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.
A convenient and high yielding method for the synthesis of R-(-)-imperanene, starting from the readily available natural lignan hydroxymatairesinol from Norway spruce, was developed. Hydroxymatairesinol was degraded in strongly basic aqueous conditions to (E)-4-(4-hydroxy-3-methoxyphenyl)-2-(4-hydroxy-3-methoxyphenylmethyl)but-3-enoic acid, which was esterified and then reduced by LiAlH(4) to afford R-(-)-imperanene. The configuration at the crucial stereocenter was preserved in the synthesis, and the obtained product was identified by optical rotation measurements and chiral HPLC analyses as the R-(-)-enantiomer (ee 86-92%).  相似文献   

2.
The oxidative transformation of the two isomers of the natural lignan hydroxymatairesinol from Norway Spruce (Picea abies) by 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), has been studied. Significant differences in the outcome of the reactions were observed when the pure isomers of hydroxymatairesinol were reacted with DDQ under the same conditions. The different stereoelectronic effects in the two isomers as well as their conformational structures seem to determine the site of reaction, which results in different reaction products. Several products were identified by GC-MS and NMR spectroscopy. Oxomatairesinol was obtained in a yield of 25%.  相似文献   

3.
Investigating reactions of functionalized p-quinone methides continues to advance our design of a reagent being developed for controlled, in situ modification of DNA via phosphodiester alkylation. Previously reported investigations of p-quinone methides derived from catechols allowed for trapping of isolable trialkyl phosphates for characterization and mechanistic information. However, lactone formation with these derivatives required long reaction times, resulting in an unfavorable mixture of trialkyl phosphate and hydrolysis products. To enhance the rate and efficacy of trialkyl phosphate formation and trapping, a phenol derived p-quinone methide has been designed to enforce a conformation favoring lactonization of the dialkyl phosphate alkylated intermediate. The relative rates of phosphodiester alkylation and subsequent trapping of the phosphotriester adduct have been examined by UV and (1)H NMR analysis for p-quinone methide precursor 1 and the corresponding control, 1'. The incorporation of a methyl group at the meta-position of 1 (relative to 1') significantly improves the rate of lactionization to provide a much higher yield of the desired product, lactonized phosphotriester 5. The control reaction with 1' afforded only a minor amount of the corresponding lactonized trialkyl phosphate 5'.  相似文献   

4.
The content of reactive groups such as OH, CO, and COOH was increased by modifying hydrolyzed lignin with sulfuric acid and sodium hydroxide. The increase was confirmed by IR spectral analysis. The sorptive capacity of the resulting hydrolyzed lignin derivatives was increased sharply by base activation. __________ Translated from Khimiya Prirodnykh Soedinenii, No. 1, pp. 70–72, January–February, 2006.  相似文献   

5.
The polyaromatic structure of lignin has long been recognized as a key contributor to the rigidity of plant vascular tissues. Although lignin structure was once conceptualized as a highly networked, heterogeneous, high molecular weight polymer, recent studies have suggested a very different configuration may exist in planta. These findings, coupled with the increasing attention and interest in efficiently utilizing lignocellulosic materials for green materials and energy applications, have renewed interest in lignin chemistry. Here we focus on quinone methides (QMs)—key intermediates in lignin polymerization—that are quenched via reaction with cell-wall-available nucleophiles. Reactions with alcohol and uronic acid groups of hemicelluloses, for example, can lead to lignin-carbohydrate crosslinks. Our work is a first step toward exploring potential QM reactions with nucleophilic groups in cell wall proteins. We conducted a model compound study wherein the lignin model compound guaiacylglycerol-β-guaiacyl ether 1, was converted to its QM 2, then reacted with amino acids bearing nucleophilic side-groups. Yields for the QM-amino acid adducts ranged from quantitative in the case of QM-lysine 3, to zero (no reaction) in the cases of QM-threonine (Thr) 10 and QM-hydroxyproline (Hyp) 11. The structures of the QM-amino acid adducts were confirmed via 1D and 2D nuclear magnetic resonance (NMR) spectroscopy and density functional theory (DFT) calculations, thereby extending the lignin NMR database to include amino acid crosslinks. Some of the QM-amino acid adducts formed both syn- and anti-isomers, whereas others favored only one isomer. Because the QM-Thr 10 and QM-Hyp 11 compounds could not be experimentally prepared under conditions described here but could potentially form in vivo, we used DFT to calculate their NMR shifts. Characterization of these model adducts extends the lignin NMR database to aid in the identification of lignin-protein linkages in more complex in vitro and in vivo systems, and may allow for the identification of such linkages in planta.  相似文献   

6.
[reaction: see text] We describe here a four-step semisynthetic method for the preparation of enantiomerically pure (-)-enterolactone starting from the readily available lignan hydroxymatairesinol from Norway spruce (Picea abies). Hydroxymatairesinol was first hydrogenated to matairesinol. Matairesinol was esterified to afford the matairesinyl 4,4'-bistriflate, which was deoxygenated by palladium-catalyzed reduction to 3,3'-dimethylenterolactone. Demethylation of 3,3'-dimethylenterolactone and reduction with LiAlH(4) yielded (-)-enterolactone and (-)-enterodiol, respectively.  相似文献   

7.
Electronic perturbation of quinone methides (QM) greatly influences their stability and in turn alters the kinetics and product profile of QM reaction with deoxynucleosides. Consistent with the electron-deficient nature of this reactive intermediate, electron-donating substituents are stabilizing and electron-withdrawing substituents are destabilizing. For example, a dC N3-QM adduct is made stable over the course of observation (7 days) by the presence of an electron-withdrawing ester group that inhibits QM regeneration. Conversely, a related adduct with an electron-donating methyl group is very labile and regenerates its QM with a half-life of approximately 5 h. The generality of these effects is demonstrated with a series of alternative quinone methide precursors (QMP) containing a variety of substituents attached at different positions with respect to the exocyclic methylene. The rates of nucleophilic addition to substituted QMs measured by laser flash photolysis similarly span 5 orders of magnitude with electron-rich species reacting most slowly and electron-deficient species reacting most quickly. The reversibility of QM reaction can now be predictably adjusted for any desired application.  相似文献   

8.
Conclusions In the reaction of the cis-isomer of 3,4-epoxycarane with alcohols in acidic media, opening of the oxide ring was less regioselective than in the case of the -isomer. Rearrangement of the trans-isomer of 3,4-epoxycarane with transannual cyclopropyl participation occurs only in reactions with a secondary alcohol.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 3, pp. 648–653, March, 1985.  相似文献   

9.
Nucleophilicity of NH(2), N3, and O(2) centers of cytosine toward a model quinone methide (o-QM) as alkylating agent has been studied using DFT computational analysis [at the B3LYP/6-311+G(d,p) level]. Specific and bulk effects of water (by C-PCM model) on the alkylation pathways have been evaluated by analyzing both unassisted and water-assisted reaction mechanisms. An ancillary water molecule, H-bonded to the alkylating agent, may interact monofunctionally with the o-QM oxygen atom (passive mechanisms) or may participate bifunctionally in cyclic hydrogen-bonded structures as a proton shuttle (active mechanisms). A comparison of the unassisted with the water-assisted reaction mechanisms has been made on the basis of activation Gibbs free energies (DeltaG(++)). The gas-phase alkylation reaction at N3 does proceed through a passive mechanism that is preferred over both the active (by -6.3 kcal mol(-1)) and the unassisted process. In contrast, in the gas phase, the active assisted processes at NH(2) and O(2) centers are both favored over their unassisted counterparts by -4.0 and -2.2 kcal mol(-1), respectively. The catalytic effect of a water molecule, in gas phase, reduces the gap between the TSs of the O(2) and NH(2) reaction pathways, but the former remains more stable. Water bulk effect significantly modifies the relative importance of the unassisted and water-assisted alkylation mechanisms, favoring the former, in comparison to the gas-phase reactions. In particular, the unassisted alkylation becomes the preferred mechanism for the reaction at both the exocyclic (NH(2)) and the heterocyclic (N3) nitrogen atoms. By contrast, alkylation at the cytosine oxygen atom is a water-catalyzed process, since in water the active water-assisted mechanism is still favored. As far as competition, among all the possible mechanisms, our calculations unambiguously suggest that the most nucleophilic site both in gas phase (naked reagents: N3 > O(2) >or= NH(2)) and in water solution (solvated reagents: N3 > NH(2) > O(2)) is the heterocyclic nitrogen atom (N3) (DeltaG(++)(gas) = +7.1 kcal mol(-1), and DeltaG(++)(solv) = +13.7 kcal mol(-1)). Our investigation explains the high reactivity and selectivity of the cytosine moiety toward o-QM-like structures both in deoxymononucleoside and in a single-stranded DNA, on the basis of strong H-bonding interactions between reactants and solvent bulk effect. It also offers two general reactivity models in water, uncatalyzed and active water-catalyzed mechanisms (for nitrogen and oxygen nucleophiles, respectively), which should provide a general tool for the planning of nucleic acid modification.  相似文献   

10.
11.
Kinetic studies of various coupling reactions between silicon derivatives (SiF, SiCl, SiOMe, SiH) and organometallic reagents (RMgX, RLi) were performed for systems involving retention and inversion of configuration. The results show that for both stereochemical courses, the rupture of the SiX bond is not the rate determining step. Results are in good agreement with the rate determining formation of a pentacoordinate silicon intermediate.  相似文献   

12.
The reaction kinetics of 2-thiophenesulphonyl chloride with anion and neutral nucleophiles was studied in H2O, D2O and in protic solvents-H2O (10% vv) and aprotic solvents-H2O (10% vv) mixtures at 25°. Analysing the rate constants measured in water by Bronsted, Ritchie and Edwards equations the conclusion drawn that, for the nucleophilic order against the sulphonyl sulphur, basicity is of prime importance, although there may well be some dependence on polarizability and solvation. Solvent isotope effects show that the reactions occur by nucleophilic catalysis rather than by a general base mechanism. Water is the solvent in which there is the greater reactivity than in either protic solvents or aprotic-protic mixtures. By solubility measurements and applying Parker's equation the contributions of solvation energies of both reactants and transition states to the free energy of activation are calculated. Solvent effects on nucleophilic reactivities are discussed in terms of S parameters (similar to Ritchie N+ parameters), and by the approach of multiparameter empirical correlations. The data point out that solvation plays a large role on nucleophilic order. A complete comprehension of the problem would require an equation that takes into some account solvent effects. The homogeneous comparison of 2-thiophenesulphonyl chloride data with those of α-disulphone, p-anisyl p-methoxybenzenesulphinyl sulphone and benzenesulphonyl chloride shows that the same factors are involved in driving the nucleophilic reactivity for these compounds.  相似文献   

13.
Finding new pathways to novel materials is an open challenge in modern solid-state chemistry. Among the reasons that still prevent a rational planning of synthetic routes is the lack of an atomistic understanding at the moment of phase formation. Metastable phases are, in this respect, powerful points of access to new materials. For the synthetic efforts to fully take advantage of such peculiar intermediates, a precise atomistic understanding of critical processes in the solid state in its many facets, that is, nucleation patterns, formation and propagation of interfaces, intermediate structures, and phase growth, is mandatory. Recently we have started a systematic theoretical study of phase transitions, especially of processes with first-order thermodynamics, to reach a firm understanding of the atomistic mechanisms governing polymorphism in the solid state. A clear picture is emerging of the interplay between nucleation patterns, the evolution of domain interfaces and final material morphology. Therein intermediate metastable structural motifs with distinct atomic patterns are identified, which become exciting targets for chemical synthesis. Accordingly, a new way of implementing simulation strategies as a powerful support to the chemical intuition is emerging. Simulations of real materials under conditions corresponding to the experiments are shedding light onto yet elusive aspects of solid-solid transformations. Particularly, sharp insights into local nucleation and growth events allow the formulation of new concepts for rationalizing interfaces formed during phase nucleation and growth. Structurally different and confined in space, metastable interfaces occurring during polymorph transformations bring about distinct diffusion behavior of the chemical species involved. More generally, stable structures emerge as a result of the concurrence of the transformation mechanism and of chemical reactions within the phase-growth fronts.  相似文献   

14.
Dalip Kumar 《Tetrahedron letters》2004,45(14):2887-2889
Standard procedures for elaborating a quinone methide precursor for conjugation to a DNA ligand was not compatible with the presence of a vinyl group. Instead, an acrylate linker was attached by Heck coupling subsequent to o-substitution of the phenolic precursor. This transformation required protection of the phenolic group and use of ethyl acrylate rather than acrylic acid. The presence of the vinyl group also rendered the quinone methide precursor more labile to alkaline conditions than its equivalent saturated derivative and required mild conditions for coupling to the pyrrole-imidazole polyamide.  相似文献   

15.
16.
Reaction of phenylalkanenitriles and diphenylacetonitrile with aromatic nitro compounds were studied using various base-solvent systems. Four independent types of reaction: substitution of halogen, substitution of nitro group, substitution of hydride anion and electron transfer were observed. Relationship—reaction pathway—conditions have been discussed.  相似文献   

17.
Characteristic features of reactions involving esters of phosphorus-containing acids in highly organized media (micelles, liquid crystals, vesicles, and emulsions) are surveyed.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 298–312, February, 1996.  相似文献   

18.
Disproportionation of TEMPO in acids leads to the formation of an N-oxoammonium salt, which can further decompose under thermal conditions, yielding the corresponding hydroxylamine, N(2)O, CO(2) and a series of dimerisation products. Overall, acid-catalysed thermal decay of TEMPO leads to ca. 80% yield of hydroxylamine.  相似文献   

19.
For investigating the stability of C(8)-fluorine bond in 8-fluoropurine nucleosides some protected 8-fluoroguanosine, 8-fluoroinosine and 8-fluoroadenosine derivatives were prepared by direct fluorination of acetyl-protected purine nucleosides with elemental fluorine in solvents such as chloroform, acetonitrile and nitromethane. Fluorination reactions conducted in chloroform medium gave better yields of 8-fluoropurines. The fluorination yields were slightly lower when acetonitrile or nitromethane was used as solvent, but the product purification was found to be much easier. When the synthesized, protected fluoronucleosides were subjected to standard basic (NH3 in methanol or 2-propanol) and acidic (HCl in methanol) deprotection conditions relevant to nucleoside chemistry, an efficient defluorination reaction took place. The kinetics of these defluorination reactions were conveniently followed, under pseudo-first-order reaction conditions, using 19F NMR spectroscopy. 1H NMR, LC-MS and mass spectroscopy identified the products of the kinetic reaction mixtures. The defluorination reaction rate constants (kobs) in basic media depended upon the electron density at C(8) while the kobs data in acidic medium were determined by the pKa of N7. An addition-elimination based mechanism (SNAr) has been proposed for the defluorination reactions of these 8-fluoropurine nucleosides.  相似文献   

20.
The structures of TADDOL‐like α‐conidendrin‐based chiral 1,4‐diols (LIGNOLs) have been studied at molecular mechanics, Hartree‐Fock (HF)/6‐31G* and DFT/B3LYP/TZVP level of theory. The molecules included were 1,1‐diphenyl, two diastereomers of 1,1,4‐triphenyl, 1,1,4,4‐tetraphenyl, and 1,1,4,4‐tetramethyl 1,4‐diol. Several conformers of each molecule were studied thorougly also including the entropy contributions. For the triphenyl 1,4‐diols, which can form π ? π interactions between phenyl rings, the DFT optimized structures differed significantly from the HF optimized ones. A property for the most stable structures, in addition to the ability to form π ? π interactions, seemed to be the possibility to have the aliphatic six‐membered ring in a boat conformation. For all of the studied LIGNOLs some conformers were found, where the two OH groups pointed almost to the same direction. By this an intramolecular hydrogen bond can be formed between them. The bridging hydrogen atom falls at the same place as a chelate‐bonded metal ion would be situated, as in the case of the analogous molecules, TADDOLs, but only a few of these molecules would be able to work well as ligands for asymmetric catalysis. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  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