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 140 毫秒
1.
Dibenzo‐24‐crown‐8 is studied herein as a flexible ligand able to adopt different conformations, as well as for the complexation of mercury. The recrystallization of dibenzo‐24‐crown‐8 (DB24C8) from dry THF gives a new polymorphic structure of this ligand. This new structure is described and compared to the literature compound. Additionally, coordination of this ligand to mercury iodide HgI2 is studied.  相似文献   

2.
This study covers phosphorylation of heterocyclic analogues of N‐methyldiphenylamine with phosphorus tribromide in pyridine solution. The reaction is found to proceed regioselectively in accordance with the orienting effect of the amino group. Mono and bis‐phosphorylated derivatives of the heterocycles have been isolated and characterized. It is pointed out that the heterocyclic systems under study exhibit reduced reactivity in electrophilic phosphorylation as compared to N‐methyldiphenylamine. The results of calculations by the PM3 method are reported for the starting molecules and their σ‐complexes. © 2001 John Wiley & Sons, Inc. Heteroatom Chem 12:652–657, 2001  相似文献   

3.
A high‐yield synthesis of 1‐triphenylmethyl‐3‐tert‐butylaziridinone ( 4 ), its physical and spectral properties, the limits of its thermal stability, and reactions with methanol, benzylamine and sodium methoxide in methanol are described.  相似文献   

4.
The reaction of 3‐amino‐5‐nitro‐1,2,4‐triazole with nitrous acid produces the corresponding diazonium salt. When the diazonium salt is treated with nitroacetonitrile, a subsequent condensation and cyclization reaction occurres to produced 4‐amino‐3,7‐dinitrotriazolo‐[5,1‐c][1,2,4] triazine (DPX‐26). X‐ray crystallographic analysis shows that the DPX‐26 has a density of 1.86 g cm?3, while it is calculated to have a heat of formation of 398.3 kJ mol?1. DPX‐26 is predicted to approach the explosive performance of RDX but displays significantly better safety properties. Oxidation of DPX‐26 using hypofluorous acid produces 4‐amino‐3,7‐dinitrotriazolo‐[5,1‐c][1,2,4] triazine 4‐oxide (DPX‐27), which is also predicted to be a high‐performance material with enhanced safety properties.  相似文献   

5.
In 3,4‐di‐2‐pyridyl‐1,2,5‐oxadiazole (dpo), C12H8N4O, each mol­ecule resides on a twofold axis and inter­acts with eight neighbours via four C—H⋯N and four C—H⋯O inter­actions to generate a three‐dimensional hydrogen‐bonded architecture. In the perchlorate analogue, 2‐[3‐(2‐pyrid­yl)‐1,2,5‐oxadiazol‐4‐yl]pyridinium perchlorate, C12H9N4O+·ClO4 or [Hdpo]ClO4, the [Hdpo]+ cation is bisected by a crystallographic mirror plane, and the additional H atom in the cation is shared by the two pyridyl N atoms to form a symmetrical intra­molecular N⋯H⋯N hydrogen bond. The cations and perchlorate anions are linked through C—H⋯O hydrogen bonds and π–π stacking inter­actions to form one‐dimensional tubes along the b‐axis direction.  相似文献   

6.
Spiro[fenchane‐2,2′‐(1,3,4)‐thiadiazoline] ( 6 ), prepared from thiofenchone and diazomethane, extrudes N2 (t1/2 22 min, 46°C, toluene) and furnishes the S‐methylide 7 which, in turn, closes the thiirane ring or else is intercepted by 1,3‐cycloadditions to dipolarophiles (tetracyanoethylene, maleic anhydride, N‐methyl‐1,2,4‐triazoline‐3,5‐dione, aromatic thioketones). When thiocarbonyl S‐methylide 7 is set free in methanol, fenchone S,O‐dimethylacetal is formed as an HX adduct. Catalysis by acetic acid converts 7 to 1‐(methylthio)‐α‐fenchene ( 25 ) by way of a Wagner‐Meerwein rearrangement. The addition of diazomethane to thiocampher leads, via thiadiazoline 12 , to thiocamphor S‐methylide; the latter undergoes a 1,4‐H shift, thus affording 2‐methylthio‐2‐bornene ( 11 ). © 2001 John Wiley & Sons, Inc. Heteroatom Chem 12:136–145, 2001  相似文献   

7.
The structures of a 14‐crown‐4 ether containing both benzo and cyclo­hexano substituents, 2,6,13,17‐tetraoxatricyclo‐[16.4.0.07,12]docosa‐1(18),19,21‐triene, C18H26O4, and its lith­ium complex, [2,6,13,17‐tetraoxatricyclo[16.4.0.07,12]docosa‐1(18),19,21‐triene‐κ4O](thio­cyanato‐N)­lith­ium(I), [Li(NCS)‐(C18H26O4)], are presented. The conformation of the free crown, (I), is not preorganized for cation binding, as its donor dipoles are oriented towards opposite sides of the crown ring. The Li+‐crown complex, (II), exhibits two formula units in the asymmetric unit. The binding conformation observed in (II) does not completely reorient the dipoles to one point, resulting in a long Li—O bond length [2.068 (5) and 2.073 (5) Å].  相似文献   

8.
An investigation of the free‐radical bulk photopolymerization of 1‐vinyl‐2‐pyrrolidinone (NVP) with an NVP‐based crosslinker, 1,6‐(bis‐3‐vinyl‐2‐pyrrolidinonyl)hexane (BNVP), and an NVP‐based comonomer, 3‐hexyl‐1‐vinyl‐2‐pyrrolidinone (VHP), was carried out. The enthalpies of polymerization were determined for NVP and VHP to be 30.8 and 35.7 kJ/mol, respectively. The rates of polymerization were determined for NVP/VHP and NVP/BNVP systems at various temperatures. These photopolymerization studies revealed that the overall rates of polymerization of these 3‐alkylated‐2‐pyrrolidinone derivatives increased with substitution onto the pyrrolidinone ring. A series of pyrrolidinone‐based additives in bulk NVP were used in model photopolymerizations of NVP for the evaluation of plasticizer effects. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 694–706, 2002; DOI 10.1002/pola.10142  相似文献   

9.
Unsolvated 5,10,15,20‐tetra‐4‐pyridylporphyrin, C40H26N8, (I), its sesquihydrate, C40H26N8·1.514H2O, (II), and its 2‐chlorophenol disolvate, C40H26N8·2C6H5ClO, (III), reveal different conformational features of the porphyrin core. In (I), the latter is severely deformed from planarity, apparently in order to optimize the intermolecular interactions and efficient crystal packing of the molecular entities. The molecular framework has a C1 symmetry. In (II), the porphyrin molecules are located on symmetry axes, preserving the marked deformation from planarity of the porphyrin core. The molecular units are interlinked into a single‐framework supramolecular architecture by hydrogen bonding to one another via molecules of water, which lie on twofold rotation axes. In (III), the porphyrin molecules are located across centres of inversion and are characterized by a planar conformation of the 24‐membered macrocyclic porphyrin ring. Two trans‐related pyridyl substituents are hydrogen bonded to the 2‐chlorophenol solvent molecules. The interporphyrin organization in (III) is similar to that observed for many other tetraarylporphyrin compounds. However, the organization observed in (I) and (II) is different and of a type rarely observed before. This study reports for the first time the crystal structure of the unsolvated tetrapyridylporphyrin.  相似文献   

10.
A simple and efficient method for the synthesis of highly substituted benzo‐ and hetero‐fused analog of 2, 3‐dihydro‐6H‐oxa‐3a‐aza‐phenalene was developed using 2H‐1, 4‐benzoxazine and α‐oxoketene dithio‐acetals. J. Heterocyclic Chem., (2011).  相似文献   

11.
The first example of an interpenetrated methyl‐modified MOF‐5 with the formula Zn4O(DMBDC)3(DMF)2, where DMBDC2? is 2,5‐dimethylbenzene‐1,4‐dicarboxylate and DMF is N,N‐dimethylformamide (henceforth denoted as Me2MOF‐5‐int ), namely, poly[tris(μ4‐2,5‐dimethylbenzene‐1,4‐dicarboxylato)bis(N,N‐dimethylformamide)‐μ4‐oxido‐tetrazinc(II)], [Zn4(C10H8O4)3O(C3H7NO)2]n, has been obtained from a solvothermal synthesis of 2,5‐dimethylbenzene‐1,4‐dicarboxylic acid and Zn(NO3)2·6H2O in DMF. A systematic study revealed that the choice of solvent is of critical importance for the synthesis of phase‐pure Me2MOF‐5‐int , which was thoroughly characterized by single‐crystal and powder X‐ray diffraction (PXRD), as well as by gas‐adsorption analyses. The Brunauer–Emmett–Teller surface area of Me2MOF‐5‐int (660 m2 g?1), determined by N2 adsorption, is much lower than that of nonpenetrated Me2MOF‐5 (2420 m2 g?1). However, Me2MOF‐5‐int displays an H2 uptake capacity of 1.26 wt% at 77 K and 1.0 bar, which is comparable to that of non‐interpenetrated Me2MOF‐5 (1.51 wt%).  相似文献   

12.
7‐Alkanoyloxy‐3,7‐dimefhyl‐7,8‐dihydro‐6H‐isochromene‐6,8‐diones 12‐15 were synthesized in 69‐16% yields from the reaction of 2,4‐dihydroxy‐3‐methyl‐6‐(2‐oxopropyl)benzaldehyde 11 with p‐toluenesulfonic acid in various carboxylic acids such as acetic acid, propionic acid, butyric acid and heptanoic acid followed by oxidation with lead tetraacetate. On the other hand, (±)‐daldinin A 5 (oleate) was not obtained using oleic acid as a medium. In the cases of heptanoic acid and oleic acid, esters 16 and 17 were produced in 23 and 9% yields, respectively. 6,8‐Dihydroxy‐3,7‐dimethyl‐2‐benzopyrylium p‐toluenesulfonate 31 is considered as the intermediate for the production of 12‐15. Overall yields of isochromenes 12‐15 were 26‐6% starting from 2‐methylresorcinol for seven steps.  相似文献   

13.
Lap shear and friction force measurements were carried out on a series of monodisperse polystyrene (PS) films below the corresponding glass‐transition temperatures. It showed that adhesion between the PS/PS interface was possible at the temperature below the bulk Tg, and the lower the molecular weight of PS, the lower the temperature at which the interfacial strength was detectable. The examination of a series of molecular weights indicated both the surface molecular motion and the magnitude of the interfacial strength were dependent on molecular weight and its distribution. And a steep increase of the friction force with increasing the test temperature was observed around 0 ∼ 30 °C. The contact angle of water versus molecular weight measurements also showed a transition at room temperature. The behavior observed in this study was supposed to be due to the increased molecular mobility, and was in good agreement with the measured surface transition temperatures by DSC. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 654–658, 2000  相似文献   

14.
The resolution of η5(2‐formyl‐3,4‐dimethylphospholyl)(triphenylphosphine)‐manganesedicarbonyl 1 has been carried out by chromatography of the acetals derived from (S,S)‐1,2‐diphenylethane‐1,2‐diol. The enantiopure 2‐diphenylphosphinomethyl 4 and diphenylmethylimino 5 derivatives have been prepared from 1 . © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:458–460, 2005; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20130  相似文献   

15.
On crystallization from CHCl3, CCl4, CH2ClCH2Cl and CHCl2CHCl2, 6‐chloro‐5‐hydroxy‐2‐pyridone, C5H4ClNO2, (I), undergoes a tautomeric rearrangement to 6‐chloro‐2,5‐dihydroxypyridine, (II). The resulting crystals, viz. 6‐chloro‐2,5‐dihydroxypyridine chloroform 0.125‐solvate, C5H4ClNO2·0.125CHCl3, (IIa), 6‐chloro‐2,5‐dihydroxypyridine carbon tetrachloride 0.125‐solvate, C5H4ClNO2.·0.125CCl4, (IIb), 6‐chloro‐2,5‐dihydroxypyridine 1,2‐dichloroethane solvate, C5H4ClNO2·C2H4Cl2, (IIc), and 6‐chloro‐2,5‐dihydroxypyridine 1,1,2,2‐tetrachloroethane solvate, C5H4ClNO2·C2H2Cl4, (IId), have I41/a symmetry, and incorporate extensively disordered solvent in channels that run the length of the c axis. Upon gentle heating to 378 K in vacuo, these crystals sublime to form solvent‐free crystals with P21/n symmetry that are exclusively the pyridone tautomer, (I). In these sublimed pyridone crystals, inversion‐related molecules form R22(8) dimers via pairs of N—H...O hydrogen bonds. The dimers are linked by O—H...O hydrogen bonds into R46(28) motifs, which join to form pleated sheets that stack along the a axis. In the channel‐containing pyridine solvate crystals, viz. (IIa)–(IId), two independent host molecules form an R22(8) dimer via a pair of O—H...N hydrogen bonds. One molecule is further linked by O—H...O hydrogen bonds to two 41 screw‐related equivalents to form a helical motif parallel to the c axis. The other independent molecule is O—H...O hydrogen bonded to two related equivalents to form tetrameric R44(28) rings. The dimers are π–π stacked with inversion‐related dimers, which in turn stack the R44(28) rings along c to form continuous solvent‐accessible channels. CHCl3, CCl4, CH2ClCH2Cl and CHCl2CHCl2 solvent molecules are able to occupy these channels but are disordered by virtue of the site symmetry within the channels.  相似文献   

16.
The gold electrode was modified by noisomes, prepared by sonicating the mixed solution of PEG 6000/Tween 80/Span 80/H2O. Electrochemical cyclic voltammetry (CV) and impedance spectroscopy (EIS) were performed for characterizing the modification. The construction of multilayer films of octadecanethiol (ODT)‐niosomes had almost no pinholes, which elicited that niosomes‐coated electrode through ODT assembly was more effective immobilization compared to direct modification method. A large semicircle formation in the entire range of frequency indicated the complete electron‐transfer control for the redox reaction, implying a perfect blocking behavior. The Nyquist plot was consisting of two depressed semicircles after storage for 36 h in air, which implied a progressive and uniform construction manner, as expected for porous coatings. The modified electrode was used to investigate the interaction between 1‐anilino‐8‐naphthalene‐sulfonate (ANS) and noisomes. It was found that the low‐frequency semicircular arcs increase in diameter, indicating that the increase in the resistance Rp attributed mainly to the niosomes/solution interface. It was due to the binding of ANS to the niosomes‐modified electrode surface, exhibited against diffusing species through the pores. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
The compound 6‐azaindole undergoes self‐assembly by formation of N(1)?H???N(6) hydrogen bonds (H bonds), forming a cyclic, triply H‐bonded trimer. The formation phenomenon is visualized by scanning tunneling microscopy. Remarkably, the H‐bonded trimer undergoes excited‐state triple proton transfer (ESTPT), resulting in a proton‐transfer tautomer emission maximized at 435 nm (325 nm of the normal emission) in cyclohexane. Computational approaches affirm the thermodynamically favorable H‐bonded trimer formation and the associated ESTPT reaction. Thus, nearly half a century after Michael Kasha discovered the double H‐bonded dimer of 7‐azaindole and its associated excited‐state double‐proton‐transfer reaction, the triply H‐bonded trimer formation of 6‐azaindole and its ESTPT reaction are demonstrated.  相似文献   

18.
The annulation reaction of vinamidinium salt containing nitrofurazanyl moiety at the β‐position gives access to the corresponding pyrazole. At nitration, two nitro groups were installed to the pyrazole ring. The synthesized 3‐(3,5‐dinitropyrazol‐4‐yl)‐4‐nitrofurazan 13 is strong NH acid and a new family energetic salts was prepared by direct neutralization with high nitrogen bases. Compound 13 crystallizes in the monoclinic space group P21/c, and charaterized by high density of 1.979 g/cm3 (at 100 K). J. Heterocyclic Chem., (2012).  相似文献   

19.
In this study, all‐solid‐state type potentiometric PVC membrane selective microsensor was developed for Metformin (MET) which is an antidiabetic drug active substance. Metformin‐tetraphenylborate (MET‐TPB) ion‐pair was used as an ionophore in the structure of the sensor membrane. It was determined that the sensor membrane at the ratio of 69 % o‐nitrophenyl octyl ether, 27 % polyvinyl chloride and 4 % MET‐TPB performed the best potentiometric performance. In a wide concentration range (1×10?5–1×10?1 mol/L), the slope, detection limit, response time, pH range, and life‐time of the sensor were determined as 55.9±1.6 mV (R2=0.996), 3.35×10?6 mol/L, 8–10 s, pH: 3–8, and ~10 weeks, respectively. The voltammetric performances of the sensor were also investigated. The prepared microsensor was successfully utilized for the determination of Metformin in a pharmaceutical drug sample by potentiometry and voltammetry. It was observed that the obtained results were in agreement with the results obtained by the UV spectroscopy method at 95 % confidence level.  相似文献   

20.
In novel superatom chemistry, it is very attractive that all‐metal clusters can mimic the behaviors of nonmetal atoms and simple nonmetal molecules. Wizardly all‐metal halogen‐like superatom Al13 with 2P5 sub shell (corresponding to the 3p5 of chlorine) is the most typical example. In contrast, how to mimic the behaviors of magnetic transition‐metal atom using all‐nonmetal cluster is an intriguing challenge for superatom chemistry. In response to this based on human intuition, using quantum chemistry methods and extending jellium model from metal cluster to all‐nonmetal cluster, we have found out that all‐nonmetal octahedral B6 cluster with characteristic jellium electron configuration 1S21P62S21D8 in the triplet ground state can mimic the behaviors of transition‐metal Ni atom with electron configuration 3s23p64s23d8 in electronic configuration, physics and chemistry. Interestingly, the characteristic order of 1S1P2S1D for the B6 nonmetal cluster with short B‐B lengths is different from that of the traditional jellium model—1S1P1D2S for metal clusters with long M‐M lengths, which exhibits a novel size effect of nonmetal cluster on jellium orbital ordering. Based on the jellium electron configuration, the B6 with the spin moment value of 2μB is a new all‐nonmetal transition‐metal nickel‐like superatom exhibiting a new kind of all‐nonmetal magnetic superatom. Finding the application of the all‐nonmetal magnetic superatom, we encapsulate the magnetic superatom B6 inside fully hydrogenated fullerene forming a clathrate B6@C60H60 with the spin moment value of 2μB. As the C60H60 cage as a polymerization unit can conserve the spin moment of endohedral B6, the clathrate B6@C60H60 is a new all‐nonmetal magnetic superatom building block. Naturally, magnetic superatom structures of the B6 and B6@C60H60 may be metastable.  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