 28 . Since β-elimination of AcOH from 27 / 28 proved difficult, the bromide 34 was prepared from 27 / 28 by photobromination and subjected to reductive elimination with Zn/Cu (→ 35 ; Scheme 3). This two-step sequence was first investigated using the model compounds 30 and 31 . Transesterification of 35 , followed by deacetylation gave 10 , which is a strong inhibitor of the Vibrio Cholerae sialidase.  相似文献   

4.
The cyclopentadienes 3a/b/c, 8b/c, 12a/b/d , and 16a/b/d were prepared as mixtures of regioisomers from the D -mannitol-derivatives 1,6,10 , and 14 and transformed into the ferrocenes 17, 18 , and 19 (73%; 38:17:45), 23, 24 , and 25 (70%; 6:42 :52), 26 (31%), and 27 (27%), respectively. Deprotection of 17–19 with HCl/MeOH gave the H2O-soluble ferrocenes 20–22 ; chloromercuration and iodination of 17 via 29 led to the C2-symmetric diiodo-ferrocene 30 . The mono(chloromercurio) derivative 28 , obtained as a by-product, was also transformed into 29 . The structure of the ferrocenes 18 and 19 , and of the bis(chloromercurio)ferrocene 29 has been established by X-ray analysis. The starting cyclopentadienes 3 were obtained in 50% yield from 1 by dialkylation of CpNa, followed by thermolysis of the spiro-annulated 2 . Similarly, dimesylate 6 (from 4 ) gave the spiro-annulated diene 7 and the annulated dienes 8 in about equal amounts; thermolysis transformed 7 into 8 (62%). The dienes 12 were prepared in 15% yield from the ditriflate 10 via 11 , the dimesylate 9 proving insufficiently reactive, and the dienes 16 (49%) from 14 via 15 .  相似文献   

5.
The synthesis of the 6-amino-6-deoxysialic-acid analogues 4, 5 , and 6 , is described. Mitsunobu reaction of the 1-C-nitroglycal 8 , (PPh3, HCOOH, DEAD) gave the formiate 10 with inversion of configuration at C(3) (Scheme 2). Treatment of 10 with aq. NH3 and subsequent protection of the amino function gave the imines 14 and 15 (Scheme 3), which were transformed into the triflates 17 . Substitution by azide, deprotection, and N-acetylation gave the anormeric 2-acetamido-3-azido-1-deoxy-1-nitro-D -mannoses 16 and the enol ether 18 . Chain elongation of the nitro azides 16 followed by hydroylsis gave the nonulosonates 20/22 , which upon reduction yielded the diols 23 and 24 , respectively (Scheme 4). The diol 23 was transformed into the sialic-acid analogues 5, 6 , and 32 by ozonolysis, transfer hydrogenation, hydorgenolysis, and deprotection (Scheme 5), and the diol 24 into 4 by a similar reaction sequence. The sialic-acid analogues 4 and 6 inhibit bacterial and viral sialidases competitively. The inbibitor constants for this enzyme from Vibrio cholerae are 0.12 mm for 4 and 0.19 mm for 6 , respectively. The activity of fowl plague virus sialidase was reduced by 17% and 36% under the influence of 4 and 6 , respectively, at a concentration of 0.1 mM . Compound 5 was inactive.  相似文献   

6.
On the Photochemistry of 2, 1-Benzisoxazoles (Anthraniles) and on the Thermal and Photochemical Decomposition of 2-Azido-acylbenzenes in Strongly Acidic Solution Anthranils 6 (Scheme 3), when irradiated with a mercury high-pressure lamp, in 96% sulfuric acid yielded, after work-up, 2-amino-5-hydroxy-acylbenzenes 8 and as side products 2-amino-3-hydroxy-acylbenzenes 9 (cf. Schemes 5–7 and Table 1). When C(5) of the anthranils 6 carries a methyl group a more complex reaction mixture is found after irradiation in 96% sulfuric acid (cf. Schemes 8 and 9): 3, 5-dimethyl-anthranil ( 6d ) yielded (after irradiation and acetylation) 2-acetyl- amino-5-methyl-acetophenone ( 15 ), 2-acetylamino-5-acetoxymethyl-acetophenone ( 18d ) and 2-acetylamino-5-acetoxy-6-methyl-acetophenone ( 12c ). The latter product was also formed after irradiation of 3, 4-dimethylanthranil ( 6c ) in 96% sulfuric acid. 3, 5, 7-Trimethyl-anthranil ( 6f ) formed under the same conditions 2-acetylamino-3, 5-dimethyl-acetophenone ( 15f ) and 2-acetylamino-5-acetoxymethyl-3-methyl-acetophenone ( 18f ). Since qualitatively the same product patterns were observed when the corresponding 2-azido-acetophenones 7 were decomposed in 96% sulfuric acid it is concluded that anthranilium ions (cf. 6b -H⊕, Scheme 11) on irradiation are transformed by cleavage of the N, O-bond into 2-acyl-phenylnitrenium ions (cf. 25b -H⊕) in the singlet ground state. The nitrenium ions are trapped directly by nucleophiles ( HSO ?4 in 96% sulfuric acid), thus, yielding the hydroxy-acetophenones 8 and 9 (Scheme 11). If C(5) is blocked by a methyl group a [1, 2]-rearrangement of the methyl group may occur (cf. Scheme 13) or loss of sulfuric acid can lead to quinomethane iminium ions (cf. 32-H⊕ , Scheme 13) which will react with HSO ?4 ions to yield, after hydrolysis and acetylation, the 5-acetoxymethyl substituted acetophenones 18d and 18f . It is assumed that the reduction products (2-acetylamino-acetophenones 15 ) are formed from the corresponding nitrenium ions in the triplet ground state.  相似文献   

7.
Cyclopentadienyl C-glycosides (= glycosyl-cyclopentadienes) have been prepared as latent fulvenes. Their reaction with nucleophiles leads to cyclopentadienes substituted with (protected) alditol moieties and, hence, to enantiomerically pure metallocenes. Treatment of 1 with cyclopentadienyl anion gave the epimeric glycosyl-cyclopentadienes 6 / 7 (Scheme 1). Each epimer consisted of a ca. 1:1 mixture of the 1, 3-and 1, 4-cyclopentadienes a and b , respectively, which were separated by prep. HPLC. Slow regioisomerisation occurred at room temperature. Diels-Alder addition of N-phenylmaleimide to 6a / b ca. 3:7 at room temperature yielded three ‘endo’-adducts, i.e., a disubstituted alkene ( 8 or 9 , 25%) and the trisubstituted alkenes 10 (45%) and 11 (13%). The structure of 10 was established by X-ray analysis. Reduction of 6 / 7 (after isolation or in situ) with LiAlH4 gave the cyclopentadienylmannitols 12a / b (80%) which were converted to the silyl ethers 13a / b (Scheme 2). Lithiation of 13a / b and reaction with FeCl2 or TiCl4 led to the symmetric ferrocene 14 (76%) and the titanocene 15 (34%), respectively. The mixed ferrocene 16 (63%) was prepared from 13a / b and pentamethylcyclopentadiene. Treatment of 6 / 7 with PhLi at ?78° gave a 5:3 mixture of the 1-C-phenylated alcohols 17a / b and 18a / b (71%) which were silylated to 19a / b and 20a / b , respectively. Lithiation of 19 / 20 and reaction with FeCl2 afforded the symmetric ferrocenes 21 and 22 and the mixed ferrocene 23 (54:15:31, 79%) which were partially separated by MPLC. The configuration at C(1) of 17–22 was assigned on the basis of a conformational analysis. The reaction of the ribofuranose 24 with cyclopentadienylsodium led to the epimeric C-glycosides 27a / b and 28a (57%, ca. 1:1, Scheme 3). The in-situ reduction of 27 / 28 with LiAlH4 followed by isopropylidenation gave 25a / b (65%) which were transformed into the ferrocene 26 (79%) using the standard method. Phenylation of 27 / 28 , desilylation, and isopropylidenation gave a 20:1 mixture of 33a / b and 34a / b (86%) which was separated by prep. HPLC. The same mixture was obtained upon phenylation of the fulvene 32 which was obtained in 36% yield from the reaction of the aldehydo-ribose 30 with cyclopentadienylsodium at ?100°. Lithiation of 33 / 34 and reaction with FeCl2 gave the symmetric ferrocene 35 (88%). Similarly, the aldehydo-arabinose 36 was transformed via the fulvene 37 (32%) into a 18:1 mixture of 38a / b and 39a / b (78%) and, hence, into the ferrocene 40 (83%). Conformational analysis allowed to assign the configuration of 33–35 , whereas an X-ray analysis of 40 established the (1S)-configuration of 38a / b and 40 . The opposite configuration at C(1) of 38a / b and 33a / b was established by chemical degradation (Scheme 4). Hydrogenation (→ 41 and 44 , resp.), deprotection (→ 42 and 45 , resp.), NaIO4 oxidation, and NaBH4 reduction yielded (+)-(S)- 43 and (?)-(R)- 43 , respectively.  相似文献   

8.
The synthesis and structure determination of adenosine‐derived monomeric building blocks for new oligonucleosides are described. Addition of Me3Si‐acetylide to the aldehyde derived from the known partially protected adenosine 1 led to the epimeric propargylic alcohols 2 and 3 , which were oxidised to the same ketone 4 , while silylation and deprotection led to 7 and 9 (Scheme 1). Introduction of an I substituent at C(8) of the propargylic silyl ethers 10 and 11 was not satisfactory. The protected adenosine 12 was, therefore, transformed in high yield into the 8‐chloro derivative 13 by deprotonation and treatment with PhSO2Cl; the iodide 15 was obtained in a similar way (Scheme 2). The 8‐Cl and the 8‐I derivatives 13 and 15 were transformed into the propargylic alcohols 17 , 18 , 25 , and 26 , respectively (Scheme 3). The propargylic derivatives 2 , 10 , 17 , 19 , 23 , 25 , and 27 were correlated, and their (5′R) configuration was determined on the basis of NOEs of the anhydro nucleoside 19 ; similarly, correlation of 3 , 11 , 18 , 20 , 24 , 26 , and 28 , and NOE's of 20 evidenced their (5′S)‐configuration.  相似文献   

9.
N-Nitroacetyl derivatives of L-proline, L-valine, and L-phenylalanine esters were prepared in two steps under mild conditions (Scheme 2). Regiospecific mono- and bis-allylation of these nitroacetyl derivatives were accomplished in presence of a Pd(0) catalyst. The bis-allyl derivatives 7–9 were obtained in 40–75% yield. The tertiary NO2 group in these compounds could be transformed into an acetylamino group by Zn/AcOH/Ac2O. The final products 11–13 are dipeptides in which the N-terminal glycine residue bears two α-allyl substituents.  相似文献   

10.
A facile synthetic approach was adopted towards the synthesis of benzo‐fused macrocyclic lactams 2a – 2g via the base‐catalyzed condensation reaction of 2,2′‐[alkanediylbis(oxy)]bis[benzaldehydes] 3a – 3c with N,N′‐substituted bis[2‐cyanoacetamide] derivatives 7a – 7c (Scheme 2). The latter compounds were obtained by the reaction of the appropriate diamines 6a – 6c with ethyl 2‐cyanoacetate ( 4 ). Attempts to prepare the oxaaza macrocycles 2 by alternative pathways were also investigated. The novel pyrazolo‐fused macrocycles 13a and 13b were obtained in 48 and 52% yield, respectively, upon treatment of 2d and 2g with NH2NH2?H2O at 100° (Scheme 4).  相似文献   

11.
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   

12.
Addition of CH2N2 to 2,3:5,6-di-O-isopropylidene-1-thio-mannono-1,4-lactone ( 1 ) gave the 2,5-dihydro-1,3,4-thiadiazole 2 and the 4,5-dihydro-1,2,3-thiadiazole 3 . First-order kinetics were observed for the thermolysis of 3 (Scheme 3) at 80–110° in C6D5Cl solution and of 2 (Scheme 3) at 20–35° in CDC13, respectively. The 1,2,3-thiadiazole 3 led to mixtures of the thiirane 9 , the starting thionolactone 1 , the thiono-1,5-lactone 8 , and the enol ether 7 , while the isomeric 1,3,4-thiadiazole 2 led to mixtures of the anomeric thiiranes 9 and 12 , the O-hydrogen S,O,O-ortholactone α-D - 14 , the S-methyl thioester 15 , the S,S,O-ortholactone 13 , and the 2,3:5,6-di-Oisopropylidene-mannono-1,4-iactone ( 16 ). Pure products of the thermolysis were isolated by semipreparative supercritical fluid chromatography (SFC), whereas preparative HPLC led to partial or complete decomposition. Thus, the β-D -mannofuranosyl β-D -mannofuranoside 10 , contaminated by an unknown S species, was isolated by preparative HPLC of the crude product of thermolysis of 3 at 115–120° and partially transformed in CD3OD solution into the symmetric di(α-D -mannofuranosyl) tetrasulfide 11 . Its structure was evidenced by X-ray analysis. Similarly, HPLC of the thermolysis product of 2 gave the enethiol 17 , the sulfide 19 , and the mercapto alcohol 18 as secondary products. Thermolysis of the thiirane 9 at 110–120° (Scheme 4) led to the anomeric thiirane 12 which was transformed into mixtures of the enethiol 17 and the enol ether 7. Addition of H2O to 17 and 7 gave the corresponding hemiacetals 18 and 20. The mechanism of the thermolysis of the dihydrothiadiazoles 2 and 3 , and the thiiranes 9 and 12 is discussed.  相似文献   

13.
The tricyclic alcohols 3–7 , derived from the corresponding ketones 1 and 2 (Scheme 1), by action of acids underwent dehydration with skeletal rearrangements. Dehydration of 3 and 4 with POCl3/pyridine (procedure A) afforded the polycyclic hydrocarbons 9, 10 , and 12, 13 , respectively. With TsOH (procedure B), on the other hand, 3 and 4 gave homo-triquinacenes 10 and 14 respectively, as well as the polycyclic ethers 11 and 15 (Scheme 2). Hydrocarbon 9 (or 12 ) was converted into 10 FSO3H to the tertiary alcohol 16 (Scheme 4). Plausible mechanisms for these transformations are summarized in Scheme 8. Dehydration of the secondary alcohols 5 and 7 was effected by procedure A. While treatment of alcohol 5 with POCl3/pyridine yielded two isomeric hydrocarbons 17 and 18 , similar dehydration of its epimeric alcohol 7 afforded hydrocarbon 21 as the sole product. The tertiary alcohol 6 was dehydrated by both procedures to yield two isomeric hydrocarbons 19 and 20 (Scheme 5). Hydrocarbon 20 was converted into 19 by procedure B (mechanisms, Scheme 10). Reaction of ketone 2 with CF3COOH gave the addition product 22 converted into vinylsulfonyl fluorides 24 and 25 by treatment with FSO3H (Scheme 6). Homo-triquinacenes 10 and 14 reacted smoothly with 4-phenyl-1,2,4-triazoline-3,5-dione to give the ‘ene’-reaction products 26 and 27 , respectively.  相似文献   

14.
Transamidation Reactions with Cyclic Amino-amides Lactames which are substituted at the nitrogen atom by a 3-aminopropyl residue are transformed under base catalysis to cyclic amino-amides enlarged by 4 ring atoms. The formed ring must be at minimum 12-membered. Scheme 2 illustrates this result: the 8-membered 7 is transamidated in 96% yield to the 12-membered ring 8 (in the presence of potassium 3-aminopropylamid in 1, 3-propanediamine), the 9-membered 10 to the 13-membered ring 11 (97%) and the 11-membered 14 to the 15-membered ring 15 . Furthermore, the 13-membered ring 27 (Scheme 5) is transformed to the 17-membered 28 . In the case of the 15-membered lactame 15 it is demonstrated that 14 is not formed back under the conditions of the transamidation. Large ring lactames which are substituted at the nitrogen atom by a 3-(alkylamino) propyl group lead under base catalysis to an equilibrium mixture, e.g. the 17-membered 26 is in equilibrium with the 21-membered 29 . This result is similar to the behavior of the corresponding open-chain amino-amides [2]. Because of transannular interactions, the 11-membered ring 2 is not stable: transamidation of the 7-membered 1 (Scheme 1) doesn't give the expected 2 , but its water elimination product 3 in small yield. The N-tosyl derivative of 2 , namely 20 , is synthesized by an independent route (Scheme 3). Detosylation of 20 yields the 7-membered 1 instead of 2 . Concerning the mechanism of this interesting reaction see Scheme 4.  相似文献   

15.
Synthesis of Triafulvene-Precursors from Trisubstituted Cyclopropanes Trisubstituted cyclopropanes 5a–f are prepared by carbene addition to the appropriate olefins. While 5a (Y = OAc) and 5c (Y = Cl) rearrange in the presence of BuLi, 5d (Y = SPh) is stable enough to allow the envisaged sequence for triafulvene (Scheme 2) : halogen-Li exchange, followed by methylation of 6d , gives 7d in a 93% yield; after base-induced elimination of HB r from 7d , the key precursor 1-methylene-2-(phenylthio)cyclopropane ( 9 , 70% yield) is isolated. Compound 9 is transformed into the corresponding sulfoxide 10 (83%), sulfone 11 (80%), and sulfonium fluoroborate 12 (95% yield) by subsequent oxidation and methylation, respectively. Some 1H-NMR results of cyclopropanes 5a–f and 7d as well as of methylidene-cyclopropanes 9–11 are discussed.  相似文献   

16.
Dialkynes of the type 3 (Scheme 1) are regioselectively deprotected by treating them either with base in a protic solvent (→ 4 ), or– after exposing the OH group– by catalytic amounts of base in an aprotic solvent (→ 5 and 8 ). The Me3Si-protected 12 (Scheme 2) is inert to catalytic BuLi/THF which transformed 11 into 9 , while K2CO3/MeOH transformed both 10 into 9 , and 12 into 13 , evidencing the requirement for a more hindered (hydroxypropyl)silyl substituent. C-Silylation of the carbanions derived from 17–19 (Scheme 3) with 15 led to 20–22 , but only 22 was obtained in reasonable yields. The key intermediate 27 was, therefore, prepared by a retro-Brook rearrangement of 23 , made by silylating the hydroxysulfide 16 with 15 . The OH group of 27 was protected to yield the {[dimethyl(oxy)propyl]dimethylsilyl}acetylenes (DOPSA's) 21, 28 , and 29 . The orthogonally protected acetylenes 20–22, 28 , and 29 were de-trimethylsilylated to the new monoprotected acetylene synthons 30–34 . The scope of the orthogonal protection was checked by regioselective deprotection of the dialkynes 39–42 (Scheme 4), prepared by alkylation of 35 (→ 39 ), or by Pd0/CuI-catalyzed cross-coupling with 36–38 (→40–42 ). The cross-coupling depended upon the solvent and proceeded best in N,N,N′,N′ -teramethylethylenediamine (TMEDA). Main by-product was the dimer 43 . On the one hand, K2CO3/MeOH removed the Me3Si group and transformed 39–42 into the monoprotected 44–47 ; catalytic BuLi/THF, on the other hand, transformed the alcohols 48–51 , obtained by hydrolysis of 39–42 , into the monoprotected dialkynes 52–55 , all steps proceeding in high yields. Addition of the protected DOPSA groups to the lactones 56 (→57–59 ) and 62 (→63 ) (Schemes 5 and 6) gave the corresponding hemiketals. Reductive dehydroxylation of 57 and 58 failed; but similar treatment of 59 yielded the alcohol 61 . Similarly, 63 was transformed into 64 which was protected as the tetrahydropyranyl (Thp) ether 65 . In an optimized procedure, 62 was treated sequentially with lithiated 31 , BuLi, and Me3SiCl (→ 66 ), followed by desilyloxylation to yield 60% of 67 , which was protected as the Thp ether 68 . Under basic, protic conditions, 68 yielded the monoprotected bisacetylene 69 ; under basic, aprotic conditions, 67 led to the monoprotected bisacetylene 70 . These procedures are compatible with the butadiynediyl function. The butadiyne 73 was prepared by cross-coupling the alkyne 69 and the iodoalkyne 71 (obtained from 70 , together with the triiodide 72 ) and either transformed to the monosilylated 76 or, via 77 , to the monosilylated 78 . Formation of the homodimers 74 and 75 was greatly reduced by optimizing the conditions of cross–coupling of alkynes.  相似文献   

17.
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   

18.
The reaction of N‐phenylbenzamides 5 with excess SOCl2 under reflux gave N‐phenylbenzimidoyl chlorides 6 , which, on treatment with KSeCN in acetone, yielded imidoyl isoselenocyanates of type 2 . These products, obtained in almost quantitative yield, were stable in the crystalline state. They were transformed into selenourea derivatives 7 by the reaction with NH3, or primary or secondary amines. In acetone at room temperature, 7 reacted with activated bromomethylene compounds such as 2‐bromoacetates, acetamides, and acetonitriles, as well as phenacyl bromides and 4‐cyanobenzyl bromide to to give 1,3‐selenazol‐2‐amines of type 9 (Scheme 2). A reaction mechanism via alkylation of the Se‐atom of 7 , followed by ring closure and elimination of aniline, is most likely (cf. Scheme 7). In the case of selenourea derivatives 7d and 7l with an unsubstituted NH2 group, an alternative ring closure via elimination of H2O led to 1,3‐selenazoles 10a and 10b , respectively (Schemes 4 and 7). On treatment with NaOH, ethyl 1,3‐selenazole‐5‐carboxylates 9l and 9s were saponified and decarboxylated to give the corresponding 5‐unsubstituted 1,3‐selenazoles 12a and 12b (Scheme 6). The molecular structures of selenourea 7f and the 1,3‐selenazoles 9c and 9d have been established by X‐ray crystallography (Figs. 1 and 3).  相似文献   

19.
To demonstrate the relevance of the kinetic acidity of individual OH groups for the regioselectivity of glycosylation by glycosylidene carbenes, we compared the glycosylation by 1 of the known triol 2 with the glycosylation of the diol D - 3 and the fluorodiol L - 4 . Deoxygenation with Bu3SnH of the phenoxythiocarbonyl derivative of 5 (Scheme 1) or the carbonothioate 6 gave the racemic alcohol (±)- 7 . The enantiomers were separated via the allophanates 9a and 9b , and desilylated to the deoxydiols D - and L - 3 , respectively. The assignment of their absolute configuration is based upon the CD spectra of the bis(4-bromobenzoates) D - and L - 10 . The (+)-(R)-1-phenylethylcarbamates 13a and 13b (Scheme 2) were prepared from the fluoroinositol (±)- 11 via (±)- 4 and the silyl ether (±)- 12 and separated by chromatography. The absolute configuration of 13a was established by X-ray analysis. Decarbamoylation of 13a ( → L - 12 ) and desilylation afforded the fluorodiol L - 4 . The H-bonds of D - 3 and L - 4 in chlorinated solvents and in dioxane were studied by IR and 1H-NMR spectroscopy (Fig. 2). In both diols, HO? C(2) forms an intramolecular, bifurcated H-bond. There is an intramolecular H-bond between HO? C(6) and F in solutions of L - 4 in CH2Cl2, but not in 1,4-dioxane; the solubility of L - 4 in CH2Cl2 is too low to permit a meaningful glycosidation in this solvent. Glycosidation of D - 3 in dioxane by the carbene derived from 1 (Scheme 3) followed by acetylation gave predominantly the pseudodisaccharides 18/19 (38%), derived from glycosidation of the axial OH group besides the pseudodisaccharides 16 / 17 (13%) and the epoxides 20 / 21 (7%), derived from protonation of the carbene by the equatorial OH group. Similarly, the reaction of L - 4 with 1 (Scheme 4) led to the pseudodisaccharides 28 / 29 (46%) and 26 / 27 (14%), derived from deprotonation of the axial and equatorial OH groups, respectively. Formation of the epoxides involved deprotonation of the intramolecularly H-bonded tautomer, followed by intramolecular alkylation, elimination, and substitution (Scheme 4). The regio- and diastereoselectivities of the glycosidation correlate with the H-bonds in the starting diols.  相似文献   

20.
Synthesis of the Sesquiterpene Ketone Shyobunon and of its Diastereoisomers Shyobunon ( 12 ) and 6-epishyobunon ( 13 ) as well as their epimers 10 and 11 were synthetized in five steps from geranyl- ( 1 ) and nerylsenecionate ( 2 ), respectively. Ester enolate rearrangement [5] of 1 and 2 furnished the key intermediates 3 and 4 in high yield and in about 80% stereoselectivity [6] (Scheme 1). Conversion of the acid mixture 3 / 4 to the cyclohexanone derivatives 7 and 8 succeeded in 35–40% yield by means of cyclization of their acidchlorides with tin tetrachloride to the mixture of 5 and 6 , followed by HCl elimination with diazabicyclononene (DBN) (Scheme 2). Selective reduction of 7 to 10 and 11 , and 8 to 12 and 13 with triphenyltinhydride completed the synthesis. The relative configuration of 10 and 11 as well as of 12 and 13 were deduced from the 13C-NMR. spectra (Scheme 4, Table 2). The structure of ‘epishyobunone’ is revised: it has the structure 13 , and not 11 as described earlier [1]. This is discussed in connection with the rearrangement of acoragermacrone ( 16 ) [18] to shyobunone ( 12 ) and 6-epishyobunone ( 13 ) (Scheme 5).  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