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.

Conformational propensities of N-t-butoxycarbonyl-glycine-(E/Z)-dehydrophenylalanine N′-methylamides (Boc-Gly-(E/Z)-ΔPhe-NHMe) in chloroform were investigated by NMR and IR techniques. The low-temperature crystal structure of the E isomer was determined by single crystal X-ray diffraction and the experimental data were elaborated by theoretical calculations using DFT (B3LYP, M06-2X) and MP2 approaches. The β-turn tendencies for both isomers were determined in the gas phase and in the presence of solvent. The obtained results reveal that the configuration of ΔPhe residue significantly affects the conformations of the studied dehydropeptides. The tendency to adopt β-turn conformations is significantly lower for the E isomer (Boc-Gly-(E)-ΔPhe-NHMe), both in gas phase and in chloroform solution.

  相似文献   

2.
Vinylogous β-Cleavage of Enones: UV.-irradiation of 4-(3′,7′,7′-trimethyl-2′-oxabicyclo[3.2.0]hept-3′-ene-1′-yl)but-3-ene-2-on On 1π,π*-excitation (λ = 254 nm) in acetonitrile (E/Z)- 2 is converted into the isomers 4–9 and undergoes fragmentation yielding 10 ; in methanol (E/Z)- 2 gives 7–10 and is transformed into 11 by incorporation of the solvent. On 1π,π*-excitation (λ λ?347 nm; benzene-d6) (E)- 2 is isomerized into (Z)- 2 , which is converted into the isomers 3 and 4 by further irradiation. 1π,π*-Excitation (λ = 254 nm; acetonitrile) of 4 gives 6 and (E)- 9 , whereas UV.-irradiation (λ = 254 nm; acetonitrile-d3) of 5 yields (E)- 7 and 8 . On 1π,π*-excitation (λ = 254 nm; acetonitrile) of (E/Z)- 12 the compounds (E)- 14 and (E)- 15 are obtained.  相似文献   

3.
The synthesis of dinuclear ruthenium sawhorse-type complexes [Ru2(μ-ArCH:Rhod)2(CO)4]n 12a–e and [Ru2(ArCH:Rhod)2(μ-ArCH:Rhod)2(CO)4] 13a–e through reaction of [Ru3(CO)10(NCMe)2] and [Ru3(CO)12] and the corresponding (Z)-5-arylidenerhodanines (ArCH:Rhod) 10a–e, respectively, are reported. These complexes are arranged in a sawhorse structure in which two bridged (Z)-5-arylidenerhodanines coordinate to the metals using sulfur and nitrogen of the rhodanine ring. A Density Functional Theory method was used to gain insight into the polymerization process by calculating dimerization Gibbs energies (ΔGdim). Values between ?10.7 and ?5.3 kcal mol?1 indicate that dimerization is a spontaneous process. A reaction pathway for formation of the sawhorse compounds [Ru2(μ-ArCH:Rhod)2(CO)4] was calculated and the rate-determining step for the mechanism is coordination of a second (Z)-5-arylidenerhodanine ligand with activation energies between 41.1 and 47.8 kcal mol?1. In order to understand the apparent thermodynamic favorability of the fragmentation step, we calculated the fragmentation energy (ΔEFrag) for the key intermediate and its energetic contributors, the interaction energy, ΔEint and the reorganization energy, ΔEreorg. Low values of ΔEFrag imply that the fragmentation is thermodynamically facile. Large values of ΔEint are countered by opposite and large values of ΔEreorg which indicate that the cleavage of the trimetallic intermediate aggregate is determined by the nature of the ligand and the balance between its interaction with the metal and the extent of structural reorganization.  相似文献   

4.
The synthesis and electrochemical properties of ferrocene conjugates are presented for the purpose of investigating adenosine 5′‐[γ‐ferrocenoylalkyl] triphosphate ( 1 a – 4 a , ferrocene (Fc)–ATP) as co‐substrates for phosphorylation reactions. Compounds 1 a – 4 a were synthesized, purified by HPLC, and characterized by NMR spectroscopy and mass spectrometry. In solution, all Fc–ATP bioconjugates exhibit a reversible one‐electron redox process with a half‐wave potential (E1/2) in the 390–430 mV range, peak separations (ΔEp) in the 40–70 mV range, and the peak current ratio (ipa/ipc) near unity. The peptide‐modified surface Glu‐Gly‐Ile‐Tyr‐Asp‐Val‐Pro was used to study the sarcoma‐related protein (Src) kinase activity by employing the Fc–ATP bioconjugates as co‐substrates. Subsequent kinase‐catalyzed transfer of the γ‐Fc‐phosphate group to the tyrosine residues of the surface‐bound peptides was characterized by a formal potential (Eo) ≈390 mV (vs. Ag/AgCl). The Fc‐coverage, estimated by time‐of‐flight secondary‐ion mass spectrometry (TOF‐SIMS) and cyclic voltammetry (CV), suggested validity of Fc–ATP conjugates as kinase co‐substrates. Depending on the length of the alkyl spacer of the Fc–ATP conjugate, different current densities were obtained, pointing to a direct correlation between the two. Molecular modeling revealed that the structural constraint imposed by the short alkyl spacer ( 1 a ) causes a steric congestion and negatively affects the outcome of phosphorylation reaction. An optimal analytical response was obtained with the Fc–ATP conjugates with linker lengths longer than six CH2 groups.  相似文献   

5.
The thermal reaction of homoazulene (=bicyclo[5.3.1]undeca‐1,3,5,7,9‐pentaene; 2 ) with dimethyl acetylenedicarboxylate (ADM) in 1,2‐dichloroethane (ClCH2CH2Cl) results, in contrast to an earlier report [5], in formation of not only dimethyl homoheptalene‐4,5‐dicarboxylate (=bicyclo[5.5.1]trideca‐1,3,5,7,9,11‐hexaene‐4,5‐dicarboxylate; 3 ), but also of a 4 : 1 mixture of 3 and dimethyl homoheptalene‐2,3‐dicarboxylate ( 13 ) in almost quantitative yield (Schemes 1 and 3). The structures of both homoheptalenes have been corroborated by X‐ray crystal‐structure analysis (Fig. 5). The double‐bond‐shifted (DBS) isomers 3 ′ and 13 ′ of 3 and 13 , respectively, could not be detected in their 1H‐NMR spectra (600 MHz threshold of detection ≥0.5%), in agreement with the AM1‐calculated ΔH values of the four isomeric homoheptalene‐dicarboxylates (cf. Table 4). Vilsmeyer formylation of homoazulene ( 2 ) gave homoazulene‐8‐carbaldehyde ( 14 ) in a yield of 67%, which, on treatment with benzylidene‐(triphenyl)‐λ5‐phosphane, gave, in almost quantitative yield, a 1.6 : 1 mixture of (Z)‐ and (E)‐8‐styrylhomoazulene ((Z)‐ 15 and (E)‐ 15 , resp.). Thermal reaction of the latter mixture with ADM in 1,2‐dichloroethane led, in a yield of 42%, to a 5 : 1 mixture of dimethyl (Z)‐ and (E)‐2‐styrylhomoheptalene‐4,5‐dicarboxylate ((Z)‐ 15 and (E)‐ 16 , resp.). Both isomers were separated by column chromatography on silica gel. Again, the DBS isomers of (Z)‐ 16 and (E)‐ 16 , i.e., (Z)‐ 16 ′ and (E)‐ 16 ′, could not be detected in the 1H‐NMR spectra (600 MHz) of pure (Z)‐ 16 and (E)‐ 16 .  相似文献   

6.
Comparison of the crystal structures of two pentadehydropeptides containing ΔPhe residues, namely (Z,Z)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔZPhe2–Gly3–ΔZPhe4–Gly5–OH) methanol solvate, C29H33N5O8·CH4O, (I), and (E,E)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔEPhe2–Gly3–ΔEPhe4–Gly5–OH), C29H33N5O8, (II), indicates that the ΔZPhe residue is a more effective inducer of folded structures than the ΔEPhe residue. The values of the torsion angles ϕ and ψ show the presence of two type‐III′β‐turns at the ΔZPhe residues and one type‐II β‐turn at the ΔEPhe residue. All amino acids are linked trans to each other in both peptides. β‐Turns present in the peptides are stabilized by intramolecular 4→1 hydrogen bonds. Molecules in both structures form two‐dimensional hydrogen‐bond networks parallel to the (100) plane.  相似文献   

7.
Photoisomerization of an aromatic analogue of retinoic acid, ethyl all-trans-9-(4-methoxy-2,3,6-trimethylphenyl)-3,7-dimethyl-nona-2,4,6, 8-tetraenoate 1 in dilute solutions of hexane, benzene, and ethanol yielded multi-component mixtures of cis isomers which were separated by HPLC. FT-1H-NMR. at 270 MHz and, in some cases, homonuclear decoupling and Overhauser experiments as well as 13C-NMR. were applied to establish the structures of 4 mono-cis, 4 (of 6 possible) di-cis, and 3 (of 4 possible) tri-cis isomers. The structures of 3 isomeric esters, namely (2Z, 4E, 6E, 8E) 6 , (2Z, 4Z, 6E, 8E) 9 , and (2Z, 4Z, 6Z, 8E) 7 were independently confirmed by direct syntheses. The 1H-NMR. data of all these compounds and the 13C-NMR. data of the all-trans and of 6 cis isomers available in sufficiently large quantities are discussed.  相似文献   

8.
The synthesis of the cyclopropyl silyl ketones 1 – 4 is described. The trimethylsilyl ketone 1 was prepared from geraniol ((E)- 5 ) in ca. 10% overall yield by cyclopropanation leading to 6 , CrO3 oxidation to the aldehyde 8 , reaction of the latter with trimethylsilyl anion to 14A + B , and CrO3 oxidation to 1 . Also for the (t-butyl)dimethylsilyl ketones 2 – 4 , an efficient four-step synthesis with overall yields of 48%, 85%, and 13%, respectively, was elaborated, starting from the allylic alcohols (E)- 5 , and 23 . The method of preparation involves as the key step a Wittig rearrangement of the silylallyl ethers ((E/Z)- 20 , 24 ) to the silyl alcohols ((E/Z)- 21 , 25 ), subsequent cyclopropanation ( 19A + B , 22A + B , 26 ), and oxidation to the cyclopropyl silyl ketones 2 – 4 .  相似文献   

9.
The difference between the expectation values of the total electronic kinetic energy operator (ΔEK), and the operators accounting for the Coulombic interactions between the electrons and nuclei (ΔVen), between all pairs of electrons (ΔVee), and between all pairs of nuclei (ΔVnn) for the product and reactant species in a wide variety of hydrocarbon reactions are calculated using single determinant basis set data reported in the literature. Following Allen, their contributions to ΔET, the difference between the corresponding total molecular energies and thus the reaction heat, are grouped together as a repulsion energy term, ΔErep = ΔEK + ΔVee + ΔVnn, and an attraction energy term ΔEattr = ΔVen. For all but 2 of the 71 individual reactions considered in this paper, the experimental reaction heat at 0°K corrected for zero-point energy contributions, (ΔH)zpe, is the result of near compensation between far larger ΔErep and ΔEattr terms, in sharp contrast to the much smaller ΔErep and ΔEattr terms which are characteristic of many molecular rotation processes. By matching the sign of (ΔH)zpe with that of ΔErep or ΔEattr, as the case may be, the reactions are classified as attractive-dominant or repulsive-dominant (46 in the former class and 23 in the latter), a property which is independent of the direction in which the reaction is written. The sign and magnitude of ΔVee, ΔVnn, and ΔVen and reaction category are discussed in relation to the various kinds of structural change involved in going from reactants to products. For the vast majority of reactions, the numerical relationship ΔVee ≈ ΔVnn has been found to hold to within a few percent.  相似文献   

10.
Configurational and Conformational Isomeric Paratopic, Rotational Dynamics Tetraepoxy[30]annulenes(6.2.6.2) and Diatropic Tetraoxa[30]porphyrin(6.2.6.2) Dications: Detection of a Tetraepoxy[31]annulene(6.2.6.2)Radical Cation The synthesis of tetraepoxy[32]annulenes(6.2.6.2) ( 4 ) by a cyclizing twofold Wittig reaction of (E,E,E)-5,5′-(hexa-1,3,5-triene-1,6-diyl)bis[furan-2-carbaldehyde] ( 6 ) and the corresponding bis-phosphonium salt 7 is described (Scheme 1). Contrary to the configuration of the educts, the obtained annulenes 4a and 4b are (Z,E,E,E,Z,E,E,E)- and (E,Z,E,E,E,Z,E,E)-configurated, respectively. The 1H-NMR spectra establish the paratropic, antiaromatic character of 4 . The annulenes 4 are highly dynamic systems, the (E)-ethenediyl bridges rotate around the adjacent σ-bonds, these rotations are frozen at −80°. The McMurry condensation of dialdehyde 6 yields the (E,E,Z,E,E,E,Z)-4,5-dihydrotetraepoxy[32]annulene(6.2.6.2) ( 13a ), where the configuration of the dialdehyde 6 – beside the hydrogenated double bond – is retained. As result of an intramolecular McMurry reaction of 6 , (Z,E,Z,Z)-dioxa[16]annulene(6.2) 14 is formed. By oxidation of the [32]annulenes(6.2.6.2) 4a and 4b , a mixture of the four stereoisomeric tetraoxa[30]porphyrin(6.2.6.2) dications 5a / 5a ′/ 5b / 5c is obtained; the configuration of the isomers is determined by COSY, NOESY, and NOE experiments. The Δδ values (26.81, 25.83, and 21.11 ppm) underline the diatropic, aromatic character of the dications 5 , the Soret bands are shifted bathochromically to 550 nm, and the Q-bands are in the NIR region (896 – 1039 nm). The dihydroannulene 13a is dehydrogenated by p-chloroanil (tetrachloro-1,4-benzoquinone) to give the annulenes 4a and 4b , its oxidation with DDQ (=4,5-dichloro-3,6-dioxocyclohexa-1,4-diene-1,2-dicarbonitrile) results in the same mixture of dications 5 . Entirely different results are obtained by reaction of the dihydroannulene 13a with DDQ. Here, the (E,E,E,Z,E,E,E,Z) tetraoxa[30]porphyrin(6.2.6.2) dication 5c – formed only in traces from 4a / 4b – is the main product. Beside 5c , a by-product (3%) can be isolated, which turns out (ESR, conductivity) to be the (E,E,E,Z,E,E,E,Z)-tetraoxa[31]porphyrin(6.2.6.2) radical cation 16 , obviously the intermediate in the oxidation sequence of the annulene to the dication. This result leads to the conclusion that the reaction of the dihydro compound 13a with p-chloroanil and DDQ follows different reaction mechanisms. For all isolated stereoisomeric tetraepoxy annulenes and tetraoxaporphyrin dications, the ΔHf values are calculated by the semiempiric AM1 method. The results are in agreement with the experimental observations. All data confirm the antiaromaticity of the tetraepoxy[32]annulenes(6.2.6.2) 4 and the aromaticity of the tetraoxa[30]porphyrin(6.2.6.2) dications.  相似文献   

11.
On triplet excitation (E)- 2 isomerizes to (Z)- 2 and reacts by cleavage of the C(γ), O-bond to isomeric δ-ketoester compounds ( 3 and 4 ) and 2,5-dihydrofuran compounds ( 5 and 19 , s. Scheme 1). - On singulet excitation (E)- 2 gives mainly isomers formed by cleavage of the C(γ), C(δ)-bond ( 6–14 , s. Scheme 1). However, the products 3–5 of the triplet induced cleavage of the C(γ), O-bond are obtained in small amounts, too. The conversion of (E)- 2 to an intermediate ketonium-ylide b (s. Scheme 5) is proven by the isolation of its cyclization product 13 and of the acetals 16 and 17 , the products of solvent addition to b . - Excitation (λ = 254 nm) of the enol ether (E/Z)- 6 yields the isomeric α, β-unsaturated ε-ketoesters (E/Z)- 8 and 9 , which undergo photodeconjugation to give the isomeric γ, δ-unsaturated ε-ketoesters (E/Z)- 10 . - On treatment with BF3O(C2H5)2 (E)- 2 isomerizes by cleavage of the C(δ), O-bond to the γ-ketoester (E)- 20 (s. Scheme 2). Conversion of (Z)- 2 with FeCl3 gives the isomeric furan compound 21 exclusively.  相似文献   

12.
The green seaweed Caulerpa taxifolia (VAHL ) C. AGARDH (Caulerpales), which, after its recent accidental introduction, is growing in the region of Cap Martin much more vigorously than in the tropics, is shown to contain the known sesquiterpenic toxins caulerpenyne ( 1 ) – in larger amounts than in tropical Caulerpales – and oxytoxin 1 ( 2 ). Novel, potentially toxic products isolated in small amounts from this seaweed include the sesquiterpenes taxifolial A ( = (5E)-6,10-dimethyl-2-[(E)2-oxoethylidene]undeca-5,9-dien-7- yne-1,3-diyl diacetate; 3 ), taxifolial B (= (1E,6E,10E)-3-[( Z )-acetoxymethylidene]-7, 11-dimethyl-12-oxododeca-1,6,10-trien-8-yne-1,4-diyl diacetate; 4 ), 10,11-epoxycaulerpenyne ( = (1E,6E)-3-[(Z)-acetoxymethylidene]-10,11-epoxy-7, 11-dimethyldodeca-1,6-dien-8-yne-1,4-diyl diacetate; 1:1 diastereoisomer mixture; 5 ), and taxifolial C ( = (2Z,6E)-3-formyl-7,11-dimethyldodeca-2,6,10-trien-8-yne-1,1, 4-triyl triacetate; 6 ), besides, as the first example of a monoterpene from the Caulerpales, taxifolial D ( = (2Z)-3,7-dimethylocta-2, 6-dien-4-ynal; 7 ).  相似文献   

13.
Eunicellane diterpenes with a C(7)=C(16) (Δ7,16; see 17 – 19 ), (Z)‐C(7)=C(8) ((7Z)Δ7,8; see 20 – 23 ), and (Z)‐C(6)=C(7) ((6Z)Δ6,7; see 10 ) bond, an uncommon feature in the case of extensive functionalization at the cyclohexane ring and the latter exhibiting uncommon configurations, were isolated from the gorgonian Eunicella cavolinii from Marseille (Figs. 5 and 2). The gorgonian Eunicella singularis (=Eunicella stricta) from the same area gave (7Z)Δ7,8, Δ7,16, and (6E)Δ6,7 analogs 24 , 25 , and 13 , respectively (Fig. 5 and Scheme). The (6E)Δ6,7 moiety of 13 – characterized by a slow 180° conformational flipping (Fig. 3) that results in broadening of NMR signals – makes the macrocycle highly strained. This may explain the spontaneous conversion of 13 to the 6‐methoxy‐7‐hydroxy derivative 14 in the presence of MeOH at −20° in the dark (Scheme). The isomeric, deacylated analogue 10 showed only little broadening of NMR signals and proved stable, in accordance with the less strained nature of this compound (Fig. 4).  相似文献   

14.
We report a new method for constructing propionate-derived trisubstituted alkene motifs in a stereoselective manner. 1-Substituted 1,1-di(pinacolatoboryl)-(E)-alk-2-enes are generated in situ from 1-substituted 1,1-di(pinacolatoboryl)alk-3-enes through ruthenium(II)-catalyzed double-bond transposition. These species undergo a chiral phosphoric acid catalyzed allylation reaction of aldehydes to produce the E isomers of anti-homoallylic alcohols. On the other hand, the corresponding Z isomers of anti-homoallylic alcohols are obtained when a dimeric palladium(I) complex is employed as the catalyst for this double-bond transposition. Thus, both E and Z isomers can be synthesized from the same starting materials. A B−C(sp2) bond remaining with the allylation product undergoes the Suzuki–Miyaura cross-coupling reaction to furnish a propionate-derived trisubstituted alkene motif in a stereo-defined form. The present method to construct the motifs with (E)- and (Z)-alkenes are successfully applied to the syntheses of (+)-isotrichostatic acid, (−)-isotrichostatin RK, and (+)-trichostatic acid, respectively.  相似文献   

15.
The Z and E isomers of 3‐[4‐(dimethylamino)phenyl]‐2‐(2,4,6‐tribromophenyl)acrylonitrile, C17H13Br3N2, ( 1 ), were obtained simultaneously by a Knoevenagel condensation between 4‐(dimethylamino)benzaldehyde and 2‐(2,4,6‐tribromophenyl)acetonitrile, and were investigated by X‐ray diffraction and density functional theory (DFT) quantum‐chemical calculations. The (Z)‐( 1 ) isomer is monoclinic (space group P21/n, Z′ = 1), whereas the (E)‐( 1 ) isomer is triclinic (space group P, Z′ = 2). The two crystallographically‐independent molecules of (E)‐( 1 ) adopt similar geometries. The corresponding bond lengths and angles in the two isomers of ( 1 ) are very similar. The difference in the calculated total energies of isolated molecules of (Z)‐( 1 ) and (E)‐( 1 ) with DFT‐optimized geometries is ∼4.47 kJ mol−1, with the minimum value corresponding to the Z isomer. The crystal structure of (Z)‐( 1 ) reveals strong intermolecular nonvalent Br…N [3.100 (2) and 3.216 (3) Å] interactions which link the molecules into layers parallel to (10). In contrast, molecules of (E)‐( 1 ) in the crystal are bound to each other by strong nonvalent Br…Br [3.5556 (10) Å] and weak Br…N [3.433 (4) Å] interactions, forming chains propagating along [110]. The crystal packing of (Z)‐( 1 ) is denser than that of (E)‐( 1 ), implying that the crystal structure realized for (Z)‐( 1 ) is more stable than that for (E)‐( 1 ).  相似文献   

16.
Deuteration of Betanidine and Indicaxanthine. (E/Z)-Stereoisomerism in Betalaines Previously unexplained 1H-NMR signals of the red-violet betanidine and of the yellow indicaxanthine in CF3COOH are interpreted with the help of deuteration experiments in CF3COOD. They confirm the 1,7-diazaheptamethinium chromophore of these pigments and further show an (E/Z)-stereoisomerism at one of the partial double bonds: both betanidine and isobetanidine, in CF3COOH solution, consist of a ~ 75:25 mixture of the (12E)- ( 1 resp. 12 ) and the (12Z)-stereoisomer ( 2 resp. 13 ); indicaxanthine, in the same solvent, is made up of a ~ 65:35 mixture of the (8E)- ( 14 ) and the (8Z)-stereoisomer ( 15 ). The interconversions 1?2 , 12?13 and 14?15 in CF3COOH are so fast that these isomers cannot be separated from each other.  相似文献   

17.
Azimines IV. Kinetics and Mechanism of the Thermal Stereoisomerization of 2,3-Diaryl-1-phthalimido-azimines1) Mixtures of (1E, 2Z)- and (1Z, 2E)-2-phenyl-1-phthalimido-3-p-tolyl-azimine ( 3a and 3b , resp.) and (1E, 2Z)- and (1Z, 2E)-3-phenyl-1-phthalimido-2-p-tolylazimine ( 4a and 4b , resp.) were obtained by the addition of oxidatively generated phthalimido-nitrene (6) to (E)- and (Z)-4-methyl-azobenzene ( 7a and 7b , resp.). Whereas complete separation of the 4 isomers 3a, 3b, 4a and 4b was not possible, partial separation by chromatography and crystallization led to 5 differently composed mixtures of azimine isomers. The spectroscopic properties of these mixtures (UV., 1H-NMR.) were used to determine the ratios of isomers in the mixtures, and served as a tool for the assignment of constitution and configuration to those isomers which were dominant in each of these mixtures, respectively. Investigation of the isomerization of the azimines 3a, 3b, 4a and 4b within the 5 mixtures at various concentrations by 1H-NMR.-spectroscopy at room temperature revealed that only stereoisomers are interconverted ( 3a ? 3b; 4a ? 4b) and that the (1E, 2Z) ? (1Z, 2E) stereoisomerization is a unimolecular reaction. These observations exclude an isomerization mechanism via an intermediate 1-phthalimido-triaziridine (2) or via dimerization of 1-phthalimido-azimines (1) , respectively. The 3-p-tolyl substituted stereoisomers 3a and 3b isomerized slightly slower than the 3-phenyl substituted ones 4a and 4b , an effect which is consistent with the assumption that the rate determining step of the interconversion of (1E, 2Z)- and (1Z, 2E)-1-phthalimido-azimines (1a ? 1b) is the stereoisomerization of the stereogenic center at N(2), N(3), either by inversion of N(3) or by rotation around the N(2), N(3) bond. The total isomerization process is assumed to occur via the thermodynamically less stable (1Z, 2Z)- and (1E, 2E)-isomers 1c and 1d , respectively, as intermediates in undetectably low concentrations which stay in rapidly established equilibria with the observed, thermodynamically more stable (1E, 2Z)- and (1Z, 2E)-isomers 1a and 1b , respectively. At higher temperatures, the azimines 3 and 4 are transformed into N-phenyl-N,N′-phthaloyl-N′-p-tolyl-hydrazine (8) with loss of nitrogen.  相似文献   

18.
The reaction of diazomethane with some (E) and (Z)-2-substituted-4-methylene-5(4)-oxazolones ( 1a-c ) under two different conditions, has been studied. (E) and (Z)-1,2-disubstituted-7-oxo-6-oxa-4-azaspiro[2.4]-hept-4-enes ( 3a-c, 4a-c ) were mainly obtained, together with multiple addition compounds. The reaction showed to be stereoselective only when the substituents were aromatic. Acid hydrolysis of compounds 3a and 4a produced a mixture of (E) and (Z)-3,5-disubstituted-tetrahydrofuran-2-ones ( 8a, 9a ). Smooth methanolysis of the ring led to (E) and (Z)-1-benzamido-cyclopropanecarboxylic esters ( 10a-c, 11a-c ), which, on acid hydrolysis, gave (E) and (Z)-1-amino-2-phenylcyclopropanecarboxylic acids 12a and 13a . The pmr spectra have been analyzed by an iterative computer method, and the computed best values obtained have been used to deduce the stereochemistry of the spiroderivatives.  相似文献   

19.
Superheteroaromatic Systems with Furan Building Blocks: Isomeric Antiaromatic Tetraepoxy[36]annulenes(6.4.6.4) and Aromatic Tetraoxa[34]porphyrin(6.4.6.4) Dications The title compounds are available by a twofold cyclizing Wittig reaction of (all‐E)‐3,3′‐(hexa‐1,3,5‐triene‐1,6‐diyldifuran‐5,2‐diyl)bis[prop‐2‐enal] ( 4 ) with (all‐E)‐(hexa‐1,3,5‐triene‐1,6‐diyl)bis(furan‐5,2‐diylmethylene)bis[triphenylphosphonium] dibromide ( 7 ). Two conformational isomers 2a / 2a ′ of (Z,E,E,E,E,Z,E,E,E,E)‐tetraepoxy[36]annulene(6.4.6.4) are obtained. The oxidation of 2a / 2a ′ yields two (E,E,Z,E,E,E,E,Z,E,E)‐tetraoxa[34]porphyrin(6.4.6.4) dications 3a / 3a ′, which are conformers, too. The oxidation of 2a / 2a ′ is accompanied by the isomerization of four ethen‐1,2‐diyl bridges. The reduction of the dications 3a / 3a ′ leads to the new (E,E,Z,E,E,E,E,Z,E,E)‐tetraepoxy[36]annulene(6.4.6.4) ( 2b ) and (E,E,E,Z,E,E,E,E,Z,E)‐tetraepoxy[36]annulene(6.4.6.4) ( 2c ). In 2b as well as in 2c , both 1,3‐butadiene‐1,4‐diyl bridges are rotating until −90°. The Δδ values, i.e., the maximum δ difference of the `inner' and `outer' perimeter protons of 3a / 3a ′ (26.62 and 25.32 ppm) are of the same size as the Δδ value of the tetramethyl[34]porphyrin(5.5.5.5) dication ( 1 ; Δδ=25.3 ppm); therefore, they might be called `superheteroaromatic' too. The Δδ values of the tetraepoxy[36]annulenes(6.4.6.4) ( 2a – c ; Δδ=2.3 – 3.3 ppm) establish that they are still paratropic; they represent the most expanded antiaromatic systems yet known.  相似文献   

20.
《Tetrahedron letters》2004,45(29):5563-5567
Reactions of (E)- and (Z)-ketoximes with trialkylphosphines and diphenyl disulfide (PhSSPh) have been compared to gain insight into the mechanisms involved and their potential applications. N-Sulfenylimine isomers and ketimines have been spectroscopically characterised. Both the E and Z isomers of erythromycin A oxime, when treated with Bu3P and PhSSPh (1:4:8 ratio), give the same N-phenylsulfenyl ketimine (of configuration E) as the major compound, whereas with Bu3P or Me3P and PySeSePy (1:8:4 ratio) afford the imine in good yield. Clarithromycin oxime behaves similarly.  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