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 62 毫秒
1.
Polypropylenimine dendrimer (DAB‐Am‐32, generation 4.0) was converted into a macroinitiator DAB‐Am‐32‐Cl via reaction with 2‐chloropropionyl chloride. Monodisperse nanoparticles containing poly(propylene imine)(NH2)32‐polystyrene were prepared by emulsion atom transfer radical polymerization (ATRP) of styrene (St), using the DAB‐Am‐32‐Cl/CuCl/bpy as initiating system. The structure of macroinitiator was characterized by FTIR spectrum, 1H NMR, and 13C NMR. The structure of poly(propylene imine)(NH2)32‐polystyrene was characterized by FT‐IR spectrum and 1H NMR; the molecular weight and molecular weight distribution of poly(propylene imine)(NH2)32‐polystyrene were characterized by gel permeation chromatograph (GPC). The morphology, size and size distribution of the nanoparticles were characterized by photon correlation spectroscopy (PCS), transmission electron microscopy (TEM), and atomic force microscopy (AFM). The effects of monomer/macroinitiator ratio and surfactant concentration on the size and size distribution of the nanoparticles were investigated. It was found that the diameters of the nanoparticles were smaller than 100 nm (30–80 nm) and monodisperse; moreover, the particle size could be controlled by monomer/macroinitiator ratios and surfactant concentration. With the increasing of the ratio of St/DAB‐Am‐32‐Cl, the number‐average diameter (Dn), weight‐average diameter (Dw) were both increased gradually. With enhancing the surfactant concentration, the measured Dh of the nanoparticles decreased, while the polydispersity increased. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 2892–2904, 2009  相似文献   

2.
The reduction of 2‐cyanopyridine by sodium in dry methanol in the presence of thiosemicarbazide produces 2‐pyridineformamide thiosemicarbazone, HAm4DH. The reactions of the potentially tridentate ligand HAm4DH with salts of Zn, Cd, and Hg gave a variety of metal‐ligand complexes. The complexes were characterized by mass spectrometry as well as IR and multinuclear NMR (1H, 13C, 13C CP/MAS, 113Cd, 199Hg) spectroscopy. The crystal structures of [Zn(Am4DH)(OAc)]2·H2O, [Hg(HAm4DH)2Br2]·C2H5OH and [Hg(μ‐S‐Am4DH)Br] were obtained. Coordination of anionic Am4DH? occurs through the pyridyl nitrogen, imine nitrogen and thiolato sulfur atoms, while the neutral ligands in [Hg(HAm4DH)2Br2] coordinate as monodentate ligands through their thione sulfur atoms. One of the acetate ligands in [Zn(Am4DH)(OAc)]2·H2O is bridging monodentate and the other bridging bidentate. [Hg(μ‐S‐Am4DH)Br] features five‐coordinate mercury centers with bridging thiolato sulfur atoms. The intermolecular arrangement is dictated by hydrogen bonding from the amino groups and by π‐π stacking of the pyridine rings.  相似文献   

3.
The crystallization of terbium 5,5′‐azobis[1H‐tetrazol‐1‐ide] (ZT) in the presence of trace amounts (ca. 50 Bq, ca. 1.6 pmol) of americium results in 1) the accumulation of the americium tracer in the crystalline solid and 2) a material that adopts a different crystal structure to that formed in the absence of americium. Americium‐doped [Tb(Am)(H2O)7ZT]2 ZT⋅10 H2O is isostructural to light lanthanide (Ce–Gd) 5,5′‐azobis[1H‐tetrazol‐1‐ide] compounds, rather than to the heavy lanthanide (Tb–Lu) 5,5′‐azobis[1H ‐tetrazol‐1‐ide] (e.g., [Tb(H2O)8]2ZT3⋅6 H2O) derivatives. Traces of Am seem to force the Tb compound into a structure normally preferred by the lighter lanthanides, despite a 108‐fold Tb excess. The americium‐doped material was studied by single‐crystal X‐ray diffraction, vibrational spectroscopy, radiochemical neutron activation analysis, and scanning electron microcopy. In addition, the inclusion properties of terbium 5,5′‐azobis[1H‐tetrazol‐1‐ide] towards americium were quantified, and a model for the crystallization process is proposed.  相似文献   

4.
The presence of water has been shown to deeply impact the stability and geometry of Zn complexes in solution. Evidence for tetra‐ and penta‐coordinated species in a pyridylmethylamine–ZnII model complex is presented. Novel 1H NMR tools such as T1‐filtered selective exchange spectroscopy and pure shifted gradient‐encoded selective refocusing as well as classical 2D (1H–1H) exchange spectroscopy, diffusion‐ordered spectroscopy and T1(1H) measurements, in combination with density functional theory methods allow the full conformational dynamics of a pyridylmethylamine–ZnII complex to be revealed. Four conformers and two families of complexes depending on the hydration states are elucidated.  相似文献   

5.
The synthesis and the characterization of two porphyrin coordination cages are reported. The design of the cage formation is based on the coordination of silver(I) ions to the pyridyl units of 3‐pyridyl appended porphyrins. 1H/109Ag NMR spectroscopy, and diffusion‐ordered spectroscopy (DOSY) experiments demonstrate that both the free base porphyrin 2H‐TPyP and the Zn‐porphyrin Zn‐TPyP form the closed cages, [ Ag4(2H‐TPyP)2 ]4+ and [ Ag4(Zn‐TPyP)2 ]4+, respectively, upon addition of two equivalents of Ag+. The complexation processes are characterized in details by means of absorption and emission spectroscopy in diluted CH2Cl2 solutions. The data are discussed in the frame of the point‐dipole exciton coupling theory; the two porphyrin monomers, in fact, experience a rigid face‐to‐face geometry in the cages and a weak inter‐porphyrin exciton coupling. An intermediate species is observed, for Zn‐TPyP , in a porphyrin/Ag+ stoichiometric ratio of about 1:0.5 and is tentatively ascribed to an oblique open form. The occurrence of a photoinduced electron‐transfer reaction within the cages is excluded on the basis of the experimental outcomes and thermodynamic evaluations. Photophysical experiments evidence different reactivities of singlet and triplet excited states in the assemblies. A lower fluorescence quantum yield and triplet formation is discussed in relation to the constrained geometry of the complexes. Unusually long triplet excited state lifetimes are measured for the assemblies.  相似文献   

6.
Ten examples of unsymmetrically benzannulated, boron‐doped polycyclic aromatic hydrocarbons (B‐PAHs) were prepared by a one‐pot protocol using 4,5‐dichloro‐1,2‐bis(trimethylsilyl)benzene ( 1 ), BBr3, and selected PAHs—among them anthracene, benzo[a ]pyrene, biphenylene, and fluoranthene. After mesitylation at the boron centers, the resulting air‐ and water‐stable products were investigated by 1H/11B{1H}/13C{1H} NMR spectroscopy, X‐ray crystallography, cyclic voltammetry, and UV/Vis absorption/emission spectroscopy. The experiments were augmented by DFT calculations. Most of the B‐PAHs are brightly luminescent (Φ PL up to 90 %) and undergo reversible reduction at moderate half‐wave potentials. The two chloro substituents of 1 are not only mandatory for accomplishing efficient diborylation, but can subsequently be used for Stille‐type coupling reactions to introduce 2‐thienyl moieties. Alternatively, Cl/H exchange is achievable with HSiEt3 in a quantitative, Pd‐catalyzed transformation.  相似文献   

7.
The possible stable forms and molecular structures of 1‐cyclohexylpiperazine (1‐chpp) and 1‐(4‐pyridyl)piperazine (1‐4pypp) molecules have been studied experimentally and theoretically using nuclear magnetic resonance(NMR) spectroscopy. 13C, 15N cross‐polarization magic‐angle spinning NMR and liquid phase1H, 13C, DEPT, COSY, HETCOR and INADEQUATE NMR spectra of 1‐chpp (C10H20N2) and 1‐4pypp (C9H13N2) have been reported. Solvent effects on nuclear magnetic shielding tensors have been investigated using CDCl3, CD3 OD, dimethylsulfoxide (DMSO)‐d6, (CD3)2CO, D2O and CD2Cl2. 1H and 13C NMR chemical shifts have been calculated for the most stable two conformers, equatorial–equatorial (e–e) and axial–equatorial (a–e) forms of 1‐chpp and 1‐4pypp using B3LYP/6‐311++G(d,p)//6‐31G(d) level of theory. Results from experimental and theoretical data showed that the molecular geometry and the mole fractions of stable conformers of both molecules are solvent dependent. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
3‐Phenyl‐ and 3‐(p‐methoxyphenyl)‐7,8‐dihydroxy and ‐6,7‐dihydroxychromenones were prepared from ethyl 3‐oxo‐2‐phenylpropanoate, ethyl 3‐oxo‐2‐(4‐methoxyphenyl)‐propanoate and the trihydroxy benzenes in H2SO4. 3‐Aryl‐7,8‐ and 3‐aryl‐6,7‐dihydroxy‐2H‐chromenones reacted with the bis‐dihalides of poly‐glycols in DMF/MeCO3 to afford 12‐Crown‐4, 15‐Crown‐4 and 18‐Crown‐6‐chromenones. The products were identified with IR, 1H NMR, low and high resolution mass spectroscopy and elemental analysis. Some 1:1 cation association constants, Kb, of the 3‐phenyl chromenone crown ethers with Li+, Na+, K+ and Rb+ cations were studied by steady state emission fluorescence spectroscopy; Kb chromenone‐crown complexes displayed crown ether‐cation binding selectivity rules properly in acetonitrile.  相似文献   

9.
The first experimental evidence that fullerenes react with alkali‐metal trichloroacetates through a nucleophilic addition‐substitution route, yielding dichloromethylenefullerenes as the final products, is reported. The intermediates, C60(CCl3)? and C70(CCl3)? anions, have been isolated in their protonated forms as ortho‐C60(CCl3)H, as well as three ortho and one para isomer of C70(CCl3)H. The structures were unambiguously determined by means of 1H, 13C, and 1H–13C HMBC NMR spectroscopy along with UV/Vis spectroscopy. The observed regiochemistry was analyzed with the aid of quantum chemical calculations. Conversion of the protonated compounds into the [6,6]‐closed C60/70(CCl2) cycloadducts under basic conditions can be effected only for the ortho isomers, whereas para‐C70(CCl3)H decomposes back into pristine C70.  相似文献   

10.
The new electrophilic trifluoromethylating 1‐(trifluoromethyl)‐benziodoxole reagents A and B (Scheme 1) have been used to selectively attach CF3 groups to the S‐atom of cysteine side chains of α‐ and β‐peptides (up to 13‐residues‐long; products 7 – 14 ). Other functional groups in the substrates (amino, amido, carbamate, carboxylate, hydroxy, phenyl) are not attacked by these soft reagents. Depending on the conditions, the indole ring of a Trp residue may also be trifluoromethylated (in the 2‐position). The products are purified by chromatography, and identified by 1H‐, 13C‐, and 19F‐NMR spectroscopy, by CD spectroscopy, and by high‐resolution mass spectrometry. The CF3 groups, thus introduced, may be replaced by H (Na/NH3), an overall Cys/Ala conversion. The importance of trifluoromethylations in medicinal chemistry and possible applications of the method (spin‐labelling, imaging, PET) are discussed.  相似文献   

11.
Three Lewis acid–base adducts t‐Bu3Ga–EPh3 (E = P 1 , As 2 , Sb 3 ) were synthesized by reactions of Ph3E and t‐Bu3Ga and characterized by heteronuclear NMR (1H, 13C (31P)) and IR spectroscopy, elemental analysis and single crystal X‐ray diffraction. Their structural parameters are discussed and compared to similar t‐Bu3Ga adducts. The strength of the donor‐acceptor interactions within 1 – 3 was investigated in solution by temperature‐dependent 1H NMR spectroscopy and by quantum chemical calculations.  相似文献   

12.
Four novel [60]fullerene pyrrolidines containing trifluoromethyl (? CF3) group have been synthesized via 1,3‐dipolar cycloaddition reaction, which have been characterized by UV‐Vis spectroscopy, fourier transform infrared spectroscopy, matrix‐assisted laser desorption ionization‐time of flight mass spectroscopy, and 1H, 13C, 19F nuclear magnetic resonance spectrometer (1H NMR, 13C NMR, 19F NMR). Their optical and electrochemical properties have been studied, and the results show that those fulleropyrrolidines containing ? CF3 group have good fluorescence and electrochemical properties. Compared with C60, they have negative shifts in varying degrees for half‐wave potentials, and may have potential applications for photovoltaic conversion materials since their lowest unoccupied molecular orbital (LUMO) levels are close to that of [6,6]‐phenyl‐C61‐butyric acid methyl ester.  相似文献   

13.
A microcrystalline carboxyl‐functionalized imidazolium chloride, namely 1‐carboxymethyl‐3‐ethylimidazolium chloride, C7H11N2O2+·Cl, has been synthesized and characterized by elemental analysis, attenuated total reflectance Fourier transform IR spectroscopy (ATR‐FT‐IR), single‐crystal X‐ray diffraction, thermal analysis (TGA/DSC), and photoluminescence spectroscopy. In the crystal structure, cations and anions are linked by C—H…Cl and C—H…O hydrogen bonds to create a helix along the [010] direction. Adjacent helical chains are further interconnected through O—H…Cl and C—H…O hydrogen bonds to form a (10) layer. Finally, neighboring layers are joined together via C—H…Cl contacts to generate a three‐dimensional supramolecular architecture. Thermal analyses reveal that the compound melts at 449.7 K and is stable up to 560.0 K under a dynamic air atmosphere. Photoluminescence measurements show that the compound exhibits a blue fluorescence and a green phosphorescence associated with spin‐allowed (1π←1π*) and spin‐forbidden (1π←3π*) transitions, respectively. The average luminescence lifetime was determined to be 1.40 ns for the short‐lived (1π←1π*) transition and 105 ms for the long‐lived (1π←3π*) transition.  相似文献   

14.
The trans‐bis(trimethylsilyl)chalcogenolate palladium complexes, trans‐[Pd(ESiMe3)2(PnBu3)2] [E = S ( 1 ) and Se ( 2 )] were synthesized in good yields and high purity by reacting trans‐[PdCl2(PBu3)2] with LiESiMe3 (E = S, Se), respectively. These complexes were characterized by 1H, 13C{1H}, 31P{1H} (and 77Se{1H}) NMR spectroscopy and single‐crystal X‐ray analysis. The reaction of 2 with propionyl chloride led to the formation of trans‐[Pd(SeC(O)CH2CH3)2(PnBu3)2] ( 3 ), a trans‐bis(selenocarboxylato) palladium complex and thus established a new method for the formation of this type of complex. Complex 3 was characterized by 1H, 13C{1H}, 31P{1H} and 77Se{1H} NMR spectroscopy and a single‐crystal X‐ray structure analysis.  相似文献   

15.
The binary systems of iron(II) and iron(III) with 2-pyridineformamide thiosemicarbazone (H2Am4DH) and its N(4)-methyl (H2Am4Me), N(4)-ethyl (H2Am4Et) and N(4)-phenyl (H2Am4Ph) derivatives were studied in aqueous solution by pH-potentiometry, ultraviolet–visible spectroscopy and EPR spectra. The formation constants of the iron(II) and iron(III) complexes were calculated from potentiometric and electronic absorption data at 25 °C and ionic strength μ = 0.1 mol·L?1 using the HYPERQUAD program. The values of the formation constant of the FeL species decrease in the order Fe:H2Am4DH > Fe:H2Am4Me ≈ Fe:H2Am4Et > Fe:H2Am4Ph in the same way as the basicity of the ligands. The species distribution diagrams show that the species FeL2 predominates at physiological pH in the Fe:H2Am4DH, Fe:H2Am4Me and Fe:H2Am4Et systems. The similar EPR spectra of these iron(III) binary systems indicate the same coordination spheres around the metallic center and the EPR g values suggests that the unpaired electron is in the dxy orbital, indicating a d xz 2 d yz 2 d xy 1 ground state configuration for the complexes. For the Fe(III):H2Am4Ph system the EPR results indicated dimerization and antiferromagnetic interaction due to the presence of only one thiosemicarbazone ligand around the metallic center.  相似文献   

16.
A novel series of pyrimidine/pyrazole linked β‐lactams have been synthesized in excellent yields using a simple and efficient methodology involving conjunction of different heterocyclic substrates. All the new products were characterized on the basis of various spectroscopic techniques viz. FT‐IR, 1H NMR, 13C NMR, elemental analysis (CHN), 1H‐1H correlation spectroscopy (1H‐1H COSY) and mass spectrometry (EIMS) in representative cases. Furthermore, theoretical calculations have also been performed on representative compounds and the results were compared with Cefuroxime axetil (a broad spectrum antibacterial agent). The phenomenon of tautomerism was also observed which was confirmed by different NMR experiments (D2O exchange study and 1H‐1H correlation spectroscopy).  相似文献   

17.
The reaction of [Fe3(CO)12] with bis[2‐(diphenylphosphino)phenyl]ether (DPEphos) in refluxing THF afforded a mononuclear complex, [Fe(CO)41‐P‐DPEphos)] (1), as major product and a binuclear complex, [Fe2(CO)6(μ‐CO)(μ‐P,P‐DPEphos)] (2), as minor product respectively. The DPEphos ligand acts as a terminal P‐donor in complex 1 and a bridging P,P‐donor in complex 2. Complexes 1 and 2 were characterized by elemental analysis, fast atom bombardment mass spectrometry, FT‐IR, 1H and 31P{1H} NMR spectroscopy. The structure of complex 1 has been tentatively assigned by density functional theory calculations and its analogy with reported complexes. Combination of complex 1 and PdCl2 furnished an active catalyst for the Suzuki–Miyaura cross‐coupling reactions of various aryl halides with arylboronic acids. Interestingly, under the same experimental condition, complex 1/PdCl2 as catalyst showed superior activity over the DPEphos/PdCl2 system. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
The terminal zinc hydride complex [Tntm]ZnH ( 2 ; Tntm=tris(6‐tert‐butyl‐3‐thiopyridazinyl)methanide) is an efficient hydrosilylation catalyst of CO2 at room temperature without the need of Lewis acidic additives. The inherent electrophilicity of the system leads to selective formation of the monosilylated product (MeO)3SiO2CH (at room temperature with a TOF of 22.2 h?1 and at 45 °C with a TOF of 66.7 h?1). In absence of silanes, the intermediate formate complex [Tntm]Zn(O2CH) ( 3 ) is quantitatively formed within 5 min. All complexes were fully characterized by 1H and 13C NMR spectroscopy and single‐crystal X‐ray diffraction analyses. Density functional theory (DFT) calculations reveal a high positive charge on zinc and the increased preference of the ligand to adopt a κ3‐coordination mode.  相似文献   

19.
The ion pair of the stereolabile C3‐symmetric, i+o proton complex [ 1? H]+ of diaza‐macropentacycle 1 and the configurationally stable Δ‐TRISPHAT ([Δ‐ 3 ]?) anion exists in the form of two diastereomers, namely, [Δ‐( 1? H)][Δ‐ 3 ] and [Λ‐( 1? H)][Δ‐ 3 ], the ratio of which, in terms of diastereomeric excess (de) decreases in the order [D8]THF (28 %)>CD2Cl2 (22 %)>CDCl3 (20 %)>[D8]toluene (16 %)>C6D6 (7 %)>[D6]acetone (0 %) at thermodynamic equilibrium. Except in the case of [D6]acetone, the latter is reached after a period of time that increases from 1 h ([D8]THF) to 24 h (CDCl3). Moreover, the initial value of the de of [ 1? H][Δ‐ 3 ] in CDCl3, before the thermodynamic equilibrium is reached, depends on the solvent in which the sample has been previously equilibrated (sample “history”). This property has been used to show that the crystals of [ 1? H][Δ‐ 3 ] formed by slow evaporation of CH2Cl2/CH3OH mixtures had 100 % de, which indicates that [ 1? H][Δ‐ 3 ] has enjoyed a crystallization‐induced asymmetric transformation. Structural studies in solution (NMR spectroscopy) and in the gas phase by calculations at the semiempirical PM6 level of theory suggest that the optically active anion is docked on the i+ (endo) external side of the proton complex such that one of the aromatic rings of [Δ‐ 3 ]? is inserted into a groove of [ 1? H]+, a second aromatic ring being placed astride the outside i+ pocket. Solvent polarity controls the thermodynamics of inversion of the [ 1? H]+ propeller. However, both polarity and basicity control its kinetics. Therefore, the rate‐limiting steps correspond to the ion‐pair separation/recombination and [ 1? H]+/ 1 deprotonation/protonation processes, rather than the inversion of [ 1? H]+, the latter being likely to take place in the deprotonated form ( 1 ).  相似文献   

20.
A novel optically active amphiphilic diblock copolymer bearing quinine pendants poly(ethylene oxide)‐b‐poly(glycidyl triazolyl‐L ‐quinine) (MPEO‐b‐PGTQ) was synthesized by “click” reaction of alkyne‐modified diblock copolymer poly(ethylene oxide)‐b‐poly(glycidyl propargyl ether) (MPEO‐b‐PGPE) and 9‐N3‐quinine. The structure and composition of copolymers were characterized by gel permeation chromatography, 1H nuclear magnetic resonance spectroscopy (1H NMR), elemental analysis and optical rotation measurements, which showed that the synthetic route could provide the copolymer with well‐defined composition and with similar optical activity compared to its parent quinine. The micellization behavior of this chiral copolymer was investigated in different solvent systems. The results from fluorescence spectroscopy, UV spectroscopy, dynamic light scattering, transmission electron microscopy, 1H NMR and circular dichroism (CD) spectroscopy indicated that the MPEO‐b‐PGTQ could form regular chiral spherical micelles in H2O and Tetrahydrofuran‐H2O (10:90, V/V) systems, and the state of aggregated chiral micelles depended on the nature of the medium. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 3640–3650, 2009  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