首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The reactions of thiobenzamide 8 with diazo compounds proceeded via reactive thiocarbonyl ylides as intermediates, which underwent either a 1,5‐dipolar electrocyclization to give the corresponding five membered heterocycles, i.e., 4‐amino‐4,5‐dihydro‐1,3‐thiazole derivatives (i.e., 10a, 10b, 10c , cis‐ 10d , and trans‐ 10d ) or a 1,3‐dipolar electrocyclization to give the corresponding thiiranes as intermediates, which underwent a SNi′‐like ring opening and subsequent 5‐exo‐trig cyclization to yield the isomeric 2‐amino‐2,5‐dihydro‐1,3‐thiazole derivatives (i.e., 11a, 11b, 11c , cis‐ 11d , and trans‐ 11d ). In general, isomer 10 was formed in higher yield than isomer 11 . In the case of the reaction of 8 with diazo(phenyl)methane ( 3d ), a mixture of two pairs of diastereoisomers was formed, of which two, namely cis‐ 10d and trans‐ 10d , could be isolated as pure compounds. The isomers cis‐ 11d and trans‐ 11d remained as a mixture. In the reactions of the thioxohydrazone 9 with diazo compounds 3b and 3d , the main products were the alkenes 18 and 23 , respectively. Their formation was rationalized by a 1,3‐dipolar electrocyclization of the corresponding thiocarbonyl ylide and subsequent desulfurization of the intermediate thiiran. As minor products, 2,5‐dihydro‐1,3‐thiazol‐5‐amines 21 and 24 were obtained, which have been formed by 1,5‐dipolar electrocyclization of the thiocarbonyl ylide, followed by a 1,3‐shift of the dimethylamino group.  相似文献   

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

3.
The ability to incorporate alkyl, alkenyl, aryl, and heteroatomic groups into substrates for the silicon-directed Nazarov cyclization and their subsequent reactions has been investigated. In general, most of the groups are compatible with the conditions for the cyclization and do not interfere even when directly attached to the divinyl ketone. The influence of substituents on the rate of the cyclization has been addressed and is consistent with a simple mechanistic picture. O- and N-Containing functions are tolerated except when attached to the α-vinyl C-atom of the divinyl ketone. The diastereoface-directing effect of a fused cyclobutane is discussed.  相似文献   

4.
The self‐assembled supramolecular host [Ga4L6]12? ( 1 ; L=N,N‐bis(2,3‐dihydroxybenzoyl)‐1,5‐diaminonaphthalene) catalyzes the Nazarov cyclization of 1,3‐pentadienols with extremely high levels of efficiency. The catalyzed reaction proceeds at a rate over a million times faster than that of the background reaction, an increase comparable to those observed in some enzymatic systems. A detailed study was conducted to elucidate the reaction mechanism of both the catalyzed and uncatalyzed Nazarov cyclization of pentadienols. Kinetic analysis and 18O‐exchange experiments implicate a mechanism, in which encapsulation, protonation, and water loss from substrate are reversible, followed by irreversible electrocyclization. Although electrocyclization is rate determining in the uncatalyzed reaction, the barrier for water loss and for electrocyclization are nearly equal in the assembly‐catalyzed reaction. Analysis of the energetics of the catalyzed and uncatalyzed reaction revealed that transition‐state stabilization contributes significantly to the dramatically enhanced rate of the catalyzed reaction.  相似文献   

5.
The intramolecular addition of unsaturated alkoxycarbenes leads in high yields and diastereoselectively to fused cyclopropanes (Scheme 1). Reaction of the halodiazirines 2 , 10 , 11 , and 20 with the unsaturated phenolates 1 , 8 , and 9 yielded intermediate alkoxydiazirines, and hence the homobenzofurans 5 , 12 – 16 , 22 , and 26 (Scheme 2). The intermediate alkoxydiazirine 25 was isolated at low temperature (Scheme 3). An equilibrium between the cyclopropane derivatives 12 and 27 , and 14 and 28 was established at 120°. At 200°, 12 rearranged to the chromene 29 , by disrotarory opening of the cyclopropane ring, followed by electrocyclization. Hydrogenation of 29 gave the (all-cis)-chroman 32 (Scheme 4). The homoindole 35 was obtained in good yields, presumably by an SRN1 reaction from 34 and 10 (Scheme 5).  相似文献   

6.
Indirubin, present in extracts of Isatis tinctoria and some other plant species, has promising cytotoxicity against a variety of cell lines by inhibition of cyclin‐dependent kinases. Chemical synthesis of its derivatives relies on the combination of isatins and 2,3‐dihydro‐1H‐indol‐3‐one (‘indoxyl’) derivatives and usually yields indigo as well as other by‐products. Inspection of the hydrolysis of the long‐known condensation products of 2‐thioxothiazolidin‐4‐one with isatins gave useful hints for an improved synthesis of indirubins: this reaction does not yield quinoline derivatives but 2‐(2,3‐dihydro‐2‐oxo‐1H‐indol‐3‐ylidene)‐2‐sulfanyl acetic acids. By substitution of the sulfanyl group in this oxindoles with anilines and straightforward cyclization under Nazarov conditions, a broad variety of indirubins substituted in the indoxyl ring system are thus available, usually in very good purity and yield. Use of naphthylamines in this reaction sequence yields various fluorescent substances with λfl at ca. 630 nm.  相似文献   

7.
Synthesis of (±)-α-Chamigrene Cis- and trans-β-ionol (cis and trans- 1 ) underwent acid catalysed dehydration to a mixture of the tetraenes 2–5 in 70–80% yield (Table 1). Irradiation of this mixtures made the 6-(Z), 8-(Z)-isomer 5 accessible (columns 3 and 4 in Table 1). Pyrolysis of the different mixtures at 170° showed, that both isomers, 3 and 5 respectively undergo electrocyclization to dehydrochamigrene ( 6 ). The latter was reduced to α-chamigrene ( 7 ) by hydrogen on Lindlar catalyst.  相似文献   

8.

The oxo-bridged dinuclear complexes [(μ-O){ReOCl2(L)}2] [L = 2-(1-ethylaminomethyl)-1-methylimidazole (eami); 2-(1-methylaminomethyl)-1-methylimidazole (mami); 2-(1-ethylthiomethyl)-1-methylimidazole (etmi)] were prepared by reaction of trans-[ReOCl3(PPh3)2] with L in acetone. X-ray crystallographic studies of the eami and etmi complexes show that these ligands coordinate in a bidentate manner, and that the cis, cis-N2Cl2 and cis, cis-NSCl2 equatorial planes are nearly orthogonal to the O=Re-O-Re=O backbone.  相似文献   

9.
The kinetics of the di- and trimerization of ethylen in organic solvents under the influence of a homogeneous catalyst containing π-tetramethylcyclobutadiene-nickeldichloride and a prereacted mixture of ethylaluminiumdichloride and tri-n-butylphosphine are reported. The primary reaction product is 1-butene, which is isomerized to 2-butene (cis/trans) during the reaction. The C6-Olefins are formed by the reaction of ethylene with 1-butene and with the 2-butenes. The following primary reaction products are obtained: 3-hexene (cis/trans), 1-hexene, 2-ethyl-1-butene, 3-methyl-1-pentene and 3-methyl-2-pentene (cis/trans). The effect of other phosphines on the reaction was also studied. The relative composition of the reaction product is strongly dependent upon the amount and the LEWIS base strength of the phosphine present. The results are in accordance with a coordinative mechanism on nickel.  相似文献   

10.
A versatile reactivity from the cage compound P(NMeNMe)3P is presented. The Staudinger reaction with Me3SiN3 is carried out. The crystal structure of the compound issued from the reaction on both sides (Me3SiN=P(NMeNMe)3P=NSiMe3) is reported. When the reaction occurs on only one side, the remaining free phosphorus atom is complexed with RuCl2(p‐cymene). P(NMeNMe)3P reacts with PCl3, leading to the heterocyclic compound ClP(NMeNMe)2PCl. This heterocycle also displays a versatile reactivity. Substitution reaction with HNiPr2 leads to iPr2NP(NMeNMe)2PNiPr2. Very complex 1H and 13C NMR spectra suggest that the cis isomer is the largely major isomer of this compound. The cis structure is confirmed by X‐ray diffraction. Besides the reaction on the P–Cl functions, the reaction on the lone pair of ClP(NMeNMe)2PCl is carried out, leading to the complex (p‐cymene)Cl2RuPCl(NMeNMe)2ClPRuCl2(p‐cymene). Characterization of this compound by X‐ray diffraction displays a cis isomer for this compound also.  相似文献   

11.
Potential-energy surfaces for various channels of the HNO+NO2 reaction have been studied at the G2M(RCC,MP2) level. The calculations show that direct hydrogen abstraction leading to the NO+cis-HONO products should be the most significant reaction mechanism. Based on TST calculations of the rate constant, this channel is predicted to have an activation energy of 6–7 kcal/mol and an A factor of ca. 10−11 cm3 molecule−1 s−1 at ambient temperature. Direct H-abstraction giving NO+trans-HONO has a high barrier on PES and the formation of trans-HONO would rather occur by the addition/1,3-H shift mechanism via the HN(O)NO2 intermediate or by the secondary isomerization of cis-HONO. The formation of NO+HNO2 can take place by direct hydrogen transfer with the barrier of ca. 3 kcal/mol higher than that for the NO+cis-HONO channel. The formation of HNO2 by oxygen abstraction is predicted to be the least significant reaction channel. The rate constant calculated in the temperature range 300–5000 K for the lowest energy path producing NO+cis-HONO gave rise to © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 729–736, 1998  相似文献   

12.
A simple and green method that uses [Ru(Me3tacn)Cl3] ( 1 ; Me3tacn=N,N′,N′′‐trimethyl‐1,4,7‐triazacyclononane) as catalyst, aqueous H2O2 as the terminal oxidant, and Al2O3 and NaCl as additives is effective in the cis‐dihydroxylation of alkenes in aqueous tert‐butanol. Unfunctionalized alkenes, including cycloalkenes, aliphatic alkenes, and styrenes (14 examples) were selectively oxidized to their corresponding cis‐diols in good to excellent yield (70–96 %) based on substrate conversions of up to 100 %. The preparation of cis‐1,2‐cycloheptanediol (119 g, 91 % yield) and cis‐1,2‐cyclooctanediol (128 g, 92 % yield) from cycloheptene and cyclooctene, respectively, on the 1‐mol scale can be achieved by scaling up the reaction without modification. Results from Hammett correlation studies on the competitive oxidation of para‐substituted styrenes (ρ=?0.97, R=0.988) and the detection of the cycloadduct [(Me3tacn)ClRuHO2(C8H14)]+ by ESI‐MS for the 1 ‐catalyzed oxidation of cyclooctene to cis‐1,2‐cyclooctanediol are similar to those of the stoichiometric oxidation of alkenes by cis‐[(Me3tacn)(CF3CO2)RuVIO2]+ through [3+2] cycloaddition (W.‐P. Yip, W.‐Y. Yu, N. Zhu, C.‐M. Che, J. Am. Chem. Soc. 2005 , 127, 14239).  相似文献   

13.
Three metal-organic frameworks (MOFs) [La(bta)(H2O)]n (1), [La(bta)(H2O)2·H2O]n (2) and [Er(bta)(H2O)2·H2O]n (3) with different structures were synthesized by reactions of 1,3,5-benzenetriacetic acid (H3bta) with the corresponding lanthanide salts under different conditions. X-ray diffraction analyses reveal that complexes 1 and 2 have the same metal atom and ligand but different structures. In 1, each bta3 ?  adopts cis, cis, cis conformation and acts as a μ6-bridge linking six lanthanum atoms to form a 2D framework; while in 2, the bta3 ?  has cis, trans, trans conformation and serves as a μ5-bridging ligand, which results in a 3D channel-like structure. In the case of 3, each bta3 ?  also adopts cis, trans, trans conformation but acts as a μ4-bridge linking four erbium atoms to form a 3D channel-like framework. The results imply that the reaction conditions have great impact on the structure of MOFs, and the flexible triacid ligand H3bta is versatile and can adopt different conformations in the formation of MOFs. The magnetic property of complex 3 was investigated in the temperature range of 1.8–300 K.  相似文献   

14.
Schiff bases derived from the condensation of β-diketones with N-methyl-S-methyldithiocarbazates yield cis dicarbonyl complexes Rh(CO)2 (Schiff) on reaction with [Rh(μ-Cl)(CO)2]2. Those derived from aromatic aldehydes form trans dicarbonyl complexes. These complexes with excess of triphenylphosphine give only Rh(CO)(PPh3)(Schiff). cis-1,5-cyclooctadiene (COD) reacts with cis dicarbonyl complexes to yield the carbonyl-free product Rh(COD)(Schiff); similar reactions have not been observed in the case of trans-dicarbonyl complexes. Oxidative addition of bromine to these complexes yields dibromo derivative in which the Schiff base acts as bidentate chelate. Rh(PPh3)2(Schiff) complexes have been obtained from the reaction of above Schiff bases with Rh(PPh3)3Cl. The structures of these new complexes have been determined based on IR and 1H NMR spectra.  相似文献   

15.
In chloroform, [ZrCl4·2(MeO)3PO] exists in both cis- and trans-isomeric forms. Three reactions can be envisaged in the presence of excess (MeO)3PO = L: (1) cis-[ZrCl4·2L] + *L?cis-[ZrCl4·L*L]+ L; (2) trans-[ZrCl4·2L] + *L ? trans-[ZrCl4·L*L] + L; (3) cis-[ZrCl4·2L]? trans-[ZrCl4·2L]. To distinguish between these possible reaction pathways, we have used 2D 1H-NMR spectroscopy. For the first time, variable-pressure 2D exchange spectra were used for mechanistic assignments. cis/trans-Isomerisation was found to be the fastest reaction (in CHCl3/CDCl3), with a small acceleration at higher pressure: it is concluded to be an intramolecular process with a slightly contracted six-coordinate transition state. The intermolecular (MeO)3PO exchange on the cis- and trans-isomer are second-order processes and are strongly accelerated by increased pressure: Ia mechanisms are suggested without ruling out limiting A mechanisms.  相似文献   

16.
The nucleophilic substitution reaction under NH3 chemical ionization (CI) conditions in cis- and trans-1,2-dihydroxybenzosuberans (1–4) has been studied with the help of ND3 CI and metastable data. The results indicate that in the parent diols 1 (cis) and 2 (trans), the substitution ion [MsH]+, is produced mainly by the loss of H2O from the [MNH4]+ ion (SNi reaction) while in their 7-methoxy derivatives 3 and 4, the ion-molecule reaction between [M? OH]+ and NH3 seems to be the major pathway for the formation of [MsH]+. The substitution ion from 1 and 2 and the [MH]+ ion from trans-1-amino-2-hydroxybenzosuberan give similar collision-induced dissociation mass-analysed ion kinetic energy spectra. Interestingly, their diacetates do not undergo the substitution reaction.  相似文献   

17.
The complex cis-[Re(apa)Cl2(PPh3)] (1) (H3apa?=?N-(2-aminophenyl)salicylideneamine) was prepared by reaction of trans-[ReOCl3(PPh3)2] with H3apa in ethanol. Apa acts as a tridentate chelate ligand via the doubly deprotonated amino nitrogen (which is coordinated as an imide), the amino nitrogen and the deprotonated phenolic oxygen atoms. The two chlorides lie cis to each other in a distorted octahedral geometry around the rhenium(V) centre.  相似文献   

18.
Rate constants for the gas-phase reactions of the four oxygenated biogenic organic compounds cis-3-hexen-1-ol, cis-3-hexenylacetate, trans-2-hexenal, and linalool with OH radicals, NO3 radicals, and O3 have been determined at 296 ± 2 K and atmospheric pressure of air using relative rate methods. The rate constants obtained were (in cm3 molecule?1 s?1 units): cis-3-hexen-1-ol: (1.08 ± 0.22) × 10?10 for reaction with the OH radical; (2.72 ± 0.83) × 10?13 for reaction with the NO3 radical; and (6.4 ± 1.7) × 10?17 for reaction with O3; cis-3-hexenylacetate: (7.84 ± 1.64) × 10?11 for reaction with the OH radical; (2.46 ± 0.75) × 10?13 for reaction with the NO3 radical; and (5.4 ± 1.4) × 10?17 for reaction with O3; trans-2-hexenal: (4.41 ± 0.94) × 10?11 for reaction with the OH radical; (1.21 ± 0.44) × 10?14 for reaction with the NO3 radical; and (2.0 ± 1.0) × 10?18 for reaction with O3; and linalool: (1.59 ± 0.40) × 10?10 for reaction with the OH radical; (1.12 ± 0.40) × 10?11 for reaction with the NO3 radical; and (4.3 ± 1.6) × 10?16 for reaction with O3. Combining these rate constants with estimated ambient tropospheric concentrations of OH radicals, NO3 radicals, and O3 results in calculated tropospheric lifetimes of these oxygenated organic compounds of a few hours. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
Taking advantage of a 6π electrocyclization–aromatization strategy, we accomplished the first and asymmetric total synthesis of rubriflordilactone B, a heptacyclic Schisandraceae bisnortriterpenoid featuring a tetrasubstituted arene moiety. The left‐hand fragment was accessed through a chiral‐pool‐based route, and linked to the right‐hand fragment by a Sonogashira coupling. The cis geometry of the electrocyclization substrates was established by hydrogenation or hydrosilylation of the alkyne. An electrocyclization–aromatization sequence finally built the multisubstituted arene. The hydrosilylation approach was of significant advantage in terms of reaction scale, reproducibility, and intermediate stability. The structure of synthetic rubriflordilactone B was validated by X‐ray crystallographic analysis, and found to be consistent with that reported for the authentic natural product based on an independent X‐ray crystallographic analysis. However, obvious differences in the NMR spectra of the synthetic and authentic samples suggest that the authentic samples subjected to X‐ray crystallography and NMR spectroscopy were two different compounds.  相似文献   

20.
The work reports the unexpected reaction of diphenyldibromo antimonates (III) with PtCl2 and cis‐[PtCl2(PPh3)2]. The reaction gives triphenylstibine containing PtII complexes viz. cis‐[PtBr2(SbPh3)2] ( 1 ), trans‐[[PtBr(Ph)(SbPh3)2] ( 2 ), [NMe4][PtBr3(SbPh3)] ( 3 ), and cis‐[PtBr2(PPh3)(SbPh3)] ( 4 ). All the complexes were characterised by elemental analyses, IR, Raman, 195Pt NMR, FAB mass spectroscopy and X‐ray crystallography. A plausible mechanism via the phenyl migration is proposed for the formation of these complexes. The average Pt–Br distance in 1 is 2.456(2) Å, in 2 2.496 Å(trans to Ph) while in 3 it is 2.476 Å (trans to Sb) implying a comparable trans influence of Ph3Sb and Ph3P.  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