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 359 毫秒
1.
A strong secondary isotope effect is observed in the preferred loss of methyl vs. trideutero-methyl from the molecular ions of appropriately labelled 4-t-butylpyridine and t-butylbenzene decomposing in the first and second field free regions of a double focusing mass spectrometer. This has been rationalised by invoking the theory of radiationless transitions2, which can account for the higher population of activated states responsible for loss of methyl vs. that for trideuteromethyl. 13C-Labelling at the central carbon atom of the t-butyl group indicates that the [M – methyl]+ ions, decomposing further by elimination of ethylene, cannot be represented exclusively by a pyridylated (or phenylated) cyclopropane ion if present at all. It is concluded that ions with structures generated by 1,2-hydrogen-, 1,2-pyridyl- (or 1,2-phenyl-) and 1,2-methyl shifts must also play a role. D-labelling further shows an extensive randomisation of side-chain hydrogen atoms in the [M-methyl]+ ions of 4-t-butylbenzene; in this case, however, the expelled ethylene also contains ring hydrogen atoms (≤2). Presumably this is caused by exchange between the side-chain and ortho-hydrogen atoms in the initially generated phenyldimethylcarbinyl carbenium ion.  相似文献   

2.
In this study, carried out with 16 compounds, the chair form with an axial S?O group (CA) is found, in the absence of C-4, C-5 and C-6 substituents, to be the most stable (ΔG>8,4 kJ mole?1), as previously reported for analogous cyclic sulfites. When methyl or tert-butyl substituents are present on the 4 and 6 carbon atoms, the conformation of the ring depends on their respective orientation towards the S?O group, and on the nature of the substituent of the nitrogen atom. For the trans isomers, the conformation remains anancomeric chair (CA) except when important gauche interactions exist: thus the strong Me/tBu gauche interaction in the 3-tert-butyl-4-t-methyl-2-r-oxo-1,2,3-oxathiazan induces a twist form with a 3,6-axis and an axial S?O group (CNA). When the 4- or 6-substituent is cis, the conformation of the sulfinamate may be either a chair form with an axial S?O group (CA), if the 4-substituent is a methyl, (even with a tert-butyl group in the 3-position which would be in the axial orientation) or a twist form with a 1,4-axis and an axial S?O group (COA) if the 4-substituent is a tert-butyl. Unlike cyclic sulfites, the equatorial SO chair form (CE) and the twist forms with a 2,3-axis and an isoclinal S?O group (CS, CS′) are rarely involved.  相似文献   

3.
In the benzilic acid type rearrangement of t-butyl α,β-dioxobutyrate (VII) the intact t-butoxycarbonyl group is shifted to the β-carbonyl carbon atom.  相似文献   

4.
UV.-Irradiation of methyl t-butyl ketone, di-t-butyl ketone, pivalaldehyde and isobutyraldehyde in benzene solution leads to chemically induced dynamic nuclear polarization (CIDNP.) of the parent compounds and of various reaction products. CIDNP.-effects and product distributions establish type I α-cleavage predominantly of triplet state molecules as the major primary photochemical step for all the four α-branched carbonyl compounds. In chlorinated solvents singlet reactions interfere with the triplet processes.  相似文献   

5.
Selective n → π* excitation of the α,β-unsaturated enone 1 in hydrocarbon solvents resulted in a deconjugation reaction to 3 , reminiscent of results previously reported for similar systems [2], whereas the photoreactivity of 1 in alcohol solvents at wavelengths >3400 Å was so small that only product 4 has been identified as yet. Excitation of the π → π* transition of compound 1 at 2537 Å initiated additional phototransformations which could not be effected by irradiation in the first absorption band. The [4.4.3]-12-oxapropellane derivative 2 was identified as one of the two new major photo-isomers. A 6:8 mixture of products 2 and 3 , plus about 1 part of an isomer of still unknown structure (see however, the Addendum), were readily formed in hydrocarbon solvents, and a 1:10 ratio of 2 and the unknown product was obtained in methanol. Abstraction of a methoxyl hydrogen by the ketone oxygen is proposed to account for the primary photochemical step in the cyclization to 2 . A hydrogen-deuterium isotope effect of 2.7 was observed in a competitive experiment using 1 and 1-d 6. 34% of one deuterium atom were exchanged for hydrogen when 1-d 6 was photolyzed to 2-d 6 in t-butyl alcohol, which suggests an intermediate of type a in the pathway 1 → 2 possessing a readily exchangeable proton. Steric considerations would require a strongly distorted, non-planar excited-state geometry of the enone group of 1 for the oxygen to approach a methoxyl hydrogen. The transformation 1 → 2 represents a novel reaction type in photochemical processes of conjugated enones which are specifically induced by π → π* excitation only.  相似文献   

6.
Substited pivalanilides give rise to fragments corresponding to anilies on electron-impact, which is due to the migration of a hydrogen atom from the t-butyl group to the nitrogen. The Hammett correlation indicates that the formation of this fragment is favoured by electron-donating groups. The other interesting feature of these spectra is the migration of the aryl group to the t-butyl carbon with the elimination of the neutral ketemine.  相似文献   

7.
The mass spectra of six aziridinones (α-lactams) are presented. The six aziridinones exhibit the same general fragmentation pattern, the primary processes being loss of the N-t-butyl group, loss of carbon monoxide and loss of t-butyl isocyanate. Several additional rearrangement or fragmentation processes are observed with specific α-lactams.  相似文献   

8.
To elucidate the effect of the introduction of a methyl group in the β-position of a vinyl monomer, propenyl alkyl ethers were copolymerized with vinyl ethers having the same alkoxy group. Propenyl alkyl ethers with an unbranched alkoxy group (ethyl or n-butyl propenyl ether) were more reactive than the corresponding vinyl ethers. This behavior is quite different from that of β-methylstyrene derivatives. However, propenyl alkyl ethers with branched alkoxy groups at the α carbon atom (isopropyl or tert-butyl propenyl ether) were less reactive than the corresponding vinyl ethers. Also, cis- isomers were more reactive than the trans isomers, regardless of the kind of alkoxy group and the polarity of the solvent.  相似文献   

9.
Alternating copolymers of styrene with β-carboalkoxy-substituted t-butyl α-cyanoacrylates undergo facile deesterification at 150–190°C, about 60°C below the deprotection temperature of poly(t-butyl methacrylate), and decarboxylation at 170–200°C. When the β-substituent is a methyl ester, the two events are clearly separated, with the deesterification occurring at a maximum rate at 165°C and decarboxylation at 193°C. Anhydride formation is negligible in this case. The copolymer with t-butyl cyanofumarate exhibits simultaneous deesterification and decarboxylation events at 180°C with concomitant minor dehydration.  相似文献   

10.
2-Lithio-1-triphenylmethylimidazoles react with t-butyl halogenoacetates to give a variety of products, the nature of which is cleanly determined by the halogen atom. With chloroacetate the products are chloromethyl ketones, while bromacetate gives di-t-butyl imidazolesuccinates, and iodoacetate yields iodoimidazoles. In each case 50% of the parent triphenylmethylimidazole is recovered from the reaction. When the triphenylmethyl substituent is replaced by the N,N-dimethylsulfamoyl group, reaction with bro-moacetate is suppressed, but t-butyl chloroacetate and iodoacetate again give chloroketones and aryl iodides respectively.  相似文献   

11.
Radical polymerizations and copolymerizations were carried out in the presence of phenyl tert-butyl nitrone(PBN) and tert-nitrosobutane (t-BuNO), and the structure of the spin adduct formed was investigated by ESR spectroscopy. PBN adducts gave the same ESR pattern, and the variation of the splitting constant was not large enough to warrant its use for the structure assignment. Therefore, the subsequent trapping experiments were performed with t-BuNO. The polymerization mixture of representative monomers showed esr patterns that are indicative of the propagating radical being trapped. These trapped radicals were not necessarily very stable and, in most cases, disappeared after long reaction periods. In the case of α-methyl substituted monomers, additional nine-line spectra were observed which were attributed to trapping of the radical species formed by hydrogen abstraction from the α-methyl group. The tert-butyl radical which was formed by decomposition of t-BuNo was probably responsible for the hydrogen abstraction. In the case of styrene, methyl acrylate, and methyl methacrylate, characteristic ESR patterns of the propagating radicals were observed with polymers which were prepared in the presence of t-BuNO and purified by reprecipitation. Simultaneous trapping of different propagating radicals was attempted in several copolymerization systems. However, this was generally unsuccessful, because of the large difference in reactivities of the propagating radical with t-BuNo.  相似文献   

12.
Synthesis of sugar phosphate derivatives by means of phosphite triester method is described. Seven glucose phosphotriester derivatives have been prepared, i.e. dimethyl, methyl n-propyl, and methyl isopropyl (1, 2:5, 6-di-O-isopropylidene-α-D-glucofuranose-3-) phosphate (5, 7 and 8); methyl bis-(1, 2:5, 6-di-O-isopropylidene-α-D-glucofuranose-3-) phosphate (6); methyl bis-(1, 2, 3, 4-tetra-O-acetyl-β-D-glucopyranose-6-) phosphate (9); methyl bis-(1, 2-O-isopropylidene-3,5-O-benzylidene-α-D-glucofuranose-6-) phosphate (10); and methyl (1, 2, 5, 6-di-O-iso-propylidene-α-D-glucofuranose-3-) (1, 2, 3, 4-tetra-O-acetyl-β-D-glucopyranose-6-) [phosphate (11). The results of the displacement of second chlorine atom of the reagent by different alcohols showed that methanol, n-propanol, isopropanol and as well as the glucose derivatives reacted normally to give the expected phosphite esters which yield the expected phosphate products after oxidation, but not the t-butanol. Removal of methyl group from a phosphotriester linkage can be easily achieved by the action of t-butyl amine and thus, t-butyl ammonium bis-(1, 2:5, 6-di-O-isopropylidene-α-D-glucofuranose-3-) phosphate t-butyl amine salt (12) has been obtained from its parent phosphotriester in nearly quantitative yield. The mass spectra data of di-O-isopropylideneglucose phosphate reveals that the cleavage of these compounds follows a general pattern and can be used for their characterization.  相似文献   

13.
Vinylidene chloride polymers containing comonomer units capable of consuming evolved hydrogen chloride to expose good radical-scavenging sites might be expected to display greater thermal stability than similar polymers containing simple alkyl acrylates as comonomer. Incorporation of a comonomer containing the phenyl t-butyl carbonate moiety into a vinylidene chloride polymer has the potential to afford a polymer with pendant groups which might interact with hydrogen chloride to expose phenolic groups. Copolymers of vinylidene chloride with [4-(t-butoxycarbonyloxy)phenyl]methyl acrylate have been prepared, characterized, and subjected to thermal degradation. The degradation has been characterized by thermal and spectroscopic techniques. The degradation of vinylidene chloride/[4-(t-butoxycarbonyloxy)phenyl]methyl acrylate copolymers is much more facile than the same process for similar copolymers containing either [4-(isobutoxycarbonyloxy)phenyl]methyl acrylate or methyl acrylate, a simple alkyl acrylate, as comonomer. During copolymer degradation, [4-(t-butoxycarbonyloxy) phenylmethyl acrylate units are apparently converted to acrylic acid units by extensive fragmentation of the sidechain. Thus, the phenyl t-butyl carbonate moiety does function as a labile acid-sensitive pendant group but its decomposition in this instance leads to the generation of a phenoxybenzyl carboxylate capable of further fragmentation.  相似文献   

14.
The X‐ray crystal structures of cis‐ and trans‐1‐(indol‐3‐yl)‐4‐methyl cyclohexane and its congeners with stepwise fluorination of the methyl group are reported. The trans‐configured compounds adopted diequatorial conformations, whereas the cis analogues adopted regular cyclohexane chair conformations with the methyl group preferentially assuming the axial position, even in the case of the CF3 group. Surprisingly, although the axial CF3 derivative displayed distinct valence deformations in the cyclohexane moiety, the observed structural changes were relatively modest. The cis derivatives with axial mono‐ and difluorinated methyl groups exhibited conformational disorder in the crystals with significant population levels for the staggered conformations that had one fluorine atom in the endo position; their respective trans counterparts adopted unique conformations, but again with one fluorine atom in the endo position. Theoretical calculations for a series of cis‐ and trans‐1,4‐dimethyl cyclohexane model compounds with stepwise fluorination of one equatorial or axial methyl group reproduced the experimentally observed structural response patterns very well, reproduced the experimentally determined nonlinear correlation of the axial–equatorial energy difference with the degree of methyl fluorination in a satisfactory manner, and provided further insights into important conformational aspects of partially fluorinated methyl groups.  相似文献   

15.
One new and one established functional cyclooctene were prepared and (co)polymerized using ring-opening metathesis polymerization. The resulting polymers were hydrogenated to yield the corresponding functional polyolefins that were structurally equivalent to copolymers of ethylene and either methyl methacrylate, t-butyl acrylate, or acrylic acid after deprotection. The copolymers that incorporate methyl methacrylate into the backbone were used as compatibilizers for poly(methyl methacrylate)/polyethylene blends. The copolymers that incorporate t-butyl acrylate into the backbone yielded elastomers that could be thermally crosslinked. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 3117–3126  相似文献   

16.
The reaction of N-phenylmaleimide, 4a , with sodium 2,6-di-t-butylphenolate, 5a , in dimethylsulfoxide (DMSO) resulted a complex oligomeric mixture. The dimer 8 was isolated from the reaction of the N-alkyl-maleimide 4b with 5a in DMSO. The reaction of 4a with 5a in tetrahydrofuran (THF), which is an aprotic solvent that is known to promote ion pairing, resulted in the isolation of a low yield of 6a . The reaction of 4a with 5a in the hydrogen-bonding solvent t-butyl alcohol gave 6a in slightly higher yield. The N-alkylpyrrolidine-2,5-diones 6c-f were obtained by the reaction of the maleimides 4c-e with the corresponding sodium phenolate 5a-b in t-butyl alcohol reaction medium. The isolated yield of product increased with the size of the N-alkyl substituent of the maleimide. Surprisingly, the reaction of the 2,6-dimethylphenolate 5c with 4d led to the isolation of the dimer 10 with the formation of a quaternary carbon atom. The yield of 6a was observed to counterion dependent, increasing in the order Na+ < Li+ < MgBr+ in t-butyl alcohol. The bismaleimides 12 and 14 were obtained by the reaction of either 11 or 13 with 5d in THF.  相似文献   

17.
Isotopic labelling and chemical substitution support the proposition that the skeletal rearrangement for water loss from molecular protonated ions of t-butoxycyclohexane involves competition between three reaction pathways. The principal reaction pathway (83%) involves migration of the t-butyl group to the 2-(6-) position of the cyclohexyl ring with reciprocal hydrogen transfer. A second reaction pathway (12%) involves ring contraction followed by reciprocal exchange of the t-butyl group with the 2-(5-) hydrogen atom of the nascent cyclopentyl ring. The third reaction pathway (5%) involves rearrangement of a proton-bound complex to permit ipso attack by isobutene. Stereospecific substitutions indicate that the principal reaction pathway is susceptible to 1,3-diaxial interactions.  相似文献   

18.
Tautomerism of benzaurins and hydration are studied. 1H and 19F chemical shifts have been determined for a number of substituted 4-hydroxyphenyl-diphenyl carbinols containing fluorine in a 3-, 3*- or 4*-position, and for similar compounds containing additional methyl groups in a position of 3, 3** or 4**. The same data have been obtained for the fuchsones prepared by dehydration of the above carbinols. On this basis chemical shifts of fluorine in different positions have been evaluated as a monitor of the transformation of 4-hydroxyphenyl group to the semiquinone moiety. The 19F NMR can be used to monitor the transformation of 4**-fluorobenzaurin and the related 3,3*-disubstituted and 3,3*,5,5*-tetramethylsubstituted compounds to the corresponding carbinols due to the addition of a water molecule and to study the tautomerism of the two latter benzaurins as well as that of 3,3*,4**trifluorobenzaurin. Furthermore, fluorine and methyl group chemical shifts are sensitive to syn-anti-isomerism in substituted fuchsones.  相似文献   

19.
By quantum-chemical calculations at the M06-2X/aug-cc-pVTZ level of theory geometrical parameters, dipole moments, polarizabilities, first hyperpolarizabilities and relative energies of the axial and equatorial conformers in gaseous phase were determined for 1-cyano-1-silacyclohexane, 1-tert-butyl-1-silacyclohexane, 1-tert-butyl-1-cyano-1-silacyclohexane, and 1-fluoro-1-silacyclohexane. For the cyano group and fluorine atom the axial position is more preferable whereas for tert-butyl group, equatorial one. Polarizabilities of conformers are similar but optical anisotropy of equatorial conformers of C5H10SiHCN and C5H10SiH(t-Bu) molecules is much larger than that of axial conformers. Upon substitution in nitriles of C1 atom by Si atom the hyperpolarizability is many times increased.  相似文献   

20.
We demonstrate herein that wild‐type cytochrome P450 BM3 can recognize non‐natural substrates, such as fluorinated C12–C15 chain‐length fatty acids, and show better catalysis for their efficient conversion. Although the binding affinities for fluorinated substrates in the P450 BM3 pocket are marginally lower than those for non‐fluorinated substrates, spin‐shift measurements suggest that fluoro substituents at the ω‐position can facilitate rearrangement of the dynamic structure of the bulk‐water network within the hydrophobic pocket through a micro desolvation process to expel the water ligand of the heme iron that is present in the resting state. A lowering of the Michaelis–Menten constant (Km), however, indicates that fluorinated fatty acids are indeed better substrates compared with their non‐fluorinated counterparts. An enhancement of the turnover frequencies (kcat) for electron transfer from NADPH to the heme iron and for C? H bond oxidation by compound I (Cpd I) to yield the product suggests that the activation energies associated with going from the enzyme–substrate (ES state) to the corresponding transition state (ES state) are significantly lowered for both steps in the case of the fluorinated substrates. Delicate control of the regioselectivity by the fluorinated terminal methyl groups of the C12–C15 fatty acids has been noted. Despite the fact that residues Arg47/Tyr51/Ser72 exert significant control over the hydroxylation of the subterminal carbon atoms toward the hydrocarbon tail, the fluorine substituent(s) at the ω‐position affects the regioselective hydroxylation. For substrate hydroxylation, we have found that fluorinated lauric acids probably give a better structural fit for the heme pocket than fluorinated pentadecanoic acid, even though pentadecanoic acid is by far the best substrate among the reported fatty acids. Interestingly, 12‐fluorododecanoic acid, with only one fluorine atom at the terminal methyl group, exhibits a comparable turnover frequency to that of pentadecanoic acid. Thus, fluorination of the terminal methyl group introduces additional interactions of the substrate within the hydrophobic pocket, which influence the electron transfers for both dioxygen activation and the controlled oxidation of aliphatics mediated by high‐valent oxoferryl species.  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