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 218 毫秒
1.
The buffered (AcOK) acetolyses of exo, (11) and endo-5, 6-dimethylidene-2-norbornyl brosylate (12) yielded exo5, 6-dimethylidene-2-norbornyl (16) and (3-methylidene-2-nortricyclyl)methyl acetates (18) . Endo-5, 6-dimethylidene-2-norbornyl (17) and 2-methylidene-3-tricyclo [3.2.1.03,6]octyl acetates (20) could not be detected. The titrimetric rate constants of the acetolysis of 11 (kt(exo)=4.49 ± 0.02) · 10?5 s?1 at 25°, ΔH=23.6 ±0.7 kcal mol?1, ΔS=0.7 ±2 calmol?1 K?1 and 12 (kt(endo)=1.9 ±0.08) · 10?9 s?1 at 25°, ΔH=27 ±1 kcal mol?1, ΔS=-8 ±2.5 calmol?1 K?1) were measured and compared with the polarimetric rate constants (kα/k(exo)=6.8 at 25°,(kα/k(exo)=1.0 at 121°) of the buffered acetolyses of the optically active brosylates (+)- 11 and (+)- 12 . Neither a common-ion (KOBs) nor a special ion effect (LiClO4) on kt(endo) could be detected, although external return might well intervene as some exo-5,6-dimethylidene-2-norbornyl tosylate (21) was formed upon solvolysis in the presence of KOTs. Acetolysis of (+)- 11 yielded completely racemized products, whereas (+)- 12 led to incomplete racemization. The buffered acetolysis of exo-(3exo-D)-5,6-dimethylidene-2-norbornyl brosylate (24) furnished (3exo-D)-( 26 :37.5%), exo-(7syn-D)-5,6-dimethylidene-2-norbornyl brosylate (27 : 37.5%) and [(5anti-D)-3-methylidene-2-nortricyclyl]methyl acetates (28 : 25 %). The acetolysis of endo-(2exo-D)-5,6-dimethylidene-2-norbornyl brosylate (25) yielded (2endo-D)-( 29 : 54%), exo-(1-D)-5,6-dimethylidene-2-norbornyl ( 30 : 36%) and [(6-D)-3-methylidene-2-nortricyclyl]methyl acetates ( 31 : 10%). Product analysis and deuterium label distribution was established by a combination of GC., 1H-NMR., 2H-{1H}-NMR. and MS. techniques. The results are rationalized by invoking anchimerically assisted ionization of the exo-brosylate 11 to symmetrical ion-pairs (cyclopropylcarbinyl cation intermediates) which undergo internal (and probably also external) return. Acetolysis of the endo-brosylate 12 is not anchimerically assisted and leads initially to non-symmetrical ion pairs. These evolve to symmetrical ion pair intermediates or, to a minor extent, are intercepted by solvent.  相似文献   

2.
The influence of the substituent at the C2 position on the hydrogen‐bonding patterns is compared for a series of five related compounds, namely (±)‐3‐exo,6‐exo‐dibromo‐5‐endo‐hydroxy‐3‐endo‐nitrobicyclo[2.2.1]heptane‐2‐exo‐carbonitrile, C8H8Br2N2O3, (II), (±)‐3‐exo,6‐exo‐dibromo‐6‐endo‐nitro‐5‐exo‐phenylbicyclo[2.2.1]heptan‐2‐endo‐ol, C13H13Br2NO3, (III), (±)‐methyl 3‐exo,6‐exo‐dibromo‐5‐endo‐hydroxy‐3‐endo‐nitrobicyclo[2.2.1]heptane‐2‐exo‐carboxylate, C9H11Br2NO5, (IV), (±)‐methyl 3‐exo,6‐exo‐dibromo‐7‐diphenylmethylidene‐5‐endo‐hydroxy‐3‐endo‐nitrobicyclo[2.2.1]heptane‐2‐exo‐carboxylate, C22H19Br2NO5, (V), and (±)‐methyl 3‐exo,6‐exo‐dibromo‐5‐endo‐hydroxy‐3‐endo‐nitro‐7‐oxabicyclo[2.2.1]heptane‐2‐exo‐carboxylate, C8H9Br2NO6, (VI). The hydrogen‐bonding motif in all five compounds is a chain, formed by O—H...O hydrogen bonds in (III), (IV), (V) and (VI), and by O—H...N hydrogen bonds in (II). All compounds except (III) contain a number of Br...Br and Br...O halogen bonds that connect the chains to each other to form two‐dimensional sheets or three‐dimensional networks. None of the compounds features intramolecular hydrogen bonding between the alcohol and nitro functional groups, as was found in the related compound (±)‐methyl 3‐exo,6‐exo‐dichloro‐5‐endo‐hydroxy‐3‐endo‐nitrobicyclo[2.2.1]heptane‐2‐exo‐carboxylate, (I) [Boeyens, Denner & Michael (1984b). J. Chem. Soc. Perkin Trans. 2, pp. 767–770]. The crystal structure of (V) exhibits whole‐molecule disorder.  相似文献   

3.
Low temperature (?130° to ?110°) addition of exo-norborn-5-en-2-ol ( 7 ) to excess HSO3F in SO2CIF yielded a mixture of exo-5-(fluorosulfonyloxy)-exo-2- and endo-2-norbornylhydroxonium ions ( 9+10 ) under kinetic control that was different from the mixture of 9+10 obtained by addition of endo-norborn-5-en-2-ol ( 8 ) to HSO3F under kinetic control. These mixtures differed from the mixture of 9+10 observed at higher temperature (?80° to ?60°) (thermodynamic control). Addition of 3-nortricyclanol ( 23 ) or exo-2, 3-epoxynorbornane ( 24 ) to HSO3F at -?120° ± 10° yielded a mixture containing the exo-2-(fluorosulfonyloxy)-anti-7- and syn-7-norbornylhydroxonium ions ( 26+27 ) as major adducts. Qualitative rates of the isomerization of 26+27 to the more stable ions 9+10 and of the isomerization 9 ? 10 were evaluated. The solvolysis of 9+10 in HSO3F yielded the exo-2, exo-5- and exo-2, endo-5-norbornanediyl bis (fluorosulfates) ( 21+22 ). Norbornadiene and quadricyclane added 2 equivalents of HSO3F and furnished kinetically a mixture of exo-2, anti-7- and exo-2, syn-7-norbornanediyl bis (fluorosulfates) ( 36+37 ) as major adducts. The latter 36+37 were isomerized into a kinetic mixture of the more stable isomers 21+22 . The rates of these isomerizations were compared. The use of DSO3F and (exo-2-D)-norborn-5-en-2-ol ( 15 ) confirmed that heterolyses of the fluorosulfates were responsible for the observed isomerization; elimination-addition processes occurred but much more slowly. The results are interpreted in terms of substituted classical and σ-bridged secondary 2-norbornyl cation intermediates. It appears that the electron withdrawing substituents FSO3 and H2O+ (HO) destabilize the σ-bridged 2-norbornyl cation more at C(5) than C(7). If the σ-bridged ions 5-Z substituted at C(5) by Z = FSO3 or H2O+ (HO) are transition states in the isomerization of the corresponding classical ions 3-Z, 4-Z , the free enthalpy difference between the ‘non-classical’ σ-bridged ion and the classical ions is not higher than the energy barrier to the quenching of the latter intermediates by FSO in HSO3F/SO2CIF.  相似文献   

4.
The solvolysis rates and products of the tertiary 2-methyl-2-exo- and -2-endo-norbornyl 2,4-dinitrophenyl ethers 1 and 2 , (X = 2,4-(NO2)22C6H3O) have been determined. The different sensitivities of the rates of these ethers to the inductive effect of substituents at C(6) indicate that graded bridging of C(2) by C(6) occurs in the ionization of the exo-ethers 1 , not, however, in the ionization of the endo-ethers 2. In both cases hydrolysis leads to 2-methyl-2-exo-norbornanols only. Consequently, substitution takes place with retention at C(2) in the exo-series 1 and with inversion at C(2) in the endo-series 2. It is concluded that stereoelectronic and polar effects, rather than steric bulk effects, determine the high exo/endo rate ratios of the parent norbornyl derivatives 1a and 2a .  相似文献   

5.
A series of exo-olefin compounds ((CH3)2C(PhY)−CH2C(=CH2)PhY) were prepared by selective cationic dimerization of α-methylstyrene (αMS) derivatives (CH2=C(CH3)PhY) with p-toluenesulfonic acid (TsOH) via β-C−H scission. They were subsequently used as reversible chain transfer agents for sulfur-free cationic RAFT polymerization of αMS via β-C−C scission in the presence of Lewis acid catalysts such as SnCl4. In particular, exo-olefin compounds with electron-donating substituents, such as a 4-MeO group (Y) on the aromatic ring, worked as efficient cationic RAFT agents for αMS to produce poly(αMS) with controlled molecular weights and exo-olefin terminals. Other exo-olefin compounds (R−CH2C(=CH2)(4-MeOPh)) with various R groups were prepared by different methods to examine the effects of R groups on the cationic RAFT polymerization. A sulfur-free cationic RAFT polymerization also proceeded for isobutylene (IB) with the exo-olefin αMS dimer ((CH3)2C(Ph)−CH2C(=CH2)Ph). Furthermore, telechelic poly(IB) with exo-olefins at both terminals was obtained with a bifunctional RAFT agent containing two exo-olefins. Finally, block copolymers of αMS and methyl methacrylate (MMA) were prepared via mechanistic transformation from cationic to radical RAFT polymerization using exo-olefin terminals containing 4-MeOPh groups as common sulfur-free RAFT groups for both cationic and radical polymerizations.  相似文献   

6.
The addition copolymerization of norbornene (NB) with functionalized monomers can lead to the modification of physical properties of poly(NB). Herein, the synthesis of new copolymer of NB with exo-norbornene lactone (exo-NBL) is reported. The copolymerization proceeded by four Pd catalytic systems, and of these, Pd(allyl)IDippCl/AgSbF6 (IDipp = 1,3-bis[2,6-diisopropylphenyl]imidazolin-2-ylidene)) was the most effective for the incorporation of exo-NBL. Specifically, the copolymerization with exo-NBL/NB feed ratio of 50/50 at r.t. by 0.1 mol% of the Pd catalyst produced poly(NB-co-exo-NBL) with Mn of 87,000, and Mw/Mn of 1.2 in 40% yield, incorporating exo-NBL of 18 mol%. The time–conversion plots and 1H diffusion ordered spectroscopy (DOSY) NMR analysis of the copolymer suggest that it has a random sequence. In contrast, no copolymer was formed from endo-NBL. This is because of steric hindrance of the endo-lactone moiety by considering the (co)polymerization of endo-5-norbornene-2-carboxylic acid methyl ester (endo-NBCO2Me). The incorporation of exo-NBL improves the solubility of poly(NB-co-exo-NBL) in several chlorinated solvents and gives high thermal stability with 10% weight loss at a temperature of more than 400°C. Two amorphous halos corresponding to intra- and interchain distances were observed in the WAXD patterns, allowing to calculate d-spacing values, which are higher than that of poly(NB).  相似文献   

7.
Cycloaliphatic polyolefins with functional groups were prepared by the Pd(II)-catalyzed addition polymerization of norbornene derivatives. Homo- and copolymers containing repeating units based on bicyclo[2.2.1] hept-5-en-2-ylmethyl decanoate (endo/exo-ratio = 80/20), bicyclo[2.2.1]hept-5-ene-2-carboxylic acid methyl ester (exo/endo = 80/20), bicyclo[2.2.1]hept-5-ene-2-methanol (endo/exo = 80/20), and bicyclo[2.2.1]hept-5-ene-2-carboxylic acid (100% endo) were prepared in 49–99% yields with {(η3-allyl)Pd(BF4)} and {(η3-allyl)Pd(SbF6)} as catalysts. The catalyst containing the hexafluoroantimonate ion was slightly more active than the tetrafluoroborate based Pd-complex.  相似文献   

8.
The Synthesis and Hydrolysis of 6-exo-Substituted 2-Methyl-2-exo-norbornyl and 2-Methyl-2-endo-norbornyl 2,4-Dinitrophenyl Ethers The synthesis of the title compounds and their hydrolysis products in aqueous dioxane are described. Upon hydrolysis, the 2-exo-ethers 1 (X=N2phO) as well as the 2-endo-ethers 2 (X=N2phO) yield the corresponding 2-methyl-2-exo-norbornanols 3 only. Therefore, the 2-exo-ethers react with retention of configuration at C(2), the 2-endo-ethers 2 with inversion at C(2).  相似文献   

9.
New alicyclic esters were synthesized by addition at heating of aliphatic monocarboxylic saturated acids C1-C4 to exo-5-phenyl-, exo-5-cyclohexyl-, and exo-5-(cyclohex-3-enyl)bicyclo[2.2.1]hept-2-enes. Among the esters obtained the acetates has more pleasant odor with fruit tint, and they may be used as a component of synthetic perfumes.  相似文献   

10.
The reaction of [RuCl2(PPh3)3] and closo-[B10H10]2? with p-IPhCOOH in CH2Cl2 solution affords two para-iodobenzoate exo-cyclized 11-vertex closo-ruthenaborane clusters [(PPh3)(p-IPhCO2)2RuB10H8] (1) and [(PPh3)2ClRu(PPh3)(p-IPhCO2)RuB10H9]?···?CH2Cl2 (2) that have been characterized by elemental analysis, FT-IR, 1H and 13C?NMR spectra and single-crystal X-ray diffraction analysis. Both clusters are based on a closo-type C 2 v 1?:?2?:?4?:?2?:?2 RuB10 stack with the metal occupying the unique six-connected apical position. In 1, the metal center has three exo-polyhedral ligands: one triphenylphosphine and two native oxygen atoms of para-iodobenzoates. The other oxygen atoms of two para-iodobenzoates are additionally bonded to B(2) and B(3) atoms respectively, resulting in two exo-cyclic five-membered Ru–O–C–O–B rings and engendering a symmetrical conformation. For 2, the metal center also has three exo-polyhedral ligands, one triphenylphosphine and one para-iodobenzoate to form one exo-cyclic five-membered Ru–O–C–O–B ring. There is an additional exo-polyhedral ruthenium atom bonding to the {RuB10} center via a {Ru–Ru} linkage and two {RuH μ B} bridges resulting in one closo distorted exo-polyhedral Ru(1)–Ru(2)–B(2)–B(4) tetrahedron.  相似文献   

11.
Reactions with Organophosphorus Compounds, 49. Synthesis and 1H NMR Spectra of (3-Acylbicyclo[2.2.1]hept-5-en-2-yl)phosphonates Reaction of the (E)-(β-acylvinyl)phosphonates 1 with cyclopentadiene yields the isomeric norbornylphosphonates 2 (endo-acyl, exo-P) and 3 (exo-acyl, endo-P) in a 7:3 ratio. With 1,3-cyclohexadiene the corresponding bicyclooctenyl derivatives 7 and 8 are obtained from 1a . The (Z)-phosphinylacrylate 4 gives with cyclopentadiene the isomers 5 (exo-CO2Me, exo-P) and 6 (endo-CO2Me, endo-P) in nearly equal amounts. The configuration of the cycloadducts has been proved by 1H NMR spectroscopy.  相似文献   

12.
The binuclear metal complex [Cu(μ-exoO2)cyclamCu(bpy)](ClO4)2·H2O (bpy?=?2,2′-bipyridine and (exoO2)cyclam?=?1,4,8,11-tetraazacyclotradecanne-2,3-dione) has been synthesized and characterized by single-crystal X-ray analysis and spectroscopic and magnetic measurements. The structure consists of homobinuclear [Cu(μ-exoO2)cyclamCu(bpy)]2+ cations, a weakly coordinated water molecule and perchlorate ions. In each binuclear unit, Cu1, coordinated by four nitrogen atoms of the macrocyclic organic ligand is connected to Cu2 via the exo-cis oxygen atoms of the macrocyclic ligand with Cu···Cu separations of 5.151?Å; Cu2 assumes square-pyramidal geometry. Magnetic properties measured at 2–300?K show antiferromagnetic exchange between adjacent copper(II) ions.  相似文献   

13.
The mass spectra of exo-2-norbornyl chloride, 1- and 2-methyl exo-2-norbornyl, exo-camphenilyl, apoisobornyl, bornyl and isobornyl chloride, and camphene hydrochloride, α- and β-fenchyl chloride and fenchene hydrochloride, and exo-isofenchyl chloride and 2,5,5-trimethyl exo-2-norbornylchloride, and camphene and α-fenchene have been examined at 12 to 16 and 80° eV and at 30 to 49 and 80°, or higher temperatures. Wagner-Meerwein rearrangements occur very readily in the ion source and compounds related by these rearrangements give very similar fragmentation patterns. Thermal decompositions are important with the tertiary chlorides especially at higher source temperatures. The rates of methanolysis of some of these chlorides were measured.  相似文献   

14.
The difficult Diels-Alder additions of α-acetoxy- and α-chloroacrylonitrile to furan can be run at 20–35° and atmospheric pressure in the presence of CuCl. Cu(BF4) · 6 H2O, Cu(OOCCH3)2 · H2O or cupric tartrate · 3H2O. Under kinetic control, the exo-carbonitrile adducts 2 and 8 , respectively, are favoured. Saponification of the 2endo-acetoxy-7-oxabicyclo[2.2.1]hept-5-ene-2exo-carbonitrile ( 2 ) furnished the 7-oxabicyclo[2.2.1]hept-5-en-2-one ( 4 ). Basic hydrolysis of the adducts ( 8 + 9 ) of α-chloroacrylonitrile to furan and its 5exo, 6exo-isopropylidenedioxy derivatives did not give the corresponding ketones, the carboxamides 14 + 15 and 16 + 17 , respectively, were isolated.  相似文献   

15.
A series of five compounds containing the bicyclo[3.3.0]octa‐2,6‐diene skeleton are described, namely tetramethyl cis,cis‐3,7‐dihydroxybicyclo[3.3.0]octa‐2,6‐diene‐2,4‐exo,6,8‐exo‐tetracarboxylate, C16H18O10, (I), tetramethyl cis,cis‐3,7‐dihydroxy‐1,5‐dimethylbicyclo[3.3.0]octa‐2,6‐diene‐2,4‐exo,6,8‐exo‐tetracarboxylate, C18H22O10, (II), tetramethyl cis,cis‐3,7‐dimethoxybicyclo[3.3.0]octa‐2,6‐diene‐2,4‐exo,6,8‐exo‐tetracarboxylate, C18H22O10, (III), tetramethyl cis,cis‐3,7‐dimethoxy‐1,5‐dimethylbicyclo[3.3.0]octa‐2,6‐diene‐2,4‐exo,6,8‐exo‐tetracarboxylate, C20H26O10, (IV), and tetramethyl cis,cis‐3,7‐diacetoxybicyclo[3.3.0]octa‐2,6‐diene‐2,4‐exo,6,8‐exo‐tetracarboxylate, C20H22O12, (V). The bicyclic core is substituted in all cases at positions 2, 4, 6 and 8 with methoxycarbonyl groups and additionally at positions 3 and 7 with hydroxy [in (I) and (II)], methoxy [in (III) and (IV)] or acetoxy [in (V)] groups. The conformations of the methoxycarbonyl groups at positions 2 and 4 are exo for all five compounds. Each C5 ring of the bicyclic skeleton is almost planar, but the rings are not coplanar, with dihedral angles of 54.93 (7), 69.85 (5), 64.07 (4), 80.74 (5) and 66.91 (7)° for (I)–(V), respectively, and the bicyclooctadiene system adopts a butterfly‐like conformation. Strong intramolecular hydrogen bonds exist between the –OH and C=O groups in (I) and (II), with O...O distances of 2.660 (2) and 2.672 (2) Å in (I), and 2.653 (2) and 2.635 (2) Å in (II). The molecular packing is stabilized by weaker C—H...O(=C) interactions, leading to dimers in (I)–(III) and to a chain structure in (V). The structure series presented in this article shows how the geometry of the cycloocta‐2,6‐diene skeleton changes upon substitution in different positions and, consequently, how the packing is modified, although the intermolecular interactions are basically the same across the series.  相似文献   

16.
Long range proton–proton coupling constants between H3exo and H5exo have been measured in a series of 4-substituted bornanones. Correlation between the magnitudes of 4J3exo–5exo and the inductive linear free energy σI has been obtained. 4-Substituted bornanone nitrimines exhibit diminished values of 4J3exo–5exo with respect to the parent ketone.  相似文献   

17.
Chemical ionization mass spectra of exo- and endo-2-norbornanols and their phenylurethane derivatives have been obtained with several reactant ions. Small differences are noted in the abundances of norbornyl and [M+H]+ ions for the phenylurethane derivatives: more norbornyl ions with the exo compounds. Relative rate constants for decomposition of [M+H]+ ions, kexo/kendo ? 1-2. No evidence was found for s?-participation in the decomposition of these ions. The i-C4H10 chemical ionization spectrum of endo-2-norbornanol contains a much greater abundance of [M-H]+ ions than the i-C4H10 chemical ionization spectrum of exo-2-norbornanol. This difference presumably results from steric hindrance toward attack of the endo hydrogen.  相似文献   

18.
LI  Xiao-Liu XU  Xiao-Ming TIAN  Jun LI  Ying-Xia 《中国化学》2005,23(11):1564-1568
Novel β-C-disaccharides containing a three carbon linkage using exo-glycal as the precursor were prepared stereoselectively. The synthesis was achieved by the tandem reactions of the stereoselective hydroboration of exo-glycal and Suzuki cross-coupling reaction with an iodovinyl sugar, and followed by hydrogenated deprotection under the catalysis of Pd(OH)2/C.  相似文献   

19.
The 1H NMR spectra of 2-exo-hydroxymethyl-3-endo-methylnorbornene and the corresponding 2-endo-3-exo isomer have been completely assigned with the aid of lanthanide reagents. This enabled a full analysis of the unusual spectrum of 2-exo-hydroxy-methyl-3-endo-methylnorbornene to be undertaken, confirming the proposed structure. The ΔEu values for 2-exo-hydroxymethyl-3-endo-methylnorbornene and the 2-endo-3-exo-isomer have been used to test the effect of rotational averaging on the calculated pseudo contact shifts. Good agreement is obtained for a dynamic model in which the Eu atom exchanges between two sites on the oxygen atom of the OH bond, and in which the rotational equilibrium about the CH? CH2OH bond is explicitly considered.  相似文献   

20.
An example of an octaphosphane of type R2P8 (R=(DDP)Ga) was isolated by treatment of cage compound (DDP)GaP4 ( 2 , DDP=(2,6-diisopropylphenyl)(4-((2,6-diisopropylphenyl)imino)pent-2-en-2-yl)amide) with (C6F5)2PBr. The initially formed endo-exo butterfly shaped pentaphosphane 7 rapidly rearranges to the more stable exoexo isomer 8 , which undergoes dimerization to decaphosphane 11 . Compound 11 unexpectedly eliminates tetraaryldiphosphane 13 to give tetracyclo[3.3.0.02,7.04,6]octaphosphane [(DDP)GaBr]2P8 ( 12 ). The reaction steps were confirmed by crystal structure analysis of the key intermediates and supported by kinetic studies using NMR techniques.  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