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 750 毫秒
1.
Boc-D -Ala-MeLeu-MeLeu-MeVal-OH (DLLL ) and its isomers DLDL, DLDD and DLLD were synthesized using several different strategic approaches and a modification of the mixed pivalic anhydride method for carboxyl activation. Alternatively, the tert-butoxy-carbonyl (Boc) or benzyloxycarbonyl (Z) amino-protecting groups and the benzyloxy (OBzl) or tert-butoxy (OtBu) carboxyl-protecting groups were used to protect the reacting peptides. By monitoring the reaction temperature, it was possible to synthesize, starting from the tripeptide DLL as peptide model, either the tetrapeptide DLLL (?20°) or the tetrapeptide DLDL (+20°), selectively. Using 1H-NMR spectroscopy to follow the mixed pivalic anhydride formation of the DLL - and DLD -tripeptides, it could be shown that anhydride formation is strongly dependent on the temperature. It is slow at ?20° (several hours) and fast at +20° (20 to 40min). The isomerization of the DLL -anhydride to the more stable DLD -anhydride can be reduced to a minimum by working at ?20°, while this isomerization proceeds to near completion at +20°.  相似文献   

2.
The title compounds, 4 and 7 , have been prepared from the corresponding α‐amino acid derivative selenocystine ( 1 ) by the following sequence of steps: cleavage of the Se? Se bond with NaBH4, p‐methoxybenzyl (PMB) protection of the SeH group, Fmoc or Boc protection at the N‐atom and Arndt–Eistert homologation (Schemes 1 and 2). A β3‐heptapeptide 8 with an N‐terminal β3‐hSec(PMB) residue was synthesized on Rink amide AM resin and deprotected (‘in air’) to give the corresponding diselenide 9 , which, in turn, was coupled with a β3‐tetrapeptide thiol ester 10 by a seleno‐ligation. The product β3‐undecapeptide was identified as its diselenide and its mixed selenosulfide with thiophenol (Scheme 3). The differences between α‐ and β‐Sec derivatives are discussed.  相似文献   

3.
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   

4.
The reaction of methyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate ( 2a ) with thiobenzoic acid at room temperature gave the endothiopeptide Bz‐AibΨ[CS]‐Pro‐OMe ( 7 ) in high yield. In an analogous manner, (benzyloxy)carbonyl (Z)‐protected proline was transformed into the thioacid, which was reacted with 2a to give the endothiotripeptide Z‐Pro‐AibΨ[CS]‐Pro‐OMe ( 12 ). The corresponding thioacid of 7 was prepared in situ via saponification, formation of a mixed anhydride, and treatment with H2S. A second reaction with 2a led to the endodithiotetrapeptide 9 , but extensive epimerization at Pro2 was observed. Similarly, saponification of 12 and coupling with either 2a or H‐Phe‐OMe and 2‐(1H‐benzotriazol‐1‐yl)‐1,1,3,3‐tetramethyluronium tetrafluoroborate/1‐hydroxy‐1H‐benzotriazole (TBTU/HOBt) gave the corresponding endothiopeptides as mixtures of two epimers. The synthesis of the pure diastereoisomer BzΨ[CS]‐Aib‐Pro‐AibΨ[CS]‐N(Me)Ph ( 21 ) was achieved via isomerization of 7 to BzΨ[CS]‐Aib‐Pro‐OMe ( 16 ), transformation into the corresponding thioacid, and reaction with N,2,2‐trimethyl‐N‐phenyl‐2H‐azirin‐3‐amine ( 1a ). The structures of 12 and 21 were established by X‐ray crystallography.  相似文献   

5.
Addition of various amines to the 3,3‐bis(trifluoromethyl)acrylamides 10a and 10b gave the tripeptides 11a – 11f , mostly as mixtures of epimers (Scheme 3). The crystalline tripeptide 11f 2 was found to be the N‐terminal (2‐hydroxyethoxy)‐substituted (R,S,S)‐ester HOCH2CH2O‐D ‐Val(F6)‐MeLeu‐Ala‐OtBu by X‐ray crystallography. The C‐terminal‐protected tripeptide 11f 2 was condensed with the N‐terminus octapeptide 2b to the depsipeptide 12a which was thermally rearranged to the undecapeptide 13a (Scheme 4). The condensation of the epimeric tripeptide 11f 1 with the octapeptide 2b gave the undecapeptide 13b directly. The undecapeptides 13a and 13b were fully deprotected and cyclized to the [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐D ‐valine]]‐ and [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐L ‐valine]]cyclosporins 14a and 14b , respectively (Scheme 5). Rate differences observed for the thermal rearrangements of 12a to 13a and of 12b to 13b are discussed.  相似文献   

6.
Twenty nine novel N‐4‐methyl‐1,2,3‐thiadiazole‐5‐carbonyl‐N′‐phenyl ureas were designed and synthesized, and their structures were confirmed by proton nuclear magnetic resonance (1H NMR), infra red spectroscopy (IR) and high‐resolution mass spectroscopy (HRMS). Compounds V‐9 , V‐11 , V‐12 , V‐15 , V‐19 , V‐21 , V‐22 and V‐24 exhibit excellent activity against Culex pipiens pallens. Compounds V‐12 and V‐22 present good insecticidal activity against Plutella xylostella L. Their median lethal concentrations (LC50) are 164.15 and 89.69 mg·L?1, respectively. Compound V‐11 also has potential wide spectrum of fungicide activity. Its median effective concentrations (EC50) detected from 3.82 µg·mL?1 against Physalospora piricola to 31.60 µg·mL?1 against Cercospora arachidicola. Compounds V‐15 and V‐24 show outstanding induction activities as same as positive controls TDL and ningnanmycin, furthermore V‐24 has the highest induction activity of 41.85%±4.43%. To elucidate the structure activity relationship in these compounds, a 3D‐QSAR model has been built. The established model showed a reliable predicting ability with q2 values of 0.643 and r2 values of 0.982.  相似文献   

7.
用DCC/DMAP合成N-苄氧羰基氨基酸薄荷酯   总被引:12,自引:0,他引:12  
李前荣  顾承志  尹浩  张毅 《有机化学》2005,25(11):1416-1419
以乙腈为溶剂, 用DCC/DMAP偶合法合成了5个N-苄氧羰基氨基酸薄荷酯, 产率90%~94%. 起始原料氨基酸包括: 甘氨酸, DL-蛋氨酸, β-丙氨酸, L-缬氨酸和L-脯氨酸. 给出了目标产物的氢谱、碳谱、红外光谱、质谱、高分辨质谱和元素分析数据.  相似文献   

8.
Polycyclic ‘cage’ ketones, such as pentacyclo[5.4.0.02,6.03,10.05,9]undecan‐8‐one ( 10 ), pentacyclo[5.4.0.02,6.03,10.05,9]undecane‐8,11‐dione ( 11 ), and adamantan‐2‐one ( 16 ) were treated with the nucleophilic dimethoxycarbene (DMC; 1 ), which was generated thermally from 2,5‐dihydro‐2,2‐dimethoxy‐5,5‐dimethyl‐1,3,4‐oxadiazole ( 4a ) in boiling toluene. In this ‘one‐pot’ procedure, the α‐hydroxycarboxylic acid ester 12 or a corresponding derivative 15 or 17 was obtained (Schemes 4–7). Additionally, ‘cage’ thione 21 was treated with DMC under the same conditions yielding dimethoxythiirane 22 (Scheme 8). Subsequent hydrolysis or desulfurization (followed by hydrolysis on silica gel) of 22 gave α‐mercaptocarboxylate 25 and the corresponding desulfurized ester 24 , respectively. In all cases, the addition of DMC occurred stereoselectively, and the addition from the exo‐face is postulated to explain the structures of the isolated products.  相似文献   

9.
(−)-D -Noviose ((−)- 22 ), a rare sugar, was synthesized starting from the optically active building block (−)- 1 in seven steps. The first step of this route, the introduction of a methyl-ether group under Lewis-acidic conditions, left the acetoxy group untouched and thereby preserved the absolute configuration in the product 2 (Scheme 2). Next, the double bond of methyl ether 2 was cis-dihydroxylated leading selectively to 3 . After saponification of the acetoxy group of 3 the two vicinal cis OH groups of 17 were selectively protected as the cyclic carbonate 18 (Scheme 4). This kind of protection was essential to achieve the proper regioselectivity in the Baeyer-Villiger rearrangement of the cyclopentanone derivative 19 that was obtained after oxidation of the remaining OH group of 18 . Lactone 20 was the major product of this rearrangement. Final reduction to the corresponding lactol (cyclic hemiacetal) with diisobutylaluminium hydride (DIBAH) at low temperature was accompanied by reductive cleavage of the protecting cyclic carbonate moiety thereby leading directly to (−)-D -noviose ((−)- 22 ).  相似文献   

10.
The protected poly‐Aib oligopeptides Z‐(Aib)n‐N(Me)Ph with n=2–6 were prepared according to the ‘azirine/oxazolone method’, i.e., by coupling amino or peptide acids with 2,2,N‐trimethyl‐N‐phenyl‐2H‐azirin‐3‐amine ( 1a ) as an Aib synthon (Scheme 2). Following the same concept, the segments Z‐(Aib)3‐OH ( 9 ) and H‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 20 ) were synthesized, and their subsequent coupling with N,N′‐dicyclohexylcarbodiimide (DCC)/ZnCl2 led to the protected heptapeptide Z‐(Aib)3‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 21 ; Scheme 3). The crystal structures of the poly‐Aib oligopeptide amides were established by X‐ray crystallography confirming the 310‐helical conformation of Aib peptides.  相似文献   

11.
The reaction of N‐alkylbenzaldimines with carbon monoxide (CO) in the presence of cobalt (Co) catalysts resulted in the formation of N‐alkylphthalimidines (Table 1). Their formation is proposed to occur by C? H activation of the aryl ring, migratory insertion of the hydride species into the benzaldimine functionality, CO coordination, and insertion into the Co? C bond, followed by reductive elimination of the N‐alkylphthalimidine and regeneration of the starting Co species (Scheme 4). Deuterium (2H)‐labeling NMR studies are consistent with this mechanism (Scheme 5).  相似文献   

12.
A novel manganese(III)–porphyrin complex, namely, catena‐poly[[chloridomanganese(III)]‐μ2‐5,10,15,20‐tetrakis(pyridin‐3‐yl)‐21H,23H‐porphinato(2?)‐κ5N21,N22,N23,N24:N5], [MnCl(C40H24N8)]n, 1 , was prepared by the hydrothermal reaction of manganese chloride with 5,10,15,20‐tetrakis(pyridin‐3‐yl)‐21H,23H‐porphine. The crystal structure was determined by single‐crystal X‐ray diffraction. The porphyrin macrocycle exhibits a saddle‐like distortion geometry. The MnIII atom has a six‐coordination geometry. Each porphyrin unit links to two neighbouring units to yield a one‐dimensional coordination polymer. These chains are further interlinked by hydrogen bonds to form a two‐dimensional network. The complex shows red photoluminescence emission bands in ethanol solution, which can be attributed to ligand‐to‐ligand charge transfer (LLCT) accompanied by partial metal‐to‐ligand charge transfer (MLCT), as revealed by TDDFT calculations.  相似文献   

13.
First Example of an H-Shift in ‘Thiocarbonyl Aminides’ (N-(Alkylidenesulfonio)aminides) Reaction of benzyl azide ( 15a ) with the sterically hindered C?S group of 4,4-dimethyl-1,3-thiazole-5(4H)-thiones 14 (Scheme 3) and 1,1,3,3-tetramethylindane-2-thione ( 17 , Scheme 4) at 80° leads to the corresponding imines in high yield, without formation of any by-product. In contrast, 15a and 2,2,4,4-tetramethyl-3-thioxocyclobutanone ( 7 ) under the same conditions yielded, in addition to imine 19 , products 20a and 21 (Scheme 5). For the formation of 20a , a reaction mechanism via [1,4]-H shift in the intermediate ‘thiocarbonyl aminides’ 23 is proposed (Scheme 6). Product 21 as well as the dithiazole derivative 22 , which is formed only in the reaction with 4-nitrobenzyl azide ( 15c ), are formal adducts of the dipole 23 . Whereas precedents are known for the formation of cycloadducts of type 22 , the pathway to 21 is not known. Two possible mechanisms of its formation are proposed in Schemes 8 and 9.  相似文献   

14.
Bromination of the monosilylated dialkynylated monomer 1a (Scheme 1), dimer 3 and tetramer 5 by N-bromosuccinimide (NBS) in the presence of CF3COOAg gave 2, 4 , and 6 , respectively, in over 93%. Similar conditions led to bromodesilylation. Either silyl group of the diprotected monomer 1c was selectively removed by bromolysis. On the one hand, bromodesilylation of 1c gave 2 in yields varying between 80 and 99%. On the other hand, bromodesilylation of 7 , obtained from 1c by hydrolytic removal of the tetrahydro-2H-pyran-2-yl (Thp) group, yielded 91% of 8 . Mechanistic considerations suggested that the deprotective bromination should be improved by replacing the Me3Si by a Me3Ge group. Indeed, bromodegermylation of 1b was quantitative and ca. 60 times faster than bromodesilylation of 1c . The Me3Si and Me3Ge groups can be used for an orthogonal protection/activation of dialkynes. This was shown by desilylating 12 to 11 (Scheme 2), while bromination yielded 13 . Both reactions proceeded in high yields; 9 was isolated as a minor by-product of 13 . The reactivity towards bromolysis decreases in the series H-DOPS > Me3Ge ≈ H > Me3Si > Thp-DOPS (DOPS = [dimethyl(oxy)-p ropyl]dimethylsilyl). Orthogonal bromolysis of DOPS- and Me3Ge-substituted dialkynes is slightly more selective than the one of Me3Si- and Me3Ge-substituted analogues. Coupling of 7 with the bromoalkyne 2 gave the dimer 15 (76%), 14 (2%), and 16 (4%) (Scheme 3). The binomial synthesis was optimized so that each cycle, doubling the size of the precursor, requires the minimal number of transformations (Scheme 4). The orthogonally protected monomer 1b , dimer 19 , and tetramer 22 were, on the one hand, hydrolyzed to the alcohols 18 (95%), 21 (91%), and 24 (91%), respectively, and, on the other hand, bromodegermylated to 2 (99%), 4 (97%), and 6 (93%). Cross-coupling of 18 with 2, 21 with 4 , and 24 with 6 gave the orthogonally protected dimer 19 (73%), tetramer 22 (87%), and octamer 25 (83%), respectively.  相似文献   

15.
The cycloadducts of isoquinolinium N‐phenyl imide 2 with C=C bonds are derivatives of 2‐amino‐1,2‐dihydroisoquinoline. Their Nβ‐vinylphenylhydrazine system is amenable to an acid‐catalyzed [3,3]‐sigmatropic shift; the formation of pentacyclic aminals is exemplified by 6 → 8 . The dimethyl maleate adduct 11 , C21H20N2O4, is exceptional by being converted on treatment with acid to bright‐yellow crystals, C24H22N2O6 (additional C3H2O2). X‐Ray crystal‐structure analysis and NMR spectra reveal structure 13 , and mechanistic studies indicated an initial β‐elimination at the N−N bond of 11 to yield 18 ; this step is followed by a retro‐Mannich‐type cleavage that gives methyl isoquinoline‐1‐acetate ( 14 ) and methyl 2‐(phenylimino)acetate ( 15 ), according to the sequence C21H20N2O4 ( 11 )→ 18 →C12H11NO2 ( 14 )+C9H9NO2 ( 15 ). In the second act of the drama, electrophilic attack by 15 ‐H+ on the ene‐hydrazine group of a second molecule of 11 furnishes 13 by a polystep intramolecular redox reaction. All rate constants must be fine‐tuned in this reaction cascade to give 13 in yields of up to 78% with an overall stoichiometry: 2 C21H20N2O4 ( 11 )→C24H22N2O6 ( 13 )+C12H11NO2 ( 14 )+aniline. Interception and model experiments confirmed the above pathway. A by‐product, C33H31N3O6 ( 62 ), arises from an acid‐catalyzed dimerization of 11 and subsequent elimination of 15 .  相似文献   

16.
Alkylbenzenes are synthesized for the first time from aliphatic hydrocarbons via an one pot, transition metal‐free coupling approach under basic conditions. The method consists of two steps: condensation of 2‐bromoprop‐2‐enyl‐ or 2‐propargylcyclohexanone with alcohols, amines, or amino alcohols, followed by base treatment (Scheme 1). Phenolic ethers and N‐phenylated polyalkyl aromatic compounds are shown to be in the scope of the demonstrated reaction (Table). The proposed mechanism suggests that the unsaturation in another part of the molecule (propargyl‐group equivalent) is transferred into the cyclohexane ring to yield a benzene ring through a series of prototropic shifts.  相似文献   

17.
4μ‐A2B2 star‐shaped copolymers contained polystyrene (PS), poly(isoprene) (PI), poly(ethylene oxide) (PEO) or poly(ε‐caprolactone) (PCL) arms were synthesized by a combination of Glaser coupling with living anionic polymerization (LAP) and ring‐opening polymerization (ROP). Firstly, the functionalized PS or PI with an alkyne group and a protected hydroxyl group at the same end were synthesized by LAP and then modified by propargyl bromide. Subsequently, the macro‐initiator PS or PI with two active hydroxyl groups at the junction point were synthesized by Glaser coupling in the presence of pyridine/CuBr/N,N,N ′,N ″,N ″‐penta‐methyl diethylenetri‐amine (PMDETA) system and followed by hydrolysis of protected hydroxyl groups. Finally, the ROP of EO and ε‐CL monomers was carried out using diphenylmethyl potassium (DPMK) and tin(II)‐bis(2‐ethylhexanoate) (Sn(Oct)2) as catalyst for target star‐shaped copolymers, respectively. These copolymers and their intermediates were well characterized by SEC, 1H NMR, MALDI‐TOF mass spectra and FT‐IR in details. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

18.
We have accomplished the synthesis of 13C‐labeled tautomycin at the C18, C19, C21, and C22 positions starting from 100 % [13C]triethylphosphonoacetate for the purpose of elucidating the dynamics and conformation of the C17–C26 moiety. NMR spectroscopy of 13C‐labeled tautomycin revealed strong binding with protein phosphatase type 1 and new features in the 13C NMR spectrum, such as the very small three‐bond coupling constants (2J).  相似文献   

19.
Attempted Synthesis of Push-Pull Diacetylenes Two alternative synthesis of push-pull diacetylenes of type 1 (5-amino-2,4-alkadiynals) are investigated. A bromination-dehydrobromination sequence starting with 5-dimethylamino-2,4-pentadienal ( 2 ) as well as the application of the well-known Cadiot-Chodkiewicz coupling reaction give new intermediates 3–5 , and 7 and 8 , respectively, but fail to give the target molecules 1 .  相似文献   

20.
Various condensed areno[g]lumazine derivatives 2 , 3 , and 5 – 7 were synthesized as new fluorescent aglycones for glycosylation reactions with 2-deoxy-3, 5-di-O-(p-toluoyl)-α/β-D -erythro-pentofuranosyl chloride ( 10 ) to form, in a Hilbert-Johnson-Birkofer reaction, the corresponding N1-(2′-deoxyribonucleosides) 15 – 21 . The β-D -anomers 15 , 17 , 19 , and 21 were deblocked to 24 – 27 and, together with N1-(2′-deoxy-β-D -ribofuranosyl)lumazine ( 22 ) and its 6, 7-diphenyl derivative 23 , dimethoxytritylated in 5′-position to 28–33. These intermediates were then converted into the 3′-(2-cyanoethyI diisopropylphosphoramidites) 34 – 39 which function as monomeric building block in oligonucleotide syntheses as well as into the 3′-(hydrogen succinates) 40 – 45 which can be used for coupling with the solid-support material. A series of lumazine-modified oligonucleotides were synthesized and the influence of the new nucleobases on the stability of duplex formation studied by measuring the Tm values in comparison to model sequences. A substantial increase in the Tm is observed on introduction of areno[g]lumazine moieties in the oligonucleotide chain stabilizing obviously the helical structures by improved stacking effects. Stabilization is strongly dependent on the site of the modified nucleobase in the chain.  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