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.
The introduction of substituents with various inductive constants in the 2 position of the diazabicyclo]2.2.2]octane ring gives rise to a change in the pKa values and affects the ease of introduction of a methyl group in the 1 and 4 positions, as well as the rate of demethylation of the bisquaternary salts from the 1 position. When 1,4-diazabicyclo[2.2.2]octane is treated with a sufficiently strong nucleophile, the ring is opened to give a piperazine derivative.See [1] for communication 7.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 2, pp. 250–255, February, 1982.  相似文献   

2.
Halogen bonding (XB) is a highly‐directional class of intermolecular interactions that has been used as a powerful tool to drive the design of crystals in the solid phase. To date, the majority of XB donors have been iodine‐containing compounds, with many fewer involving brominated analogues. We report the formation of adducts in the vapour phase from a series of dibromoperfluoroalkyl compounds, BrCF2(CF2)n CF2Br (n = 2, 4, 6), and 1,4‐diazabicyclo[2.2.2]octane (DABCO). Single‐crystal X‐ray diffraction studies of the colourless crystals identified 1,4‐diazabicyclo[2.2.2]octane–1,4‐dibromoperfluorobutane (1/1), C4Br2F8·C6H12N2, (I), 1,4‐diazabicyclo[2.2.2]octane–1,6‐dibromoperfluorohexane (1/1), C6Br2F12·C6H12N2, (II), and 1,4‐diazabicyclo[2.2.2]octane–1,8‐dibromoperfluorooctane (1/1), C8Br2F16·C6H12N2, (III), each of which displays a one‐dimensional halogen‐bonded network. All three adducts exhibit N…Br distances less than the sum of the van der Waals radii, with butane analogue (I) showing the shortest N…Br halogen‐bond distances yet reported between a bromoperfluorocarbon and a nitrogen base [2.809 (3) and 2.818 (3) Å], which are 0.58 and 0.59 Å shorter than the sum of the van der Waals radii.  相似文献   

3.
Introduction1,4 Diazabicyclo[2 .2 .2 ]octane (DABCO)wasre portedtocatalyzeorganicreactionsduetoitsstrongbasici ty .1,2 Severalchiraltrans 2 ,3 disubstitutedDABCOshavebeensynthesizedandappliedtotheasymmetricBaylis Hillmanreaction3andvicinalhydroxylation .4ThefirstsynthesisofthetitlecompoundwasreportedbySoai5from (2S ,5S) bis(phenylmethyl)piperazine (1) ,asshowninScheme 1.Butthisprocedureislengthy ,andtheoverallyieldisnotsosatisfactory .Besides ,thereport edmethodforthepreparationof 1is…  相似文献   

4.
The title compound, C18H19Cl2NO4·C6H12N2·H2O, is a cocrystal hydrate containing the active pharmaceutical ingredient felodipine and diazabicyclo[2.2.2]octane (DABCO). The DABCO and water molecules are linked through O—H...N hydrogen bonds into chains around 21 screw axes, while the felodipine molecules form N—H...O hydrogen bonds to the water molecules. The felodipine molecules adopt centrosymmetric back‐to‐back arrangements that are similar to those present in all of its four reported polymorphs. The dichlorophenyl rings also form π‐stacking interactions. The inclusion of water molecules in the cocrystal, rather than formation of N—H...N hydrogen bonds between felodipine and DABCO, may be associated with steric hindrance that would arise between DABCO and the methyl groups of felodipine if they were directly involved in hydrogen bonding.  相似文献   

5.
A new crystalline form of 1,4‐diazabicyclo[2.2.2]octane (DABCO) monohydrate, C6H12N2·H2O, crystallizing in the space group P31, has been identified during screening for cocrystals. There are three DABCO and three water molecules in the asymmetric unit, with two DABCO molecules exhibiting disorder over two positions related by rotation around the N...N axis. As in the monoclinic C2/c (Z′ = 2) polymorph, the molecular components are connected via O—H...N hydrogen bonds into a polymeric structure that consists of linear O—H...N(CH2CH2)3N...H—O segments, which are approximately mutually perpendicular. The two polymorphic forms of DABCO monohydrate can be considered as structural analogues of NaCl, with the nearly globular DABCO molecules showing distorted cubic closest packing and all octahedral interstices occupied by water molecules.  相似文献   

6.
The natural abundance 15N-NMR chemical shifts of selected aliphatic amines, 2-substituted pyridine type compounds, bialicyclic tertiary amines have been measured as a function of the nature of the solvent. In the case of cyclic aliphatic amines, like piperidine, morpholine, piperazine, thiomorpholine, the nitrogen is more shielded in concentrated solution compared to that in dilute solution whereas in the hydrogen bonding and protonating solvents there is a prominent deshielding. 2-Substituted pyridines studied can be further divided into four sub groups. The site of hydrogen bonding and protonation in 2-amino, 2-hydroxy and 2-mercapto pyridines have been conclusively proved from the 15N-NMR chemical shifts and the well-known tautomeric forms of the above compounds. Similarly in the case of 2-(2-thienyl)pyridine and 2-(3-thienyl)pyridine, the site of donation has been proved as the nitrogen of the pyridine ring in both the compounds. In a similar manner, the site of hydrogen bonding and protonation in two individual compounds 2-anilinopyridine and 2-(2-pyridyl)benzimidazole have also been established. Among the bialicyclic amines, 1,2-diazabicyclo[2.2.2]octane (DABCO) behaved differently from the other two compounds. In both 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) and 1,5-diazabicyclo[4.3.0]non-5-ene (DBN), it was possible to show that N1-nitrogen in both the compounds is the site of donation. The effect of the second donor site on the 15N-NMR chemical shift, the site of donation in the selected compounds and some typical compounds reported in literature have been presented and discussed.  相似文献   

7.
Unsymmetrical piperazines are key constituents of many pharmaceuticals. Given that the selective introduction of an aryl and alkyl motif onto the piperazine is not always straightforward, direct arylation and alkenylation of 1,4‐diaza‐bicyclo[2.2.2]octane would obviate the inefficiencies associated with the preparation of these target molecules. We have utilized alkyl halides, aryl or alkenyl triflates, and 1,4‐diaza‐bicyclo[2.2.2]octane for the synthesis of N‐alkyl‐N ′‐aryl or alkenylpiperazines. The optimum conditions are developed using CuCl, t‐BuOL i in NMP . Alkenyl triflates requires N ,N ′‐dimethylethylenediamine and higher temperature to afford the desired cross‐coupled product. Substrates bearing electron‐deficient and electron‐rich groups were successfully coupled under the optimum reaction conditions.  相似文献   

8.
Studying the axial ligation behavior of metalloporphyrins with nitrogenous bases helps to better understand not only the biological function of heme‐based protein systems, but also the catalytic properties of porphyrin‐based reaction sites in other biomimetic synthetic support environments. Unlike iron porphyrin complexes, little is known about the axial ligation behavior of Mn porphyrins, particularly in the solid state with Mn in the +3 oxidation state. Here, we present the syntheses and crystal and molecular structures of three new high‐spin manganese(III) porphyrin complexes with the different amine‐based axial ligands imidazole (im), piperidine (pip), and 1,4‐diazabicyclo[2.2.2]octane (DABCO), namely bis(imidazole)(5,10,15,20‐tetraphenylporphyrinato)manganese(III) chloride chloroform disolvate, [Mn(C44H28N4)(C3H4N2)2]Cl·2CHCl3 or [Mn(TPP)(im)2]Cl·2CHCl3 (TPP = 5,10,15,20‐tetraphenylporphyrin), (I), bis(piperidine)(5,10,15,20‐tetraphenylporphyrinato)manganese(III) chloride, [Mn(C44H28N4)(C5H11N)2]Cl or [Mn(TPP)(pip)2]Cl, (II), and chlorido(1,4‐diazabicyclo[2.2.2]octane)(5,10,15,20‐tetraphenylporphyrin)manganese(III)–1,4‐diazabicyclo[2.2.2]octane–toluene–water (4/4/4/1), [Mn(C44H28N4)Cl(C6H12N2)]·C6H12N2·C7H8·0.25H2O or [Mn(TPP)Cl(DABCO)]·(DABCO)·(toluene)·0.25H2O, (IV). A fourth complex, chlorido(pyridine)(5,10,15,20‐tetraphenylporphryinato)manganese(III) pyridine disolvate, [Mn(C44H28N4)Cl(C5H5N)]·2C5H5N or [Mn(TPP)Cl(py)]·2(py), (III), acquired using different crystallization methods from published data, is also reported and compared to the previous structures.  相似文献   

9.
Halogen bonding is an intermolecular interaction capable of being used to direct extended structures. Typical halogen‐bonding systems involve a noncovalent interaction between a Lewis base, such as an amine, as an acceptor and a halogen atom of a halofluorocarbon as a donor. Vapour‐phase diffusion of 1,4‐diazabicyclo[2.2.2]octane (DABCO) with 1,2‐dibromotetrafluoroethane results in crystals of the 1:1 adduct, C2Br2F4·C6H12N2, which crystallizes as an infinite one‐dimensional polymeric structure linked by intermolecular N...Br halogen bonds [2.829 (3) Å], which are 0.57 Å shorter than the sum of the van der Waals radii.  相似文献   

10.
The photoreduction of 4,4'-bipyridine (44BPY) by diazabicyclo[2.2.2]octane and triethylamine (TEA) is investigated by using picosecond transient absorption and time-resolved resonance Raman spectroscopy in various acetonitrile-water mixtures. The results are interpreted on the basis of a preferential solvation effect resulting from the presence of a specific interaction between 44BPY and water by hydrogen bonding. Below 10% water, the free 44BPY species is dominant and leads upon photoreduction to a contact ion pair that undergoes efficient intrapair proton transfer if TEA is the amine donor. Above 10% water, most of the 44BPY population is H-bonded and leads upon photoreduction to a hydrated ion pair in which the intrapair proton transfer is inhibited. Instead, the 44BPY(-*) species is protonated by water through the hydrogen bond with a rate constant that increases by more than 3 orders of magnitude on going from 10% to 100% water. The dependence of this rate constant on the solvent mixture composition suggests that the reaction of intracomplex proton transfer is controlled by the hydration of the residual OH(-) species by three molecules of water, leading to a trihydrated HO(-)(H(2)O)(3) species.  相似文献   

11.
Boron–nitrogen dative bonds provide a suitable motif for reversible, yet strong and directed interactions, leading to the highly efficient self‐assembly of small organic building blocks into supramolecular cage structures. A bipyramidal [2+3] assembly, as the first example of a supramolecular cage mediated by B?N dative bonds that exists as a discrete species in solution, is quantitatively obtained from a tribenzotriquinacene‐based trisboronate ester and 1,4‐diazabicyclo[2.2.2]octane. Thermodynamic equilibria of cage formation are investigated by isothermal titration calorimetry and fully reversible cage opening can be observed at elevated temperatures.  相似文献   

12.
The coadsorption of CO and triethylenediamine (TEDA) (also called 1,4-diazabicyclo[2.2.2]octane, DABCO) on a high-area gamma-Al2O3 surface has been investigated with use of transmission FTIR spectroscopy. It has been found that TEDA binds more strongly to both Lewis acid sites and to Br?nsted Al-OH sites than does CO. Competition experiments indicate that TEDA displaces CO to less strong binding sites. Evidence for weak CO...TEDA interactions is found in which small nu(CO) redshifts are produced. Comparison between different amines such as triethylenemonoamine (TEMA) (also called 1-azabicyclo[2.2.2]octane, ABCO), trimethylamine (TMA), and ammonia indicates that the nu(CO) redshift increases with increasing amine polarizability, indicating that the redshift is mainly due to dipole image damping effects on the CO oscillator frequency. The direct bonding between the exposed N lone pair electrons of the TEDA molecule and CO does not occur. First principles theoretical studies have characterized the bonding of CO with gamma-Al2O3 Lewis acid sites of various types as well as TEDA bonding to both Lewis acid sites and to Al-OH groups. The theoretical studies also indicate that strong bonding of adsorbed CO with TEDA molecules does not occur, and that the observed decrease in the binding energy of CO when coadsorbed with TEDA on gamma-Al2O3 is expected.  相似文献   

13.
[reaction: see text] In a study directed toward the use of the chloroacetyl protecting group in carbohydrate synthesis, the sterically hindered tertiary amine diazabicyclo[2.2.2]octane (DABCO) was found to give complete and selective cleavage of the chloroacetyl group in the presence of other ester functions such as benzoyl and acetyl groups at primary and/or secondary positions.  相似文献   

14.
Three‐component reaction of 4‐hydroxy‐2H‐benzo[h]chromen‐2‐one, aromatic aldehydes, and malononitrile in the presence of 1,4‐diazabicyclo[2.2.2]octane (DABCO) in ethanol at room temperature affords good yields of novel dihyrobenzo[h]pyrano[3,2‐c]chromene derivatives. The synthesized compounds examined by MTT assays for cytotoxic activity in two human cancer cell lines (MOLT‐4, HL‐60). Most of the evaluated compounds showed low inhibitory activity against tumor cell line at micromolar concentrations.  相似文献   

15.
Simple pentafluorobenzyl‐substituted ammonium and pyridinium salts with different anions can be easily obtained by treatment of the parent amine or pyridine with the respective pentafluorobenzyl halide. Hexafluorophosphate is introduced as the anion by salt metathesis. In the case of the ammonium salt 4 , water co‐crystallisation seems to suppress effective anion–π interactions of bromide with the electron‐deficient aromatic system, whereas with salts 5 and 6 such interactions are observed despite the presence of water. However, due to asymmetric hydrogen‐bonding interactions with ammonium side chains, the anion of 5 is located close to the rim of the pentafluorophenyl group (η1 interaction). In 6 the CH–anion hydrogen bonding is more symmetric and fixes the anion on top of the ring (η6). A similar structure‐controlling effect is observed in case of the 1,4‐diazabicyclo[2.2.2]octane derivatives 7 . Here the position of the anion (Cl, Br, I) is shifted according to the length of the weak CH–halide interaction. The hexafluorophosphate 7 d reveals that this “non‐coordinating” anion can be located on top of an aromatic π system. In the methyl‐substituted pyridinium salts 9 and 10 different locations of the bromide anions with respect to the π system are observed. This is due to different conformations of the mono‐ versus disubstituted pyridine, which leads to different directions of the weak, but structurally important, HMe? Br bonds.  相似文献   

16.
任颜卫  陆家贤  江鸥  程晓飞  陈俊 《催化学报》2015,(11):1949-1956
稀土金属有机骨架(Ln-MOFs)是利用有机配体和稀土离子之间配位自组装形成的具有超分子多孔网络结构的类沸石材料,其优点是稳定性好,一般不溶于常规的有机和无机溶剂,并且孔径、孔形及孔表面性质可通过其构建分子的选择或修饰进行灵活设计和制备.稀土离子性能独特,有机配体种类繁多,将稀土离子与有机配体可控组装可获得许多结构多样、性能优异的Ln-MOFs材料.这些功能材料已在气体吸附与分离、发光器件、化学传感以及磁性材料等多方面显示出潜在应用价值.特别是Ln-MOFs材料作为非均相催化剂具有热稳定性高、比表面积大以及稀土离子配位环境多样等优点,近年来受到国内外研究者关注和重视.后合成修饰法(PSM)是利用MOFs骨架中不饱和配位的金属离子或潜在的有机反应基团,通过配位键或共价键方式引入有机或无机分子,制备具有新功能的骨架材料.本文采用PSM策略,将三种不同的有机二胺后合成修饰到具有配位不饱和位点的稀土金属有机骨架[Er(btc)]的孔道中,得到三种固体碱催化剂:Er(btc)(ED)0.75(H2O)0.25(2), Er(btc)(PP)0.55(H2O)0.45(3)和Er(btc)(DABCO)0.15(H2O)0.85(4).其中, btc为1,3,5-均苯三甲酸, ED为乙二胺, PP为哌嗪, DABCO位为三乙烯二胺.单晶结构分析表明,在[Er(btc)(H2O)]·DMF0.7(1)中,铒离子与六个btc配体的六个羧酸氧原子和一个水分子配位,形成变形的五角双锥几何构型.每个btc配体连接六个铒离子构成具有一维开放孔道(0.7 nm′0.7 nm)的三维立体结构.重要的是,孔道中的配位水分子和游离DMF分子可通过真空加热除去而不影响其骨架结构(热稳定性达500 oC),这将有利于对其进行后合成修饰.热重分析(TGA)表明,催化剂2在25–300 oC失去孔道中配位的乙二胺和水分子;催化剂3在250 oC之前失去孔道中的哌嗪和水分子;催化剂4则在100 oC之前失去孔道中配位的三乙烯二胺和水分子.粉末X射线衍射(PXRD)结果显示,后合成修饰过程并没有改变催化剂骨架的稳定性,其稳定性在空气中超过30 d.氮气吸附实验表明, Ln-MOF 1的比表面积和孔体积分别为2000 m2/g和0.75 cm3/g,平均孔尺寸为0.65 nm,与晶体结构分析结果基本一致.相比之下,后合成修饰的催化剂2的比表面积明显降低,为650 m2/g,而催化剂3和4由于后修饰较大体积的二胺分子(哌嗪和三乙烯二胺),表现出可忽略的氮气吸附能力.上述结果表明,催化剂2具有较高的有机胺负载量、较高的热稳定性和多孔性.采用苯甲醛和丙二腈的Knoevenagel缩合反应研究了三种固体碱的非均相催化能力.结果表明,在相同反应条件下,催化剂2具有很好的首次催化能力(99%),优于催化剂3(93%)和4(63%).并且,催化剂2循环使用三次后,催化能力几乎没有改变,而催化剂3和4的催化能力则逐渐降低,催化剂4在第三次使用时已无催化能力.滤出实验显示,催化剂2在反应过程中无活性物种离去进入液相体系中,即无乙二胺分子从催化剂骨架孔道中离去,证明其为非均相催化本质.而催化剂3和4则在反应过程中有二胺分子离去,进入反应液相中,从而导致其循环使用催化能力降低.催化剂2的底物择形催化反应结果显示,先是丙二腈分子进入催化剂孔道中,形成碳负离子,然后亲核进攻醛分子生成产物.因此,体积较大的腈衍生物因不能进入孔道而不能发生反应,而体积较大的醛分子则不受影响,能顺利地发生反应.  相似文献   

17.
Crystallisation studies of ethyl resorcinarene with diquats 2b and 3a (1,4-dimethyl-1,4-diazoniabicyclo[2.2.2]octane dibromide and 1,4-diazoniabicyclo[2.2.2]octane dichloride, respectively) resulted in hydrogen bonded molecular capsules in which the cations are encapsulated in between the cavities of two resorcinarene molecules and anions are located in the middle of the lower rim ethyl chains.  相似文献   

18.
The reactions of 2‐phenoxy‐3,5‐dinitropyridine ( 1 ) with a series of substituted anilines ( 4a–d ) in dimethyl sulfoxide (DMSO) in the presence of 1,4‐diazabicyclo[2.2.2]octane (DABCO) yield the 2‐anilino derivatives without the accumulation of intermediates. The kinetics is compatible with a two‐step reaction involving initial nucleophilic attack followed by either base‐catalyzed or uncatalyzed conversion to the product. The base‐catalyzed pathway is likely to involve rate‐limiting proton transfer from the zwitterionic intermediate to base. The results are compared with those for reactions of 1,3,5‐trinitrobenzene (2) and phenyl 2,4,6‐trinitrophenyl ether ( 3 , R = Ph ) with anilines. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 198–203, 2009  相似文献   

19.
The functionalisation of well-known rigid metal-organic frameworks (namely, [Zn(4)O(bdc)(3)](n), MOF-5, IRMOF-1 and [Zn(2)(bdc)(2)(dabco)](n); bdc = 1,4-benzene dicarboxylate, dabco = diazabicyclo[2.2.2]octane) with additional alkyl ether groups of the type -O-(CH(2))(n)-O-CH(3) (n = 2-4) initiates unexpected structural flexibility, as well as high sorption selectivity towards CO(2) over N(2) and CH(4) in the porous materials. These novel materials respond to the presence/absence of guest molecules with structural transformations. We found that the chain length of the alkyl ether groups and the substitution pattern of the bdc-type linker have a major impact on structural flexibility and sorption selectivity. Remarkably, our results show that a high crystalline order of the activated material is not a prerequisite to achieve significant porosity and high sorption selectivity.  相似文献   

20.
Abstract— Various nitrogen containing compounds have previously been shown to quench singlet oxygen (10z). When measuring the dimol 1O2 light emission arising from the H 2O2 /OCI- reaction, we found that certain cyclic diamines increase the emission of light, while other amines were inhibitory. This increase of light emission was seen with both 1, 4diazabicyclo[2.2.2]octane and N, N'-dimethylpi-perazine but not with acyclic analogues. Sodium azide inhibited both the normal and enhanced light emission. The enhanced light emission shows spectral properties characteristic of lO2 dimol emission.  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