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 265 毫秒
1.
A regioselective Biginelli‐like reaction of alicyclic mono‐ketones, aromatic aldehydes, and urea in ionic liquid [BPY]BF4 has been investigated. The process is controlled by the size of alicyclic mono‐ketones and the steric hindrance of aromatic aldehydes. The reaction of cyclopentanone with urea and aromatic aldehydes afforded 7‐arylidene‐3,4,6,7‐tetrahydro‐4‐aryl‐1H‐cyclopenta[d]pyrimidin‐2(5H)‐ones ( 4 ). When cyclohexanone was used as the source of active methylene to react with urea and aldehydes with slight steric hindrance groups under the same condition, 8‐arylidene‐3,4,5,6,7, 8‐hexahydro‐4‐arylquinazolin‐2(1H)‐ones ( 6 ), a homologue of 4 , were yielded, whereas 4,8‐bisaryloc‐tahydro‐1H‐pyrimido[5,4‐i]‐quinazoline‐2,10(3H,11H)‐diones ( 7 ) were obtained via the simple one‐pot reaction of cyclohexanone, urea, and aromatic aldehydes with high steric hindrance groups. The possible transitional states and mechanism of the regioselective process were discussed.  相似文献   

2.
N-l-Diamantylmaleimide was synthesized by reaction of maleic anhydride with 1-aminodiamantane, followed by dehydration with acetic anhydride and sodium acetate. Poly(N-1-adamantylmaleimide) ( IIa ) and poly(N-l-diamantylmaleimide) ( IIb ) were polymerized using 2,2′-azobisisobutyronitrile (AIBN) as an initiator under different experimental conditions such as various initiator concentrations, solvents, polymerization temperatures, and polymerization times. Polymerizations of N-l-adamantylmaleimide in benzene at 60°C or in bulk gave polymers with molecular weights (2000–9500). The experimental results indicated that the propagation may be interrupted by steric hindrance of bulky and rigid substituents such as the adamantyl or diamantyl groups. In addition, the effect of chain transfer to monomer contributes to the relatively low activation energy. The glass transition temperatures of Ia and Ib were 204 and 216°C, respectively. The temperatures at 5% weight loss of the polymers IIa and IIb were above 412°C. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
The syntheses of optically active t-butyl 2-(p-tolylsulfinyl)propionate and -butyrate ( 2a and 2b , respectively) are described, and it is shown that aldol-type condensation of the corresponding enolates is limited by steric hindrance. Optically active (2a) reacted, however, in high yield with aliphatic aldehydes and lead to 80% of asymmetric induction.  相似文献   

4.
It has been established that the cyclization of -(2-carboxy-3-naphthylamino)propionic acid into N-acetyl-4-oxo-1, 2, 3, 4-tetrahydrobenzo[g]quinoline takes place with the participation of acetic anhydride and an alkali metal acetate. In the absence of the alkali metal acetate cyclization takes place in a different direction. A mechanism for this reaction has been proposed as taking place through the formation of an internal mixed anhydride, which decomposes under the reaction conditions into N-acetyl-4-oxo-1, 2, 3, 4-tetrahydrobenzo[g]-quinoline and carbon dioxide.For part V, see [1].  相似文献   

5.
Three biphenyl‐3,5‐dicarboxylic acid (H2 L ) based coordination polymers, namely, [Mn3( L )3(2,2′‐bpy)2]n ( 1 ), {[Mn( L )(phen)] · (MeOH)}n ( 2 ), and [Mn( L )(dipt)]n ( 3 ), (2,2′‐bpy = 2,2′‐bipyridine, phen = 1,10‐phenanthroline, and dipt = 2,9‐dimethyl‐1,10‐phenanthroline) were synthesized and characterized by single‐crystal X‐ray diffraction and analyses of their magnetic properties. 1 is a trinuclear manganese structure with a 2D motifs, which can join by hydrogen bond bridges to give 3D supramolecular architectures. 2 has a dinuclear center forming a 1D supramolecular ladder chain. The mononuclear complex 3 displays 1D metal‐organic chains driven by μ2‐ L linkers. Their structural differences were investigated, revealing that the influence of steric hindrance on the structures of acid‐based coordination polymers is realized through changing the N‐heterocyclic ancillaries of diverse steric hindrance. Obviously, with decreasing of the steric hindrance of the N‐donor ligand, complexes 1 – 3 show structures from 1D to 2D and mononuclear to multinuclear. Magnetic susceptibility measurements indicate that 1 and 2 have dominating antiferromagnetic couplings between metal ions, whereas compound 3 is paramagnetic.  相似文献   

6.
Three series of water‐soluble semi‐rigid thermo‐responsive polymers with well‐defined molecular weights based on mesogen‐jacketed liquid crystal polymers, poly[bis(N‐(2‐hydroxypropyl) pyrrolidone) 2‐vinylterephthalate] [P(2‐HPPVTA)], poly[bis(N‐(1‐methyl‐2‐hydroxyethyl) pyrrolidone) 2‐vinylterephthalate] [P(1‐M‐2‐HEPVTA)] and poly[bis(N‐hydroxypropyl pyrrolidone) 2‐vinylterephthalate] (PHPPVTA) have been synthesized via reversible addition‐fragmentation chain transfer polymerization. The steric hindrance effects on liquid crystalline property and thermo‐responsive behaviors of semi‐rigid water‐soluble polymers (P(2‐HPPVTA), P(1‐M‐2‐HEPVTA), and PHPPVTA) were carefully investigated. From molecular structure, the steric hindrance of P(1‐M‐2‐HEPVTA) is stronger than that of P(2‐HPPVTA). Polarized light microscope and one‐dimensional wide‐angle X‐ray diffraction revealed that both the P(2‐HPPVTA) and P(1‐M‐2‐HEPVTA) display a columnar nematic phase, indicating that the steric hindrance effect do not affect liquid crystalline behavior of the polymers. The dynamic light scattering results demonstrated that P(1‐M‐2‐HEPVTA) exhibited lower cloud point compared with that of P(2‐HPPVTA) at the same mass concentration and the same molecular weight. The more significant molecular weight and concentration dependence on cloud point have been observed in P(2‐HPPVTA) solution than in P(1‐M‐2‐HEPVTA) solution. We also discovered that the cloud points of both P(2‐HPPVTA) and P(1‐M‐2‐HEPVTA) solution are lower in D2O than in H2O. It is noted that the cloud point of PM‐2 is 9.9 °C lower in D2O than in H2O, much less pronounced than the cloud point difference of PH‐2. The differences of thermo‐responsive behaviors between P(2‐HPPVTA) and P(1‐M‐2‐HEPVTA) were resulted from the steric hindrance effect existed in their side groups. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013 , 51, 3429–3438  相似文献   

7.
The reactivity of metal-hydride complexes can be harnessed by the modification of ancillary ligands. With the aim of improving the hydride-donor ability of the key Mn−H intermediate and reducing steric hindrance, we herein report the rational design of a versatile and efficient NHC-based NNC-pincer Mn catalyst for hydrogenation reactions. This newly developed catalyst exhibited higher activity than the corresponding NNP-pincer Mn catalyst owing to its reduced steric hindrance and enhanced Mn−H σ-bonding orbital energy level through a π-antibonding interaction. Using this highly active NNC-pincer Mn catalyst, a rich array of polar unsaturated compounds (>80 examples) including esters, N-heteroarenes, amides, carbonates, and urea derivatives, were successfully hydrogenated under relatively mild conditions. This work represents a rare example of a general phosphine-free Mn-catalyzed hydrogenation system.  相似文献   

8.
Linear 1:1:1 coupling of dimethylzinc, 1,3-dienes, and carbonyl compounds in this order is facilitated by catalytic amounts of [Ni(acac)(2)] to give (E)-3-hexen-1-ols in good yields under mild conditions [Eq. (a)]. Increasing steric hindrance at the carbonyl group favors formation of the 1:2:1 adduct, and this is the sole product when the carbonyl compound is acetone. acac=acetylacetonate.  相似文献   

9.
ortho‐Methyl effects are exploited to tune steric hindrance between side‐chain N,N′‐diaryls and polycyclic dihydrodibenzo[a,c]phenazine, and in turn control the conformations of N,N′‐diphenyl‐dihydrodibenzo[a,c]phenazine (DPAC) and its ortho‐methyl derivatives M x ‐M y (x=0, 1 or 2, y=1 or 2, x and y correlate with the number of methyl groups in the ortho‐positiond of N,N′‐diphenyl). The magnitude of steric hindrance increases as x and y increase, and the V‐shaped dihydrodibenzo[a,c]phenazine skeleton is gradually tuned from a bent (DPAC) to planar ( M2‐M2 ) structure in the ground state. As a result, the relaxation of the excited‐state structure of DPAC and its numerous analogues could be mimicked by model structures M x ‐M y, demonstrating for the first time the the conformation change from bent‐to‐planar and hence a large range of energy‐gap tuning of polycyclic conjugated structures controlled by the steric hindrance.  相似文献   

10.
Abstract

Cyclic mixed anhydride, N-methyl-2-benzoyloxy-2-oxo-1,3,2-oxazaphosphorinane (1a) has been synthesised and the rate of its fragmentation involving nitrogen migration from phosphorus to carbonyl carbon has been measured. (1a) was found to be ca. 60 times less reactive than the non-cyclic, O-methyl-N,N-dimethyl analogue. The crystal and molecular structure of (1a) has been determined using x-ray diffraction. Pna21, a=22.229(6), b=7.597(2), c=7.210(2) Å; V=1217.6(6) Å3. Final R=3.08% for 1037 reflections with I(rel) > 2[sgrave]I(rel) and 157 parameters. The observed conformation of the molecule of (1a) is very different from that required for the fragmentation to occur; in order to achieve the geometry postulated for the transition state significant rotations about the P[sbnd]O and O[sbnd]C bonds would be necessary and steric hindrance by the 4,6-axial hydrogens would be expected.  相似文献   

11.
[reaction: see text] A general method for the conversion of pyrrole anions to 2-arylpyrroles has been developed. Using a palladium precatalyst and sterically demanding 2-(dialkylphosphino)biphenyl ligands, (pyrrolyl)zinc chloride may be cross-coupled with a wide range of aryl halides, including aryl chlorides and aryl bromides, at low catalyst loadings and under mild conditions. A high degree of steric hindrance is tolerated. Certain ring-substituted pyrrole anions have also been arylated with aryl bromide substrates.  相似文献   

12.
Summary: The complexes cis‐P,P′‐(η5‐cylopentadienyl)‐{5,17‐dibromo‐11,23‐bis(diphenylphosphino)‐25,26,27,28‐tetrapropoxy‐calix[4]arene}nickel(II ) tetrafluoroborate ( 1 ) and dibromo‐{5,17‐dibromo‐11,23‐bis(diphenylphosphino)‐25,26,27,28‐tetrapropoxycalix[4]arene}nickel(II ) ( 2 ), both of which contain a constrained chelating diphosphine, were evaluated for the polymerization of norbornene. Combined with methylaluminoxane, they result in remarkably active systems for the production of high‐molar‐mass vinyl‐type polynorbornene. Turnover frequencies of up to 7.5 × 105 mol(NBE) · mol(Ni)−1 · h−1 are observed. A plausible explanation for their high performances relies on a periodic P–Ni–P bite angle enlargement that temporarily increases the steric hindrance about the catalytic centre, which in turn favours the insertion steps.

Molecular structure of 2 .  相似文献   


13.
Aminomethylating 9-methyl-3, 4-benzoacridine with cycloalkylamine (pyrrolidine, piperidine, and morpholine) hydrochlorides in alcohol is described. The isomeric 9-methyl-1, 2-benzoacridine does notaminomethylate under similar conditions, apparently because the annulated benzene ring at positions 1, 2 in the acridine gives rise to steric hindrance, and the reaction product is 9--hydroxyethyl-1, 2-benzoacridine, converted by further heating to 9-vinyl-1, 2-benzoacridine.  相似文献   

14.
Diels‐Alder reactions of tetracyclone with various dienophiles under solvent free conditions were studied. In the case of cyclic dienophiles that exhibit more steric hindrance, decarbonylation of [4+2] adducts were carried out.  相似文献   

15.
Asymmetric alternating copolymerization of meso-epoxides and carbon dioxide (CO2) is an efficient route to produce isotactic polycarbonates with main-chain chirality involving two contiguous stereogenic centers. Based on multi-chiral induction and bimetallic synergistic catalysis, herein, we design novel binaphthol-linked dinuclear aluminum complexes bearing quaternary ammonium salts anchored on the ligand with multiple chiralities for asymmetric copolymerization of CO2 and meso-epoxides, affording the completely alternating polycarbonates with up to 99% enantioselectivity. Also, they are discovered to be highly active and enantioselective in catalyzing alternating copolymerization of phthalic anhydride and meso-epoxides in a controlled manner. The linked-chain length and steric hindrance at the para-position on the phenolate moieties of the ligand have significant effects on both activity and enantioselectivity. Notably, the bifunctional, multichiral complexes prove to be highly enantioselective and selective even at a low catalyst loading of 20 ppm, indicating an intramolecular synergistic effect between the tethered quaternary ammonium salts and the metallic ions.  相似文献   

16.
The treatment of N-[2-(1H-indol-3-yl)ethyl]alkanamide, 1 (1), with phosphorus oxychloride under controlled conditions gave l-alkyl-2,3,4,9-tetrahydro-1H-pyrido[3,4-b]indol-1-ol, 2 . The reaction of 2 with acetic anhydride or with methyl isocyanate at room temperature resulted in the formation of amido carbinol 3 and urea carbinol 7, respectively. The former was transformed into amido ester 4 by boiling acetic anhydride. When the reaction of 3 with acetic anhydride was carried out in the presence of excess triethylamine at 105°, C-N bond cleavage of the tetrahydropyridine ring took place with concurrent bis(N-acetylation) to give the enol ester derivative 5 . The structures of all compounds are consistent with chemical and spectral evidence.  相似文献   

17.
Photochromic acrylates containing both biphenylene and spiro-oxazine moieties with a chiral substituent and the related polymers were prepared and yielded photochromic chiral liquid crystalline systems. The photochromic acrylates containing both an undecamethylene group and a (2S, 3S)-2-chloro-3-methylpentanoyloxy group (A11SOP) or a (-)-menthoxyacetoxy group (A11SOM) gave a supercooled mesophase; the latter reflected right-handed visible light (blue colour) at room temperature. On the other hand, the photochromic acrylate containing both the (R)-(-)-2-methylpropylene and (2S, 3S)-2-chloro-3-methylpentanoyloxy groups (A3SOP) showed no mesophase. The related homopolymers, PA11SOP and PA11SOM, did not exhibit mesophases because of steric hindrance between the side groups of the polymers. However, only PA11SOM exhibited shear-induced birefringence under 100-104°C. Several copolymers consisting of the nematogenic monomer, 4-[4-(6-acryloyloxyhexyloxy)benzoyloxy]benzonitrile (A6CN), and A11SOP or A11SOM possessed a smectic phase due to reduction of the steric hindrance between the potentially smectogenic A11SOP or A11SOM moieties.  相似文献   

18.
A series of cis-5-norbornene-endo-2,3-dicarboxylic anhydride(NDCA,M1) derivatives(M2-M4) with different types of nonpolar substituted groups were synthesized and characterized by ~1H/~(13)C-NMR and mass spectrometry(MS). Ringopening metathesis polymerization(ROMP) of these monomers using the Grubbs third generation catalyst(G3) generated high molecular weight polymers with much improved solubility compared with the NDCA's homopolymer. It was found that the solubility of these polymers increased with increased substituent's steric hindrance. The living polymerization of NDCA derivative containing the bulkiest substituent(M4) catalyzed by G3 in tetrahydrofuran was confirmed by the kinetic studies with low polydispersity indices(PDI)( 1.30). By using sequential ROMP, well-defined diblock copolymers containing anhydride groups were synthesized.  相似文献   

19.
The 13C NMR spectra of stereoisomeric trans- and cis-1,3-dimethyl-2-oxa[3](1, 1′)ferrocenophanes, 1,2,3-trithia[3](1, 1′)ferrocenophane and 1,3-dithia[3](1, 1′)ferrocenophane have been recorded and analysed. The presence of steric hindrance in some of these compounds was confirmed and the conformations of the molecules were determined.  相似文献   

20.
The effect of steric hindrance on the reaction of silicon-containing aliphatic amines with phenylglycidyl ether (PGE) has been studied using 13C-NMR, a technique described by Sojka and Moniz [J. Appl. Polym. Sci., 20 , 1977 (1976)]. It was found that steric hindrance produced a large change in the reaction rate of the primary and secondary amine hydrogens, thus leading to a separation of the polymerization and cross-linking reactions.  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