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.
Three 1:1 adducts have been obtained by heating benzaldoxime dehydrodimer with styrene. The main product possesses the structure of bis-nitrone type. A radical addition mechanism is proposed.  相似文献   

2.
The structure of the title compound, C9H8N4, comprises non‐planar mol­ecules that associate via pyrimidine N—H?N dimer R(8) hydrogen‐bonding associations [N?N 3.1870 (17) Å] and form linear hydrogen‐bonded chains via a pyrimidine N—H?N(pyridyl) interaction [N?N 3.0295 (19) Å]. The dihedral angle between the two rings is 24.57 (5)°. The structure of the 1:1 adduct with 4‐amino­benzoic acid, C9H8N4·C7H7NO2, exhibits a hydrogen‐bond­ing network involving COOH?N(pyridyl) [O?N 2.6406 (17) Å], pyrimidine N—H?N [N?N 3.0737 (19) and 3.1755 (18) Å] and acid N—H?O interactions [N?O 3.0609 (17) and 2.981 (2) Å]. The dihedral angle between the two linked rings of the base is 38.49 (6)° and the carboxyl­ic acid group binds to the stronger base group in contrast to the (less basic) complementary hydrogen‐bonding site.  相似文献   

3.
Polyesters of the formula, [(OR)n? O? CO? C(CH3)2? N?N? C(CH3)2? CO]m, where (OR)n are poly(ethylene oxide), Poly(propylene oxide), or PTHF units, were used to prepare block copolymers with styrene. Ester and ether groups were cleaved with HI, NaOCH3, and diisobutylaluminum hydride. The resulting polystyrene is telechelic with two COOH and OH groups, respectively. The number of styrene blocks per polymer molecule is 3–4.  相似文献   

4.
We report the reactions of catecholborane (HBcat; 1 ) with unsaturated and saturated NHCs as well as CAACMe. Mono‐NHC adducts of the type HBcat?NHC (NHC=nPr2Im, iPr2Im, iPr2ImMe, and Dipp2Im) were obtained by stoichiometric reactions of HBcat with the unsaturated NHCs. The reaction of CAACMe with HBcat yielded the B?H activated product CAACMe(H)Bcat via insertion of the carbine‐carbon atom into the B?H bond. The saturated NHC Dipp2SIm reacted in a 2:2 ratio yielding an NHC ring‐expanded product at room temperature forming a six‐membered ?B?C=N?C=C?N? ring via C?N bond cleavage and further migration of the hydrides from two HBcat molecules to the former carbene‐carbon atom.  相似文献   

5.
Glow discharge polymerization between hexamethyldisilazane (HMDSZ) and trimethylsilyldimethylamine (TMSDMA) was compared by means of infrared spectroscopy and ESCA analysis. Infrared spectra pointed out differences in chemical structure between the polymers prepared from the two monomers, although the two polymers were mainly composed of resembling units such as Si? CH3, Si? CH2, Si? H, Si? O? Si, and Si? O? C groups: (i) The polymers prepared from TMSDMA contained N → O group, but the polymers from HMDSZ did not contain this group. (ii) Influences of the W/FM parameter (W is the input energy of rf power, F the flow rate of the monomer, and M the molecular weight of the monomer) appeared on decreasing the C? N group and increasing the C?O group in the TMSDMA system, but little influence appeared in the HMDSZ system. ESCA spectra (C1s, Si2p, and N1s core levels) supported the differences between the two polymers elucidated by infrared spectroscopy, and pointed out differences in susceptibility of the Si? N bond to plasma: The N? Si sequence of TMSDMA was completely ruptured in discharge to yield polymers, and the Si? NH? Si sequence of HMDSZ remained in considerable amount.  相似文献   

6.
The reaction of the oxygen-bridged frustrated Lewis pairs (FLPs) tBu2P−O−Si(C2F5)3 ( 1 ) and tBu2P−O−AlBis2 ( 2 ) with azobenzene, promoted by UV irradiation, led to a selective complexation of the cis-isomer. The addition product of 2 is stable, while the adduct of 1 isomerizes in solution in an ortho-benzidine-like [3,3]-rearrangement by cleavage of the N−N bond, saturation of the nitrogen atoms with hydrogen atoms and formation of a new bond between two phenyl ortho-carbon atoms. Similar rearrangements take place with different para-substituted azobenzenes (R=Me, OMe, Cl) and di(2-naphthyl)diazene, while ortho-methylated azo compounds do not form adducts with 1 . All adducts were characterized by multinuclear NMR spectroscopy and elemental analyses and the mechanism of the rearrangement was explored by quantum-chemical calculations.  相似文献   

7.
The title compound [(t-C4H9O)3SiS]2Co(NCCH3)] 1 was obtained by reaction of anhydrous cobalt(II) chloride with tri-tert-butoxysilanethiol and triethylamine in acetonitrile as a solvent. The compound crystallizes as deep-blue orthorhombic plates with a = 17.779(4), b = 45.363(9), c = 9.096(2) Å, space group Fdd2 and Z = 8. The structure was solved by Patterson synthesis and refined to the R value of 0.0343. The crystal consists of mononuclear complexes in which the cobalt atom is five-fold coordinated to two sulphurs, two oxygens and one nitrogen in a distorted trigonal bipyramidal arrangement. The relevant bond distances and angles are: Co? S, 2.2680(7); Co? N, 2.065(4); Co? O1, 2.283(2); S? Si, 2.0666(8) Å; S? Co? S′, 119.14(4); N? Co? S, 120.43(2); O1? Co? O1′, 178.81(10)°.  相似文献   

8.
Four triorganotin(IV) complexes constructed from tetrafluorophthalic acid (H2tfp) with a 1?:?1?:?1 molar ratio of H2tfp: Et3N: R3SnCl gave two of type {[R3Sn (tfp)].Et3NH}4 (R?=?Me 1, R?=?n-Bu 2), and two of type [R3Sn (tfp).Et3NH] n (R?=?PhCH2 3, Ph 4). All the complexes are characterized by elemental, IR, 1H, 13C and 119Sn NMR analyses. Complexes 1 and 4 were also confirmed by X-ray crystallography. Complex 1 is tetranuclear with a 28-membered C16O8Sn4 macrocyclic ring system with a cavity. The supramolecular structure of 1 has been found to consist of a three-dimensional network built up by intermolecular N–H?···?O, C–H?···?O hydrogen bonds and C–F?···?F weak interactions. Complex 4 is an infinite polymeric structure. The salient feature of the supramolecular structure of 4 is that of a two-dimensional plane, in which intermolecular N–H?···?O and C–H?···?π hydrogen bonds are important.  相似文献   

9.
Thioacetic acid and dithioacetic acid react with alkynederivatives of the type (CH3)2N? C?C? CO? R ( 1 ) in the same way as other carboxylic acids: The addition to dimethylaminopropinal ( 1a ) at low temperatures yields, after rearrangement of the very instable primary adducts, Z-3-acetoxy-N,N-dimethyl-thioacrylamide ( Z-16 ) and Z-3-thioacetoxy-N,N-dimethylthioacrylamide ( Z-17 ) respectively. The structure of the two compounds can be proved by spectroscopic evidence of 16 and 18 , the latter being formed by elimination of thioketene from 17 . According to the distribution of S-atoms in 16 and 17 , two reaction pathways including 4-membered rings can be ruled out. Thus the rearrangement of 3-acyloxy-N,N-dimethyl-acrylamides most probably proceeds by a mechanism including a dipolar six-membered intermediate. This mechanism cannot be valid for the rearrangement of the adducts 2 of hydrohalogen acids, alcohols and amines to the alkyne-derivatives 1 . The acid-catalysed reaction of 3-chloro-3-dimethylamino-propenal ( 2 , X?Cl), labelled at position 1 with 13C, yields 3-chloro-N,N-dimethyl-acrylamide ( 3 , X?Cl), containing the label exclusively at position 3 . This result supports a mechanism including an immonium-oxetene 21 (X?Cl) as intermediate. - The experiments are in accord with kinetic investigations.  相似文献   

10.
A computational study based on density functional theory was undertaken to identify possible reaction pathways for the formation and decomposition of peroxynitrite at models of the active sites of the nonheme superoxide scavenging enzymes superoxide reductase (SOR) and iron superoxide dismutase (FeSOD). Two peroxynitrite isomers and their possible protonated states were investigated, namely Fe? OONO?, Fe? N(O)OO?, Fe? OONOH, and Fe? N(O)OOH. Peroxynitrite formation at the active sites was assumed by either the interaction of a peroxynitrite cis/trans anion with the pentacoordinated iron active site or the interaction between a nitric oxide bound adduct and superoxide; both scenarios were found to be facile for all models investigated. The ferrous adducts of the Fe? OONO?isomer were found to undergo instant heterolytic cleavage of the O? ONO bond to yield nitrite, whereas for the ferric adducts, the homolytic cleavage of the O? ONO bond to yield nitrogen dioxide was found to be energetically facile. For the Fe? N(O)OO? isomer, the active site models of FeSOD and SOR were only able to accommodate the cis isomer of peroxynitrite. Ferric adducts of the cis Fe? OONO? isomer were found to be energetically more stable than their trans counterparts and were also more stable than the cis adducts of the Fe? N(O)OO? isomer; conversely, the protonated forms of all adducts of the Fe? OONOH isomer were found to be lower in energy than their equivalent Fe? N(O)OOH adducts. Multiple reaction pathways for the decomposition of the formed peroxynitrite adducts (whether the anions or the protonated forms) were proposed and explored. The energy requirements for the decomposition processes ranged from exothermic to highly demanding depending on the peroxynitrite isomer, the type of model (whether an SOR or FeSOD active site), and the oxidation state of iron. © 2014 Wiley Periodicals, Inc.  相似文献   

11.
The oxoboryl complex trans‐[(Cy3P)2BrPt(B?O)] ( 2 ) reacts with the Group 13 Lewis acids EBr3 (E=Al, Ga, In) to form the 1:1 Lewis acid–base adducts trans‐[(Cy3P)2BrPt(B?OEBr3)] ( 6 – 8 ). This reactivity can be extended by using two equivalents of the respective Lewis acid EBr3 (E=Al, Ga) to form the 2:1 Lewis acid–base adducts trans‐[(Cy3P)2(Br3Al‐Br)Pt(B?OAlBr3)] ( 18 ) and trans‐[(Cy3P)2(Br3Ga‐Br)Pt(B?OGaBr3)] ( 15 ). Another reactivity pattern was demonstrated by coordinating two oxoboryl complexes 2 to InBr3, forming the 1:2 Lewis acid–base adduct trans‐[{(Cy3P)2BrPt(B?O)}2InBr3] ( 20 ). It was also possible to functionalize the B?O triple bond itself. Trimethylsilylisothiocyanate reacts with 2 in a 1,2‐dipolar addition to form the boryl complex trans‐[(Cy3P)2BrPt{B(NCS)(OSiMe3)}] ( 27 ).  相似文献   

12.
The molecules of the title compound, C18H16N4O3, exhibit a very polarized molecular–electronic structure. The mol­ecules are linked into chains by a combination of an asymmetric three‐centre N—H?(N,O) hydrogen bond [H?N 2.19, H?O 2.54, N?N 3.041 (4) and N?O 2.977 (4) Å, and N—H?N 168, N—H?O 112 and N?H?O 67°] and an N—H?π(arene) interaction [H?Cg 2.67 Å, N?Cg 3.496 (4) Å and N—H?Cg 163°; Cg is a benzyl ring centroid].  相似文献   

13.
4-Vinylpyridine ( 1a ) combines with 3 moles of dienophilic N-alkyl-maleinimides ( 2 ) in the presence of polymerization inhibitors. The first step of the reaction probably consists of 1:1-addition with participation of an aromatic double bond, comparable to the analogous behavior of styrene and its derivatives under similar conditions. The unstable intermediates 3 , like other Schiff bases (imines), add 2 further moles of the N-alkyl-maleinimides forming the spiro compounds 4 . These are split in an acidic medium into the N-alkyl-5,6,7,8-tetrahydroisoquinoline-7,8-dicarboximides ( 5 ), and N,N′-dialkyl-2-butene-1,2,3,4-tetracarboxylic 1,2,:3,4-diimides ( 6 ). LiAlH4 reduction of these two types of compounds leads to N-alkyl-1 H-(3,4-d)-pyrrolo-2,3,3a,4,5,9b-hexahydroisoquinolines ( 7 ) and to N,N′dialkyl-3,3′-bipyrrolidyls ( 8A ) and their dehydro-products 8B , respectively. From the reaction of 2-vinylpyridine ( 1b ) with N-alkyl-maleinimides ( 2 ) the 1:2-addition products 9 can be isolated in the presence of polymerization inhibitors, which are derivatives of N-alkyl-5,6,7,8-tetrahydroquinoline-5,6-dicarboximides ( 9 ). This again corresponds to the reaction type of cycloadditions with styrene. Furthermore 1:3 adducts are formed which according to 1H- and 13C-NMR.-data most likely have the structure 10 , representing a new type of cycloaddition involving the pyridine nitrogen.  相似文献   

14.
In the title compound, [Li(C5H3N4O2)(H2O)2]n, the coordinate geometry about the Li+ ion is distorted tetrahedral and the Li+ ion is bonded to N and O atoms of adjacent ligand mol­ecules forming an infinite polymeric chain with Li—O and Li—N bond lengths of 1.901 (5) and 2.043 (6) Å, respectively. Tetrahedral coordination at the Li+ ion is completed by two cis water mol­ecules [Li—O 1.985 (6) and 1.946 (6) Å]. The crystal structure is stabilized both by the polymeric structure and by a hydrogen‐bond network involving N—H?O, O—H?O and O—H?N hydrogen bonds.  相似文献   

15.
CrIII Phthalocyaninates: Synthesis, Properties, and Crystal Structure of l-Bis(triphenylphosphine)iminium trans-Di(nitrito(O))phthalocyaninato(2–)chromate(III) [Cr(H2O)2Pc2?]Ix reacts with excess (PNP)NO2 in dimethylformamide to yield less soluble greenblack l-bis(triphenylphosphine)iminium trans-di(nitrito(O))phthalocyaninato(2–)chromate(III), l(PNP)trans[Cr(ONO)2Pc2?], which crystallizes in the triclinic space group P1 (No. 2) with Z = 2. The Cr atom is in the center of the Pc2? ligand and the two nitrite ions are monodentate O-coordinated in a mutually trans arrangement to the Cr atom. The Cr? O and Cr? Niso bond distances are 1.9898(14) und 1.981(2) Å, respectively. The geometric data of the coordinated nitrite ion are: d(N? O) = 1.307(2) Å; d(N? O) = 1.205(2) Å; ?(O? N? O) = 113.7(2)°; ?(Cr? O? N) = 116.85(12)°. The non-bonding O atoms are trans to the Cr atom. The Pc2? ligand is slightly saddled. Three weak spin-allowed trip-quartet(TQ) transitions (in 103 cm?1): TQ1 (8.20) < TQ2 (11.3) < TQ3 (20.33) and the characteristic π-π* transitions of the Pc2? ligand: B (14.68) < Q1 (27.1) < Q2 (29.0) < N (35.4) are observed in the UV-VIS-NIR spectrum. Prominent luminescence spectra are obtained by excitation within the TQ1 region, in which the spin-forbidden trip-sextet transition at 7376 cm?1 dominates at low temperatures (T < 50 K). The vibrational spectra are discussed. In coincidence of the excitation lines with TQ3, vs(Cr? O) at 378 cm?1 is selectively resonance Raman (RR) enhanced. vas(Cr? O) is observed in the FIR spectrum at 391 cm?1. The following internal vibrations (in cm?1) of the nitrito ligand are in the MIR spectrum: vas(N? O)/1447 > vas(N? O)/1018/1029 > δ(O? N? O)/828 and in the RR-spectrum: vs(N? O)/1410 > vs(N? O)/952, the last followed by three overtones.  相似文献   

16.
In ferrocene‐1,1′‐diyl­bis­(di­phenyl­methanol)–4,4′‐ethyl­enedi­pyridine (1/1), [Fe(C18H15O)2]·C12H12N2, there is an intra­molecular O—H?O hydrogen bond in the ferrocenediol component and a single O—H?N hydrogen bond linking the two components into a finite monomeric adduct. Ferrocene‐1,1′‐diyl­bis­(di­phenyl­methanol)–ethyl­enedi­amine (1/1), [Fe(C18H15O)2]·C2H8N2, crystallizes with Z′ = 2 in space group P, and there are two independent four‐component aggregates in the structure, both of which are centrosymmetric. In the first type of aggregate, the molecular components are linked by O—H?N and N—H?O hydrogen bonds, in which both di­amine N atoms participate; in the second type of aggregate, the di­amine component is disordered over two sets of sites, but only one N atom is involved in the hydrogen bonding.  相似文献   

17.
Three metal-organic coordination polymers based on 4-(1H-l,3-benzimidazol-2-yl)pyridine 1-oxide (BImPyO) with the molecular structures [Cu2(C12H8N3O)2]n (1), [Cu(C12H8N3O)]n (2), and [Zn(C12H8N3O)Cl]n (3) were synthesized under hydrothermal conditions. They showed diverse coordination modes and were further characterized by elemental analysis, infrared spectroscopy, and single crystal X-ray structure analysis, respectively. In 1 and 2, BImPyO generated a 1-D chain by adopting μ2-kN?:?kN′ coordination to bridge two Cu(II) ions with bis-N-chelation. In 3, by adopting μ3-kN?:?kN′:kO coordination, BImPyO bridged three crystallographically independent Zn(II) ions to form a 3-D framework; 3 exhibits strong fluorescent emission in the solid state at room temperature.  相似文献   

18.
We have carried out a computational study of the reactive properties of chlorooxirane, the metabolically produced epoxide of vinyl chloride that is believed to be a direct-acting carcinogenic form of this molecule. An ab initio SCF-MO procedure (GAUSSIAN 70) was used to compute the energy requirements for stretching the C? Cl and both C? O bonds (SN 1 reactivity) and to determine the course of the epoxide's possible SN 2 reactions with ammonia, taken as a model for nucleophilic sites on DNA. The epoxide was assumed to be protonated; both the oxygen- and chloro-protonated forms were considered. At each step along the various reaction pathways, the structure of the system was reoptimized. For the oxygen-protonated epoxide, the C1? O bond has a significantly lower energy barrier to stretching than does the C2? O. (The carbon bearing the chlorine is designated C1.) However, both are very much higher than that of the C? Cl bond in the chloro-protonated form, confirming our earlier finding of the relative weakness of this bond. In the SN 2 processes involving ammonia, intermediate complexes are formed with both carbons of the oxygen-protonated epoxide, the C2-complex being the more stable. However, the most stable ammonia complex occurs at C1 of the chloro-protonated epoxide. Our calculated results, both the energies and also the geometry changes, allow us to propose two possible mechanisms for the formation of the 7-N-(2-oxoethyl) derivative of guanine that has been observed to be the major in vivo DNA alkylation product of vinyl chloride and has been suggested as possibly being responsible for its carcinogenicity. One of these mechanisms is SN 1 and starts with the chloro-protonated epoxide; the other is SN 2 and involves the oxygen-protonated form.  相似文献   

19.
A cyano‐bridged bimetallic 4f–3d complex, tri­aqua‐1κ3O‐μ‐cyano‐1:2κ2N:C‐penta­cyano‐2κ5C‐tetrakis(2‐pyrrolidone‐1κO)­chromium(III)­dysprosium(III) dihydrate, [CrDy(C4H7NO)4(CN)6(H2O)3]·2H2O, has been prepared and characterized by X‐ray crystallographic analysis. The structure consists of a neutral cyano‐bridged Dy–Cr dimer. A hydrogen‐bonded three‐dimensional architecture is formed through N—H?O, O—H?N and O—H?O hydrogen bonds.  相似文献   

20.
Great efforts have been made to convert renewable biomass into transportation fuels. Herein, we report the novel properties of NbOx‐based catalysts in the hydrodeoxygenation of furan‐derived adducts to liquid alkanes. Excellent activity and stability were observed with almost no decrease in octane yield (>90 % throughout) in a 256 h time‐on‐stream test. Experimental and theoretical studies showed that NbOx species play the key role in C? O bond cleavage. As a multifunctional catalyst, Pd/NbOPO4 plays three roles in the conversion of aldol adducts into alkanes: 1) The noble metal (in this case Pd) is the active center for hydrogenation; 2) NbOx species help to cleave the C? O bond, especially of the tetrahydrofuran ring; and 3) a niobium‐based solid acid catalyzes the dehydration, thus enabling the quantitative conversion of furan‐derived adducts into alkanes under mild conditions.  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