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 cellobiose-derived dialkynes 14 and 15 were prepared by glycosidation of the acceptor 9 with the thioglycosides 12 (82%) and 13 (85%), respectively. The acceptor 9 was prepared from the known alcohol 2 via the lactone 7 in five steps (48% overall), and the donors 12 and 13 were prepared from the alkynylated anhydroglucose derivative 10 (60% overall). Acetolytic debenzylation of 14 and 15 (→ 16 and 17 , resp.) followed by deacylation of 16 yielded 60% of the cellobiose-derived dialkyne 18 . Deacylation of 14 (→ 19 ), methoxymethylation (→ 20 ) and trimethylgermylation led to the orthogonally protected dialkyne 21 (69% overall). Protodesilylation of 21 with K2CO3/MeOH gave 22 (90%), while the Me3Ge group was selectively removed with CuBr (19 mol-%) in THF/MeOH to give 20 (95%). Treatment of 21 with aqueous HCl solution led to 19 (80%). Bromodegermylation of 21 (NBS/AgOOCCF3) led to a mixture of 23 (85%) and 24 (11%). Similar conditions using CuBr instead of AgOOCCF3 gave exclusively the bromoalkyne 23 (93%). The temperature dependence of the δ values of the OH resonances of 18 in (D6)DMSO evidence a strong intramolecular H-bond between C(5′)? O…?HO? C(5).  相似文献   

2.
A 7-deazaadenosine ( = tubercidin; c7A; 1 ) building block for solid-phase oligoribonucleotide synthesis was prepared. The amino group of 1 was protected with the (dimethylamino)methylidene residue (→ 3 ), and the monomethoxytrityl group was introduced at OH? C(5′) (→ 4 ). Protection of OH? C(2′) was carried out by silylation, showing that use of the (i-Pr)3Si group resulted in high 2′-O-selectivity (→ 5b , 80%). Reaction of 5b with PCl3 afforded the phosphonate 7 which was used in solid-phase oligoribonucleotide synthesis. The autocatalytic hydrolysis of hammerhead ribozymes using pG-G-G-A-G-U-C-A-G-U-C-C-C-U-U-C-G-G-G-G-A-C-U-C-U-G-A-A-G-A-G-G-C-G-C as substrate strand (S) and modified G-C-G-C-C-G-A-A-A-C-U-C-C-C as enzyme strand (E) was studied. When c7A replaced A13 or A14, a small decrease of catalytic activity was observed, while modification in position A15 enhanced the autocatalytic hydrolysis. The results demonstrate, that the atom N(7) of adenosine in any of these positions is not crucial for ribozyme action.  相似文献   

3.
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   

4.
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   

5.
The novel 8,14‐secoursatriene derivative 6 was synthesized starting from ursolic acid ( 1 ) via methyl esterification of the 17‐carboxylic acid group and benzoylation of the 3‐hydroxy group (→ 2 ; Scheme 1), ozone oxidation of the C(12)?C(13) bond (→ 3 ), dehydrogenation with Br2/HBr (→ 4 ), enol acetylation of the resulting carbonyl group (→ 5 ; Scheme 2), and ring‐C opening with the aid of UV light (→ 6 ). Ring‐C‐opened dienone derivative 7 of ursolic acid was also obtained via selective hydrolysis of 6 (Scheme 2). Both compounds 6 and 7 are key intermediates for the preparation of chiral decalin synthons from ursolic acid.  相似文献   

6.
Z-Protected diphenyl (decarboxy-dipeptidyl)phosphonates 5a - c with a (decarboxysarcosinyl)phosphonate moiety are prepared from Z-L-alanine ( 1a ). Z-L-valine ( 1b ), and Z-L-phenylalanine ( 1c ) by the following series of steps: coupling with methyl sarcosinate (→ 2a – c ), saponification (→ 3a – c ), Hofer-Moest oxidative decarboxyiation by electrolysis in MeOH (→ 4a – c ), and Arbuzov reaction with P(OPh)3/TiCl4 (Scheme 3). Double deprotonation and alkylation lead to non-stereoselective incorporation of side chains next to the phosphonate group (products of type 6 – 8 , nine examples, see Scheme 4). In the cases of 6a – c and 8c , the diastereoisomers could be separated and the configuration of the newly formed stereogenic center deduced. We assign the L,D-configuration to the diastereoisomers for which the 31 P-NMR signal appears at higher field.  相似文献   

7.
The C? C coupling of the two bicyclic, unsaturated dicarboximides 5 and 6 with aryl and heteroaryl halides gave, under reductive Heck conditions, the C‐aryl‐N‐phenyl‐substituted oxabicyclic imides 7a – c and 8a – c (Scheme 3). Domino‐Heck C? C coupling reactions of 5, 6 , and 1b with aryl or heteroaryl iodides and phenyl‐ or (trimethylsilyl)acetylene also proved feasible giving 8, 9 , and 10a – c , respectively (Scheme 4). Reduction of 1b with LiAlH4 (→ 11 ) followed by Heck arylation and reduction of 5 with NaBH4 (→ 13 ) followed by Heck arylation open a new access to the bridged perhydroisoindole derivatives 12a , b and 14a , b with prospective pharmaceutical activity (Schemes 5 and 6).  相似文献   

8.
Intramolecular Diels–Alder (IMDA) transition structures (TSs) and energies have been computed at the B3LYP/6‐31+G(d) and CBS‐QB3 levels of theory for a series of 1,3,8‐nonatrienes, H2C?CH? CH?CH? CH2? X? Z? CH?CH2 [? X? Z? =? CH2? CH2? ( 1 ); ? O? C(?O)? ( 2 ); ? CH2? C(?O)? ( 3 ); ? O? CH2? ( 4 ); ? NH? C(?O)? ( 5 ); ? S? C(?O)? ( 6 ); ? O? C(?S)? ( 7 ); ? NH? C(?S)? ( 8 ); ? S? C(?S)? ( 9 )]. For each system studied ( 1 – 9 ), cis‐ and trans‐TS isomers, corresponding, respectively, to endo‐ and exo‐positioning of the ? C? X? Z? tether with respect to the diene, have been located and their relative energies (ErelTS) employed to predict the cis/trans IMDA product ratio. Although the ErelTS values are modest (typically <3 kJ mol?1), they follow a clear and systematic trend. Specifically, as the electronegativity of the tether group X is reduced (X?O→NH or S), the IMDA cis stereoselectivity diminishes. The predicted stereochemical reaction preferences are explained in terms of two opposing effects operating in the cis‐TS, namely (1) unfavorable torsional (eclipsing) strain about the C4? C5 bond, that is caused by the ? C? X? C(?Y)? group’s strong tendency to maintain local planarity; and (2) attractive electrostatic and secondary orbital interactions between the endo‐(thio)carbonyl group, C?Y, and the diene. The former interaction predominates when X is weakly electronegative (X?N, S), while the latter is dominant when X is more strongly electronegative (X?O), or a methylene group (X?CH2) which increases tether flexibility. These predictions hold up to experimental scrutiny, with synthetic IMDA reactions of 1 , 2 , 3 , and 4 (published work) and 5 , 6 , and 8 (this work) delivering ratios close to those calculated. The reactions of thiolacrylate 5 and thioamide 8 represent the first examples of IMDA reactions with tethers of these types. Our results point to strategies for designing tethers, which lead to improved cis/trans‐selectivities in IMDAs that are normally only weakly selective. Experimental verification of the validity of this claim comes in the form of fumaramide 14 , which undergoes a more trans‐selective IMDA reaction than the corresponding ester tethered precursor 13 .  相似文献   

9.
Solid-phase synthesis of the oligo(2′-deoxynucleotides) 19 and 20 containing 2′-deoxy-β-D -xylocytidine ( 4 ) is described. For this purpose, 1-(2-deoxy-β-D -threo-pentofuranosyl)cytosine ( = 1-(2-deoxy-β-D -xylofuranosyl)-cytosine; 4 ) was protected at its 4-NH2 group with a benzoyl (→ 5 ) or an isobutyryl (→ 8 ) residue, and a dimethoxytrityl group was introduced at 5′-OH (→ 7, 10 ; Scheme 2). Compounds 7 and 10 were converted into the 3′-phosphonates 11a,b . While 19 could be hybridized with 21 and 22 under formation of duplexes with a two-nucleotide overhang on both termini ( 19 · 21 : Tm 29°; 19 · 22 : Tm 22°), the decamer 20 bearing four xCd residues could no longer be hybridized with one of the opposite strands. Moreover, the oligonucleotides d[(xC)8? C] ( 13 ), d[(xC)4? C] ( 14 ), d[C? (xC)4? C] ( 15 ), and d[C? (xC)3? C] ( 16 ) were synthesized. While 13 exhibits an almost inverted CD spectrum compared to d(C9) ( 17 ), the other oligonucleotides show CD spectra typical for regular right-handed single helices. At pH 5, d[(xC)8? C] forms a stable hemi-protonated duplex which exhibits a Tm of 60° (d[(CH+)9] · d(C9): Tm 36°). The thermodynamic parameters of duplex formation of ( 13H + · 13 ) and ( 17H + · 17 ) were calculated from their melting profiles and were found to be identical in ΔH but differ in ΔS ( 13H + · 13 : ΔS = ?287 cal/K mol; 17H + · 17 : ΔS = ?172 cal/K mol).  相似文献   

10.
The free energies of the rotational barriers, ΔG*, about ?CH? NMe2 bond in N′-heteroaryl N,N-dimethylformamidines (A), about ?CH? NEt2 bond in N-heteroaryl N,N-diethylformamidines (B), and about ?C(Me)? NMe2 bond in N′-heteroaryl N,N-dimethylacetamidines (C) have been found to be in the range 17.5–20.1 kcal/mole for type A, 18.8–21.6 kcal/mole for type B and 13–14 kcal/mole or below for type C of compounds, respectively. The compounds of the types A and B exist in the forms IIa, IIIa, IV, V, and VI, while the compounds of the type C exist in the forms IIb and IIIb.  相似文献   

11.
The synthesis of oligonucleotides containing N7-(2-deoxy-β-D -erythro-pentofuranosyl)guanine (N7Gd; 1 ) is described. Compound 1 was prepared by nucleobase-anion glycosylation of 2-amino-6-methoxypurine ( 5 ) with 2-deoxy-3,5-di-O-(4-toluoyl)-α-D -erythro-pentofuranosyl chloride ( 6 ) followed by detoluoylation and displacement of the MeO group ( 8→10→1 ). Upon base protection with the (dimethylamino)methylidene residue (→ 11 ) the 4,4-dimethoxytrityl group was introduced at OH? C(5′) (→ 12 ). The phosphonate 3 and the phosphoramidite 4 were prepared and used in solid-phase oligonucleotide synthesis. The self-complementary dodecamer d(N7G? C)6 shows sigmoidal melting. The Tm of the duplex is 40°. This demonstrates that guanine residues linked via N(7) of purine to the phosphodiester backbone are able to undergo base pairing with cytosine.  相似文献   

12.
Replacement and Oxidation Reactions of N-Dichlorophosphanyl Triphenylphosphazene, Ph3P?N? PCl2 The title compound ( 1 ) reacts with MeOH, EtOH, PhOH, EtSH, and water forming N-phosphanyl or N-phosphinoyl phosphazenes, resp., Ph3P?N? PX2 (X ? OPh( 8 ), SEt( 9 )) or Ph3P?N? PH(O)X (X ? Cl( 3 ), OH( 4 ), OMe( 5 ), OEt( 7 )). The reaction of 1 with P(NEt2)3 yields Ph3P?N? P(NEt2)2 ( 10 ). Ph3P?N? PF2( 11 ) and Ph3P?N? PH(O)F ( 12 ) are obtained by chlorine-fluorine exchange. The N-phosphanyl compounds 1 , 8 , 9 and 11 are oxidized by NO2 yielding the corresponding N-phosphoryl derivatives, Ph3P?N? P(O)X2 (X ? Cl( 2 ), OPh( 13 ), SEt ( 14 ), F( 15 )). The thiophosphoryl compounds, (Ph3P?N? P(S)X2 (X ? Cl( 16 ), OPh( 17 ), F( 18 )) are obtained by oxidizing 1 , 8 , and 11 with sulfur.  相似文献   

13.
The C(2) isotropic chemical shift values in solid‐state CP/MAS 13C NMR spectra of conformational polymorphs Form I (δ 28.5) and III (δ 22.9) of (1S,4S)‐sertraline HCl ( 1 ) were correlated with a γ‐gauche effect resulting from the respective 162.6° antiperiplanar and 68.8° (+)‐synclinal C(2)? C(1)? N? CH3 torsion angles as measured by X‐ray crystallography. The similarity of the solution‐state C(2) chemical shifts in CD2Cl2 (δ 22.8) and DMSO‐d6 (δ 23.4) with that for Form III (and other polymorphs having C(2)? C(1)? N? CH3 (+)‐synclinal angles) strongly suggests that a conformational bias about the C(1)? N bond exists for 1 in both solvents. This conclusion is supported by density functional theory B3LYP/6‐31G(d)‐calculated relative energies of C(1)? N rotameric models: (kcal) 0.00 [73.8 °C(2)? C(1)? N? CH3 torsion angle], 0.88 (168.7°), and 2.40 (?63.4°). A Boltzmann distribution of these conformations at 25 °C is estimated to be respectively (%) 80.3, 18.3, and 1.4. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

14.
(R)-3-Hydroxybutyric acid (from the biopolymer PHB) and pivalaldehyde give the crystalline cis - or (R,R)-2-(tert-butyl)-6-methyl-1,3-dioxan-4-one ( 1a ), the enolate of which is stable at low temperature in THF solution and can be alkylated diastereoselectively ( →3, 4, 5 , and 7 ). Phenylselenation and subsequent elimination give an enantiomerically pure enol acetal 10 of aceto-acetic acid. Some reactions of 10 have been carried out, such as Michael addition (→ 11 ), alkylation on the CH3 substituent (→ 13 ), hydrogenation of the C?C bond (→ 1a ) and photochemical cycloaddition (→ 16 ). The overall reactions are substitutions on the one stereogenic center of the starting β-hydroxy acid without racemization and without using a chiral auxiliary.  相似文献   

15.
The protected disaccharide 44 , a precursor for the synthesis of allosamidin, was prepared from the glycosyl acceptor 8 and the donors 26–28 , best yields being obtained with the trichloroacetimidate 28 (Scheme 6). Glycosidation of 8 or of 32 by the triacetylated, less reactive donors 38–40 gave the disaccharides 46 and 45 , respectively, in lower yields (Scheme 7). Regioselective glycosidation of the diol 35 by the donors 38–40 gave 42 , the axial, intramolecularly H-bonded OH? C(3) group reacting exclusively (Scheme 5). The glycosyl acceptor 8 was prepared from 9 by reductive opening of the dioxolane ring (Scheme 3). The donors 26–28 were prepared from the same precursor 9 via the hemiacetal 25 . To obtain 9 , the known 10 was de-N-acetylated (→ 18 ), treated with phthalic anhydride (→ 19 ), and benzylated, leading to 9 and 23 (Schemes 2 and 3). Saponification of 23 , followed by acetylation also gave 9 . Depending upon the conditions, acetylation of 19 yielded a mixture of 20 and 21 or exclusively 20 . Deacetylation of 20 led to the hydroxyphthalamide 22 . De-N-acetylation of the 3-O-benzylated β-D -glycosides 11 and 15 , which were both obtained from 10 , was very sluggish and accompanied by partial reduction of the O-allyl to an O-propyl group (Scheme 2). The β-D -glycoside 30 behaved very similarly to 11 and 15 . Reductive ring opening of 31 , derived from 29 , yielded the 3-O-acetylated acceptor 32 , while the analogous reaction of the β-D -anomer 20 was accompanied by a rapid 3-O→4-O acyl migration (→ 34 ; Scheme 4). Reductive ring opening of 21 gave the diol 35 . The triacetylated donors 38–40 were obtained from 20 by debenzylidenation, acetylation (→ 36 ), and deallylation (→ 37 ), followed by either acetylation (→ 38 ), treatment with Me3SiSEt (→ 39 ), or Cl3CCN (→ 40 ).  相似文献   

16.
CH-Acidity in α-position to the N-Atom of N, N -Dialkylamides with Sterically Protected Carbonyl Groups Contribution to the Nucleophilic Amino Alkylation Sterically protected amides 1 such as the 2,4,6-triisopropyl-benzoic acid derivatives 3, 8b and 10 undergo readily H/Li-exchange with s-butyllithium at the CH3N- or CH2N-groups. The resulting organolithium compounds (cf. 9, 11 ) are alkylated and hydroxyalkylated with primary haloalkanes, aldehydes, and ketones under chain elongation in the amine position of the amides. The (E/Z)-rotamers of the dialkylamides 7 and 8 are separated by chromatography; the amides 4 – 6 , 12 , and 13 formally derived from β-hydroxyamines are obtained in the (Z)-form only. The configurational (E/Z)-assignments follow from NMR. and IR. data. The erythro and threo configuration of the two diastereomeric amides 12a and 12b are tentatively concluded from Eu(fod)3-1H-NMR.-shift experiments. The results strongly suggest that the H/Li-exchange takes place regioselectively at the CH? N group which is in cis-position to the C?O double bond (→ 14 ). The methyl 2,4,6-tri(t-butyl)benzoate ( 18 ) can also be deprotonated to the lithium acyloxymethanide 19 which is trapped by alkylation with 1-iodooctane (→ 20 ). – The steric protection of the carbonyl groups in the products 4 – 8, 10, 12, 13 , and 20 prevents their ready hydrolysis to amines and alcohols, respectively. Therefore, triphenylacetic acid derivatives 21 rather than 2,4,6-triisopropylbenzoic acid derivatives for use in the electrophilic substitution of equation (1) are recommended. The trityl group in 21 may be considered a C-leaving-group (C? C protective group, cf. 22, 23 ). The acetamide 25 reacts readily (→ 26 ) and then with electrophiles to give products 27a – c . As shown in the Table, the amides 27 are cleaved under a variety of conditions with formation of triphenylmethane. LiAlH4 produces a tertiary amine, CH3Li a secondary amine, and dissolving alkali metals/naphthalene under aprotic conditions mixtures of secondary amine and its formamide (hydrolysed by acid treatment). Thus the overall process (2) is feasible.  相似文献   

17.
Induced by a stoichiometric excess of dichloro(diisopropoxy)titanium, 1-(trimethylsiloxy)cyclohexene and p-substituted β-nitrostyrenes (Y = H,CH3,CH3O,CN) combine in CH2Cl2 solution at ?90° preferentially with relative topicity ul – opposite to the corresponding reaction of enolates or enamines. The primary products are the bicyclic nitronates 3–5 which can be separated, and which are cleaved by KF in MeOH to give the aryl(nitroethyl)-substituted cyclohexanones 1 and 2 (Tables 1 and 2, two typical procedures are given). The major products (2:1 to 4:1) are the hitherto not readily available diastereoisomers 2 of l-configuration. Instead of being solvolyzed, the bicyclic nitronate 5 can be used for nitroaldol additions (→ 6 ) and for [3 + 2]-dipolar cycloadditions (→ 7 ), diastereoselectively furnishing products with 4 asymmetric C-atoms (not counting acetal centers). The Michael addition described here is yet another example of an ul-combination of trigonal centers induced by Lewis acids, overriding the influence of the configuration of the donor component.  相似文献   

18.
Solid compounds of Cd(II), Hg(II) and Pb(II) with the sodium salt of morin-5′-sulfonic acid (NaMSA) were obtained. The molecular formula of the complexes are: Cd(C15H8O10SNa)2?·?6H2O, CdOH(C15H8O10SNa)?·?4H2O, Hg(C15H8O10S)?·?4H2O and Pb(C15H8O10S)?·?3H2O. Some of their physicochemical properties such as UV-Vis, infrared, 13C NMR and mass spectra, thermogravimetric analysis, and solubility were studied. On the basis of spectroscopic data NaMSA was bound to Cd2+ via 4C=O and 3C?–?oxygen and the Hg2+ and Pb2+ ions by 5C–OH, 4C=O and 3C–OH.  相似文献   

19.
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   

20.
NaSMe in toluene leads to regioselective de-C-silylation of the bis[(trimethylsilyl)ethynyl]saccharide 2 , but to decomposition of butadiynes such as 1 or 12 . We have, therefore, combined the known reagent-controlled, regioselective desilylation of 2 and of 12 (AgNO2/KCN) with a substrate-controlled regioselective de-C-silylation, based on C-silyl groups of different size. This combination was studied with the fully protected 3 which was mono-desilylated to 4 or to 5 (Scheme 1). Triethylsilylation of 5 (→ 6 ) was followed by removal of the Me3Si group (→ 7 ), introduction of a (t-Bu)Me2Si group (→ 8 ) and removal of the Et3Si group yielded 9 ; these high-yielding transformations proceed with a high degree of selectivity. Iodination of 4 gave 10 . The latter was coupled with 5 to the homodimer 11 and the heterodimer 12 , which was desilylated to 13 . The second building block for the tetramer was obtained by coupling 14 (from 7 ) with 5 , leading to 15 and 16 . Removal of the Me3Si group (→ 17 ) and iodination led to 18 which was coupled with 13 to the homotetramer 20 and the heterotetramer 19 (Scheme 2). Deprotection of 19 gave 21 , which was, on the one hand, iodinated to 22 , and, on the other hand, protected by the (t-Bu)Me2Si group (→ 23 ). Removal of the Et3Si group (→ 24 ) and coupling afforded the homooctamer 26 and the heterooctamer 25 . Yields of iodination, silylation, and desilylation were consistently high, while heterocoupling proceeded in only 50–55%. Cleavage of the (i-Pr)3SiC and MeOCH2O groups of 11 (→ 27 ), 15 (→ 28 ), 20 (→ 29 ) and 26 (→ 30 ) proceeded in high yields (Scheme 3). Complete deprotection in two steps of the heterocoupling products 16 (→ 31 → 32 ), 19 (→ 33 → 34 ), and 25 (→ 35 → 36 ) gave the unprotected dimer 32 , tetramer 34 , and octamer 36 in high yields (Scheme 4). Only the dimer 32 is soluble in H2O; the 1H-NMR spectra of 32 , 34 , and 36 in (D6)DMSO (relatively low concentration) show no signs of association.  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