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.
2.
An efficient synthesis of alverine via iron-catalyzed double cross-coupling of (2E)-3-chloro-N-[(2E)-3-chloroprop-2-en-1-yl]-N-ethylprop-2-en-1-amine with phenylmagnesium bromide is described.  相似文献   

3.
Summary 1. A method for the separation and quantitative determination of - and -DNP derivatives of , -DABA, ornithine, and lysine has been proposed.2. A quantitative evaluation of the N N migration in four peptides of , -DABA has been carried out.3. The degree of migration in inactive polymyxin M and the DNP-peptides from a partial hydrolyzate of the inactivated antibiotic has been determined.4. The amino acid composition of the DNP peptide B1 from a partial hydrolyzate of active DNP-polymyxin M has been established.5. The quantitative aspect of the conversion of and diketopiperazines has been studied.Khimiya Prirodnykh Soedinenii, Vol. 2, No. 4, pp. 277–284, 1966  相似文献   

4.
3-Aryl-4-formylsydnones 1a-d react with symmetrical 1,2-dicarbonyl compounds, such as benzil (2a), 4,4′-dimethoxybenzil (2b), 4,4′-difluorobenzil (2c), and di-2-thienylethanedione (2d), in glacial acetic acid, using ammonium acetate as the ammonia source, to yield 4,5-diaryl-2-sydnonyl-substituted imidazoles 3a-6d under conventional heating. In a similar treatment, 4,5-diaryl-2-sydnonyl-1-substituted imidazoles 8a-10a can be prepared by the one-pot condensation of 3-(4-ethoxyphenyl)-4-formylsydnone (1d), benzil derivatives, ammonium acetate, and primary amines. However, such reactions, which take 1-3 days at high temperature under classical conditions, are completed successfully within a few minutes under microwave irradiation.  相似文献   

5.
6.
7.
8.
9.
Syntheses are reported for the novel P-chirogenic bicyclic prophosphatranes P(RNCH2CH2)2N(OCH2CH2)—(8) in which the two R groups [i.e., 1-methylenenaphthyl and 1,2-methoxybenzyl (8)] and in its corresponding phosphine oxide (9) are different. Also synthesized was the transannulated protonated phosphatrane cation in the salt [HP(RNCH2CH2)3N]Cl in which the three R groups [i.e., 1-methylenenaphthyl, 1,2-methoxybenzyl, and (S)-1-phenethyl (14)] are different, and also in its corresponding deprotonated prophosphatrane form P(RNCH2CH2)3N (15). Good analytical resolution of racemic 9 is reported, whereas only partial resolution was achieved for diastereomeric 14.  相似文献   

10.
11.
A facile one-step synthetic protocol toward multifunctionalized cyclopropanes 4 is developed from substituted chalcones 1 and sulfones 2 in good yields via a [2C+1C] annulation.  相似文献   

12.
Conclusions The reaction of p-diethylbenzene and H2S on a chromium-containing oxide catalyst gives a thiophene analog of phenanthrene, namely, benzo[2,1-b3,4-b]dithiophene, and a thiophene analog of anthracene, namely, benzo[1,2-b4,5-b]dithiophene.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 12, pp. 2857–2859, December, 1988.  相似文献   

13.
A concise five-step approach to indolizidinones 10 and 11, two advanced intermediates for the asymmetric synthesis of polyhydroxylated indolizidine alkaloids, has been developed by using N-Cbz pyrrolidin-2-yl pyridin-2-yl sulfide 13 as the chiral building block. The method features a SmI2-mediated coupling of sulfide 13 with functionalized aldehyde 14 and a tandem N-deprotection-lactamization, which constitutes a stepwise “2 + 4” annulation method for the construction of the indolizidinone ring system of 12a.  相似文献   

14.
A seven-step synthesis ofS-(+)-hydroprene (S-1) in 20 % overall yield starting fromS-(+)-3,7-dimethyl-1,6-octadiene (2) of 55+-10 % optical purity is described. The introduction of an optical enhancement step in the synthetic sequence at the stage ofS-(–)-3,7-dimethyl-1-octanol (9) raises the optical purity ofS-1 from 50 % to 80 %.For part 13, see. ref.1 Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 342–348, February, 1993.  相似文献   

15.
This report describes a manganese-catalyzed radical [3 + 2] cyclization of cyclopropanols and oxime ethers, leading to valuable multi-functional 1-pyrrolines. In this redox-neutral process, the oxime ethers function as internal oxidants and H-donors. The reaction involves sequential rupture of C–C, C–H and N–O bonds and proceeds under mild conditions. This intermolecular protocol provides an efficient approach for the synthesis of structurally diverse 1-pyrrolines.

Described herein is a novel manganese-catalyzed radical [3 + 2] cyclization of cyclopropanols and oxime ethers, leading to valuable multi-functionalized 1-pyrrolines.

Pyrroline and its derivatives appear frequently as the core of the structure of natural products and biologically active molecules (Fig. 1A).1 Such compounds also serve as versatile feedstocks in various transformations, such as 1,3-dipolar cyclization, ring opening, reduction and oxidation, leading to diverse and valuable compounds.2–4 Over the past few decades, great effort has been devoted to the preparation of pyrrolines. This has resulted in several elegant approaches that rely on photoredox catalysis (Fig. 1B).5 The groups of Studer,5a,b Leonori,5c and Loh5df disclosed intramolecular addition of the intermediate iminyl radical to alkenes to construct pyrrolines. Generally, the synthetic value of a method can be further improved by using an intermolecular reaction pattern. For example, Alemán et al. recently reported a radical-polar cascade reaction involving the addition to ketimines of alkyl radicals formed in hydrogen atom transfer (HAT) reactions.5g That the existence of benzylic C–H bonds in the substrates is requisite for the HAT, compromises the substrate scope. Despite the appealing photochemical processes, development of new redox approaches to enrich the product diversity of pyrrolines, especially with inexpensive transition-metal catalysts, is still in demand.Open in a separate windowFig. 1(A) Importance of pyrrolines, and (B and C) synthetic approaches to pyrrolines.Prompted by extensive applications of cyclopropanols in synthesis6 and our achievements in manganese-catalyzed ring-opening reactions,7 we conceived a radical [3 + 2] cyclization using cyclopropanol as a C3 synthon and oxime ethers as a nitrogen source (Fig. 1C). Hypothetically, single-electron oxidation of cyclopropanol by Mnn generates the β-keto radical (I), which undergoes a radical [3 + 2] cascade reaction with an oxime ether to give the alkoxy radical species (II). Conversion of II to the intermediate (III), the pyrroline precursor, requires an extra H-donor to support a HAT process and an oxidant for recovery of Mnn to perpetuate the catalytic cycle. In this scenario, the strategic inclusion of oxime ether is crucial to the overall transformation. The oxime ether is not only an internal oxidant and H-donor, but should also be subject to in situ deprotection by cleaving the N–O bond during the reaction. The choice of a proper Mnn/Mnn−1 pair with suitable redox potentials is also vital to the catalytic cycle.Herein, we provide proof-of-principle studies for this hypothesis. The desired radical [3 + 2] cyclization of cyclopropanols and O-benzyl oxime ethers is accomplished with manganese catalysis. This redox-neutral process involves sequential rupture of C–C, C–H, and N–O bonds under mild conditions. The intermolecular protocol provides an ingenious approach to the synthesis of multi-functionalized 1-pyrrolines.With these considerations in mind, phenylcyclopropanol (1a) and oxime ether (2a) were initially chosen as model substrates to evaluate reaction parameters in the presence of manganese salt ( Created by potrace 1.16, written by Peter Selinger 2001-2019 N bond of 2a (entry 2). The optimization of organic solvents was then conducted (entries 3–8), and it was found that the use of fluorinated alcohols, such as trifluoroethanol (TFE) and hexafluoroisopropanol (HFIP) as solvents provided excellent yields (entries 7 and 8). Decreasing the amount of Mn(acac)3 to 1.2 equiv. gave a comparable yield (entry 9), but further reducing the amount compromised the yield (entry 10). Replacing Mn(acac)3 with Mn(OAc)3 or MnCl2 significantly decreased the reaction yield (entries 11 and 12). However, the use of Mn(acac)2 gave a similar yield to Mn(acac)3 (entries 13 vs. 9). The above results prompted us to think over the counteranion effect that the acetylacetone (acac) anion may be requisite to the reaction. Indeed, the synergistic use of stoichiometric MnCl2 and acetylacetone led to a good yield of the desired product (entry 14). More importantly, a comparable yield was obtained with only 0.2 equiv. of MnCl2 and added acetylacetone, realizing this reaction under a catalytic amount of Mn salts (entry 15). Given that the low solubility of the Mn salt may lead to poor efficiency, a reaction with 0.067 M concentration was carried out and gave a 89% yield (entry 16). Further reducing the amount of acetylacetone to 1.0 equiv. had no influence on the outcome of the reaction (entry 17), but the reaction efficiency slightly decreased when 0.6 equiv. of acetylacetone was used as the additive (entry 18). Use of a decreased amount (1.0 equiv.) of acetic acid led to the best yield (91%, entry 19), whereas the reaction in the presence of 0.5 equiv. acetic acid (entry 20) or without acetic acid (entry 21) also gave high yields. It is noted that acetic acid is not crucial to the reaction using MnCl2 as catalyst, as the reaction could generate cat. HCl in situ. The reaction with substoichiometric amount (0.6 equiv.) of acac gave a decreased but also good yield (entry 22). Reducing the catalytic loading of MnCl2 to 10 mol% slightly compromised the yield (entry 23).Optimization of the synthesis of 1-pyrrolines
EntryaMn salt (equiv.)Additive (equiv.)SolventYield (%)
1Mn(acac)3 (1.7)NoneCH3CN33
2bMn(acac)3 (1.7)NoneCH3CNTrace
3Mn(acac)3 (1.7)NoneDCM31
4Mn(acac)3 (1.7)NoneAcetone25
5Mn(acac)3 (1.7)NoneDMSOTrace
6Mn(acac)3 (1.7)NoneDMFTrace
7Mn(acac)3 (1.7)NoneTFE80
8Mn(acac)3 (1.7)NoneHFIP82
9Mn(acac)3 (1.2)NoneHFIP83
10Mn(acac)3 (0.9)NoneHFIP55
11Mn(OAc)3·2H2O (1.2)NoneHFIP36
12MnCl2 (1.2)NoneHFIPTrace
13Mn(acac)2 (1.2)NoneHFIP88
14MnCl2 (1.2)acac (3.6)HFIP80
15MnCl2 (0.2)acac (3.6)HFIP81
16cMnCl2 (0.2)acac (3.6)HFIP89
17cMnCl2 (0.2)acac (1.0)HFIP89
18cMnCl2 (0.2)acac (0.6)HFIP83
19c,dMnCl2 (0.2)acac (1.0)HFIP91
20c,eMnCl2 (0.2)acac (1.0)HFIP83
21c,bMnCl2 (0.2)acac (1.0)HFIP80
22c,dMnCl2 (0.2)acac (0.6)HFIP82
23c,dMnCl2 (0.1)acac (1.0)HFIP81
Open in a separate windowaReaction conditions: 1a (0.45 mmol), 2a (0.3 mmol), AcOH (2.0 equiv.), and Mn salt (as shown) in solvent (2.0 mL), at room temperature (rt) under N2, for 16 h.bWithout AcOH.c0.067 M reaction.d1.0 equiv. AcOH.e0.5 equiv. AcOH. acac = acetylacetone.With the optimized conditions in hand for the synthesis of 1-pyrrolines, the compatibility of various cyclopropanols was inspected (Scheme 1). Common functional groups on the phenyl ring, including halides (3b–3d), ester (3f), ether (3j), were compatible under the reaction conditions. Regardless of the presence of electron-withdrawing or -donating substituents at the para-position of this phenyl ring, the reactions readily proceeded with generally high yields (3b–3j). The cyclopropanol (1k) with an ortho-methyl substituent underwent a cyclization reaction with excellent yield, demonstrating that steric effects had little effect on product of the reaction (3k). By replacing the phenyl group with a naphthyl or thienyl group, the corresponding products (3l and 3m) were produced with slightly lower yields. When 2-substituted cyclopropanols were utilized, these reactions gave rise to a portfolio of trisubstituted 1-pyrrolines (3n–3u).The relative configuration of 3u was determined by comparison with a reported structure.8 Remarkably, this protocol provided a convenient method for the construction of an N-containing spiro skeleton (3t). The reaction with alkyl cyclopropanols could also furnish the desired products (3v–3x) smoothly and with good yields.Open in a separate windowScheme 1Scope of cyclopropanols. Reaction conditions: 1 (0.3 mmol), 2a (0.2 mmol), AcOH (0.2 mmol), MnCl2 (0.04 mmol), and acac (0.2 mmol) in HFIP (3.0 mL), at rt under N2. The d.r. values were determined by 1H NMR analysis with crude reaction mixture, and major isomers are shown with relative configurations. aThe reaction is scaled up for 10 times.Next, we studied the scope of oxime ethers (Scheme 2). Steric hindrance from the ester moiety in the oxime ethers appeared not to influence the reaction outcome. Oxime ethers bearing various esters, such as phenyl (3y), biphenyl (3z and 3ab), 2-naphthyl (3aa), 2,4-di-tert-butylphenyl (3ac and 3ad), and 2,6-dimethylphenyl (3ae) esters all reacted smoothly. In addition, the substrate with tert-butyl ester also readily underwent cyclization to afford the desired product 3af with excellent yield. Remarkably, the trifluoromethyl-substituted pyrroline (3ag) was afforded almost quantitatively from the corresponding ketoxime ether. However, if the trifluoromethyl group was replaced by a methyl or phenyl group, the reaction failed to give rise to the desired product (3ah or 3ai), and this might be attributed to poorer electrophilic nature of the methyl or phenyl substituted substrate.Open in a separate windowScheme 2Scope of oxime ethers. Reaction conditions: 1 (0.3 mmol), 2 (0.2 mmol), AcOH (0.2 mmol), MnCl2 (0.04 mmol), and acac (0.2 mmol) in HFIP (3.0 mL), at rt under N2. The d.r. values were determined by 1H NMR analysis with crude reaction mixture, and major isomers are shown with relative configurations.To illustrate the utility of this protocol, we carried out a set of synthetic applications using 1-pyrroline (3a) (Scheme 3). Upon treatment with acetyl chloride and pyridine at 42 °C, 1-pyrroline (3a) could be readily converted into the acyclic amino acid derivative (4). The reaction between 3a and LiAlH4 gave rise smoothly to the corresponding alcohol (5). In the presence of 2,3-dichloro-5,6-dicyano-1,4-benzoquin-4-one (DDQ) and triethylamine, the 2,5-disubstituted pyrrole (6) was obtained. Moreover, treatment of 3a with MeOTf and NaBH4 delivered the N-methyl proline derivative (7).9Open in a separate windowScheme 3Synthetic applications. Reaction conditions: (a) AcCl, pyridine, dry DCE, 42 °C, 63% yield; (b) LiAlH4, THF, reflux, 90% yield; (c) DDQ, Et3N, DCM, rt, 53% yield; (d) MeOTf, DCM, and then NaBH4, THF, 40% yield, cis : trans = 6.6 : 1.To probe the mechanistic pathways, we performed a radical trapping experiment in the presence of 2.0 equiv. of radical scavenger TEMPO. The radical trapping product (8) was detected by high-resolution mass spectrometry (HRMS) (Scheme 4A, top). In addition, the reaction was obviously suppressed when 1,1-diphenylethylene was added under standard condition (Scheme 4A, bottom). These results suggested that this process engaged in a radical pathway. Kinetic studies illustrated that the reaction immediately started with 20 mol% Mn(acac)2 but an approximate 15 min of induction period was appeared by using Mn(acac)3, which probably indicated that the reaction was initiated with Mn(ii) rather than Mn(iii), and the Mn(ii)/Mn(i) cycle might be involved in the transformation (Scheme 4B, for details see ESI).Open in a separate windowScheme 4(A and B) Mechanistic studies, and (C) proposed mechanism.On the basis of these results, a plausible mechanism for this radical process was proposed in Scheme 4C. Initially, the interaction between cyclopropanol (1a) and Mn(ii) salt gives rise to the alkoxy manganese species (I), which undergoes a ligand-to-metal charge transfer (LMCT) process, leading to the alkoxy radical (II).5f Subsequent ring-opening of the alkoxyl radical (II) provides the alkyl radical (III). The addition of intermediate (III) to the oxime ether, possibly activated by HFIP or HOAc, furnishes the N-centered radical (IV), which intramolecularly attacks the ketone to afford a new alkoxy radical (V).10 The subsequent 1,5-hydrogen atom transfer (HAT) process delivers the alkyl radical (VI) at the α-position adjacent to the O atom, thus driving N–O bond cleavage to generate the N-centered radical (VII),5b,11 and benzaldehyde which was detected by TLC. This radical intermediate (VII) undergoes a single electron transfer (SET) mediated by the reduced-state Mn(i) species, and protonation to yield the cyclic pyrrolidine (VIII). Finally, dehydration of this intermediate produces 1-pyrroline (3a).  相似文献   

16.
Derivatives of a new heterocyclic system thieno[2,35,6]pyrimido[3,4-a]indole, were obtained by the reaction of substituted N,O-diacetylindoxyls with excess N-(2-methyl-3-ethoxycarbonyl-4-thienyl)hydrazine. The reaction of N-acetylindoxyl and 4-hydrazinouracil forms 12-amino-1,3-dioxo-2,4,6-trimethylpyrimido[5,45,6]pyrimido[3,4-a]indole.Deceased.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 10, 1343–1345, October, 1987.  相似文献   

17.
Reaction of 2-chlorophenothiazine with acryloyl and methacryloyl chloride, and dehydrochlorination of 10-(-chloropropionyl)-2-chlorophenothiazine, has given 10-acryloyl- and 10-methacryloyl-2-chlorophenothiazine. The latter reacts with secondary amines to give chloracisin and its analogs.Translated from Khimiya Geterotsiklicheskikh Soedinenii, Vol. 6, No. 5, pp. 601–604, May, 1970.  相似文献   

18.
The reaction of ammonia or a primary amine with alkyl(cycloaklyl, phenyl)-,-dihalopropyl ketones gives 2-. 2,4-, and 1,2- or 1.2,4-substituted pyrroles in high yields.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 6, pp. 790–793, June, 1976.  相似文献   

19.
20.
C2C4 olefins were synthesized from methanol on modified ZSM-34 type zeolite catalyst at 400°C. Methanol vapor of 12 vol.% was converted completely under flow conditions at 800 h–1 space velocity, and in molar selectivity 42.9% C2H4, 33.4% C3H6, and 3.3% C4H8 were produced.
C2C4 ZSM-34 400°C. 12 .%-, 800 h–1 : C2H4-42,9%, C3H6-33,4% C4H8-3,3%.
  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