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 31 毫秒
1.
The molecular and crystal structures of 1-(4-fluorophenyl)-1,4-dihydro-1H-tetrazole-5-thione (I) and its complex with cadmium(II) (II) are studied by single crystal XRD. Free ligand I is thione; it has a nonplanar structure (the torsion angle between the tetrazole and benzene rings is 54.99(7)°) and forms H-bonded centrosymmetric dimers via two N–H…S hydrogen bonds in the crystal. The dimers contain a central planar eight-membered {S=C–N–H…S=C–N–H…} ring. Complex II has a chain structure with the composition [(C7H4N4FS)2Cd]n. The environment of the Cd(II) atom consists of two nitrogen atoms and two sulfur atoms from four ligands I and represents a distorted tetrahedron. When complex II forms, ligand I converts into the thiol form. Infinite 1D chains contain eight-membered {←S=C–N–Cd←S=C–N–Cd} rings in a chair conformation. The chains in the crystal are arranged in layers parallel to the (101) plane due to secondary intermolecular F…F and π–π-stacking interactions.  相似文献   

2.
2-N-Tosylaminobenzaldehyde ferrocenoylhydrazone (H2L) and the octahedral nickel(II) complex [Ni(HL)2] · 2CH3OH (I) are synthesized and structurally characterized (CIF files CCDC 981876 (H2L) and 981877 (I)). The crystal structures of both compounds include two independent molecules with different mutual orientations of the tosyl and ferrocene fragments. In a single crystal of H2L, the independent molecules are joined by intermolecular hydrogen bonds into infinite linear chains extended along the crystallographic axis x. The π-stacking interaction between the cyclopentadienyl rings is observed along with hydrogen bonds in a single crystal of complex I.  相似文献   

3.
Free base and cobalt(II) complexes of six meso-tetraphenylporphyrinoids containing nonpyrrolic heterocycles and of three meso-thienylporphyrins were investigated using electrospray ionization tandem mass spectrometry (ESI-MS/MS). Their fragmentation was studied in a quadrupole ion trap as a function of the porphyrinoid macrocycle structure and compared with the fragmentation behavior of the benchmark compound meso-tetraphenylporphyrin. In situ oxidation of the neutral cobalt(II) complexes under ESI conditions produced singly charged cobalt(III) porphyrinoid ions; the free bases were ionized by protonation. For the porphyrinoids with an intact porphyrin core, the major fragmentation pathways observed were the losses of the meso-substituent (for meso-phenyl groups) and characteristic fragmentations of one or more meso-substituents (for the meso-thienyl group). Complex fragmentation pathways were observed for porphyrinoids with modifications to the porphyrin core but chemically reasonable structures could be assigned to most fragments, thus delineating general patterns for the behavior of pyrrole-modified porphyrins under CID conditions.
Figure
?  相似文献   

4.
Compounds with both aromatic and perfluoro aromatic moieties often assemble in the form of π-stacks with alternating arene perfluoroarene arrangements. The 1,4 and 1,3-diamines of N-(2,3,5,6-tetrafluoropyridin-4-yl)benzene (1) and (2), C11H7F4N3, however, show a behavior more commonly observed for diblock perfluoro alkyl derivatives: the fluorinated and the hydrocarbon moieties of the molecules segregate from each other and form fluoro and hydrocarbon microdomains with parallel layers of π-stacked tetrafluoropyridine rings and diamino-phenyl moieties. The structural features of the two derivatives are, despite the different substitution patterns and cell settings, strikingly similar. The main feature in the non-fluorinated domains of 1 and 2 is a set of N–H···H and N–H···π interactions that connect the amino groups of four molecules with each other, which are augmented by some weaker C–H···π and C–H···F interactions, but no π–π stacking is observed. The fluoro microdomains in both 1 and 2 are characterized by extremely short interplanar distances between the tetrafluoropyridine rings of only about 3 Å. The molecules in the layers are shifted sideways against each other by half a pyridyl ring thus bringing the fluorine atoms to atop of the aromatic carbon atoms and vice versa.  相似文献   

5.
The binding sites of two ruthenium(II) organometallic complexes of the form [(η6-arene)Ru(N,N)Cl]+, where arene/N,N = biphenyl (bip)/bipyridine (bipy) for complex AH076, and biphenyl (bip)/o-phenylenediamine (o-pda) for complex AH078, on the peptides angiotensin and bombesin have been investigated using Fourier transform ion cyclotron resonance (FTICR) mass spectrometry. Fragmentation was performed using collisionally activated dissociation (CAD), with, in some cases, additional data being provided by electron capture dissociation (ECD). The primary binding sites were identified as methionine and histidine, with further coordination to phenylalanine, potentially through a π-stacking interaction, which has been observed here for the first time. This initial peptide study was expanded to investigate protein binding through reaction with insulin, on which the binding sites proposed are histidine, glutamic acid, and tyrosine. Further reaction of the ruthenium complexes with the oxidized B chain of insulin, in which two cysteine residues are oxidized to cysteine sulfonic acid (Cys-SO3H), and glutathione, which had been oxidized with hydrogen peroxide to convert the cysteine to cysteine sulfonic acid, provided further support for histidine and glutamic acid binding, respectively.
Fig. a
?  相似文献   

6.
Lili Yin  Yuexin Lin  Li Jia 《Mikrochimica acta》2014,181(9-10):957-965
We show that magnetic nanoparticles can be functionalized with graphene oxide (GO-MNPs) in two reaction steps, and that such nanoparticles can be used as adsorbents for the removal of phthalate esters (PAEs) from water samples. The GO-MNPs were characterized by scanning electron microscopy, transmission electron microscopy, Fourier-transform infrared spectroscopy, zeta potential, and vibrating sample magnetometer. The impacts of contact time, sample pH, ionic strength and sample volume on the adsorption process were investigated. The maximum adsorption capacity for diethyl phthalate was calculated to be 8.71 mg g?1 according to the Langmuir adsorption isotherm. The adsorption efficiency was tested by removal of PAEs. More than 99 % of the total quantity of PAEs (0.12 mg L?1) in 500 mL real water samples can be removed when GO-MNPs (275–330 mg) were used as an adsorbent. In addition, other species (estriol and fluorene) containing benzene rings were also almost completely removed with the PAEs using GO-MNPs, indicating that GO-MNPs are suitable for the removal of the species containing π-electron system through π-π interactions.
Fig. a
Magnetic nanoparticles can be functionalized with graphene oxide (GO-MNPs) in two reaction steps, and that such nanoparticles can be used as adsorbents for the removal of phthalate esters from water samples.  相似文献   

7.
The nature and transmission mechanism of substituent effects in (E)-β-substituted styrenes, C6H5–CH=CH–X, have been investigated from the structural changes induced by a variable substituent on the phenyl group. The molecular structures of 46 (E)-β-substituted styrenes were determined from MO calculations at the B3LYP/6-311++G** level of theory. The structural variation of the phenyl probe is best represented by two orthogonal linear combinations of the internal ring angles, S F STY and S R STY . Regression analysis of S F STY using appropriate explanatory variables reveals a composite field effect, the main component of which originates from the long-range effect of the substituent enhanced by field-induced π-polarization of the vinylene spacer and resonance-induced field effects. The electronegativity of the substituent also plays a role in determining the value of S F STY . Comparison with coplanar 4-substituted biphenyls reveals that the components of the field effect in the two molecular systems are of the same nature (apart from the electronegativity contribution, which is not present in biphenyl derivatives). However, the structural variation of the phenyl probe is more pronounced in (E)-β-substituted styrenes due to the shorter distance between substituent and probe. Analysis of π-charge distribution shows that the aptitude of the substituents to exchange π-electrons with the styrene and coplanar biphenyl frames is nearly the same. Nevertheless, the π-charge variation on the phenyl probe of (E)-β-substituted styrenes is 57 % greater than the corresponding quantity in coplanar 4-substituted biphenyls. Thus, the vinylene spacer is more effective than the phenylene spacer in transmitting π-charges. The S R STY parameter is related to the amount of π-charge transferred from the –CH=CH–X moiety into the π-system of the benzene ring, or vice versa, due to the resonance effect of the variable substituent and, to a lesser extent, to field-induced π-polarization.  相似文献   

8.
Ion mobility mass spectrometry (IMS-MS) is used to investigate the abundance pattern, n z (m) of Poly-(ethyleneglycol) (PEG) electrosprayed from water/methanol as a function of mass and charge state. We examine n z (m) patterns from a diversity of solution cations, primarily dimethylammonium and triethylammonium. The ability of PEG chains to initially attach to various cations in the spraying chamber, and to retain them (or not) on entering the MS, provide valuable clues on the ionization mechanism. Single chains form in highly charged and extended shapes in most buffers. But the high initial charge they hold under atmospheric pressure is lost on transit to the vacuum system for large cations. In contrast, aggregates of two or more chains carry in all buffers at most the Rayleigh charge of a water drop of the same volume. This shows either that they form via Dole’s charge residue mechanism, or that highly charged and extended aggregates are ripped apart by Coulombic repulsion. IMS-IMS experiments in He confirm these findings, and provide new mechanistic insights on the stability of aggregates. When collisionally activated, initially globular dimers are stable. However, slightly nonglobular dimers projecting out a linear appendix are segregated into two monomeric chains. The breakup of a charged dimer is therefore a multi-step process, similar to the Fenn-Consta polymer extrusion mechanism. The highest activation barrier is associated to the first step, where a short chain segment carrying a single charge escapes (ion-evaporates) from a charged drop, leading then to gradual field extrusion of the whole chain out of the drop.
Figure
?  相似文献   

9.
The copper(II) complexes CuLCl2 (1) and CuLBr2 (2), with the chelating pyrazolylpyrimidine ligand 4-(3,5-dimethyl-1H-pyrazol-1-yl)-2-methyl-6-phenylpyrimidine (L), have been synthesized. A single crystal X-ray diffraction study revealed that 1 and 2 have molecular mononuclear structures. The molecules of 1 and 2 form chains in the crystal structures of these compounds due to the formation of π-π-stacking interactions between the pyrimidine and the phenyl rings. The complexes, in combination with the co-catalyst methylaluminoxane (MAO), reveal catalytic activity in ethylene polymerization, while the free ligand L is inactive.  相似文献   

10.
Herein, we report on the synthesis of film-forming poly(styrene-co-butyl acrylate-co-acrylic acid)/SiO2 [P(St-BA-AA)/SiO2] nanocomposites by in situ formation of SiO2 nanoparticles from TEOS via sol–gel process in the presence of poly(acrylic acid) (PAA)-functionalized poly(styrene-co-butyl acrylate) [P(St-BA)] particles fabricated by soap-free emulsion polymerization. The formed silica particles could be absorbed by polyacrylate chains on the surface of PAA-functionalized P(St-BA) particles; thus, raspberry-like polymer/silica nanocomposites would be obtained. Transmission electron microscopy, Fourier transform infrared spectroscopy, attenuated total reflectance infrared spectrum, ultraviolet–visible transmittance spectra, and thermogravimetric analysis were used to characterize the resulting composites. The results showed that the hybrid polymer/silica had a raspberry-like structure with silica nanoparticles anchored on the surface of polymer microspheres. The thermal, fire retardant, and mechanical properties and water resistance of the film were improved by incorporating silica nanoparticles, while the optical transmittance was seldom affected due to nanosized silica particles uniformly dispersed in the film.
Figure
Film-forming polymer/silica nanocomposites with raspberry-like morphology have been successfully prepared via soap-free emulsion polymerization followed by the sol–gel process. The number and the size of SiO2 particles coated on the surface of polymer particles can be adjusted by the amounts of TEOS and ammonia. After the film formation of polymer/silica nanocomposites, silica nanoparticles are homogeneously dispersed within the film without aggregation.  相似文献   

11.
The crystal structure and luminescent properties of boron difluoride 1-naphthyl-butandionate-1,3 (1) (C10H7COCHCOCH3BF2) are determined and analyzed. The molecules of compound 1 are packed in the crystal into infinite stacks due to π-stacking interactions. A feature of the structure of 1 is a fragment of four neighboring molecules, the π-systems of which participate in π…π interactions of different types: naphthyl…naphthyl (head-to-head) and naphthyl…chelate (head-to-tail). The data of time resolved luminescent spectroscopy and quantum chemical computations show that in crystals and concentrated solutions of 1, one excimer structure occurs corresponding to the head-to-tail overlap.  相似文献   

12.
Three novel coordination polymers, {[Co(L)(SO4)(H2O)](CH3OH)} (1), {[Cd(L)2(SiF6)](H2O)} (2) and [Zn(L)(NO3)2] (3), synthesized from 1,4-di(benzimidazole-1-yl)benzene (L), have been characterized by infrared spectroscopy, elemental analysis and single crystal X-ray diffraction. Compounds 1–3 exhibit different structures. Complex 1 has a 3-D diamond network containing 1-D CoII chains connected by SO 4 2? · anions; 2 has a 3-D-Po framework with 1-D porous channels along the c axis; and 3 has a 1-D zig-zag chain structure with a 2-D supramolecular network based on π-π interactions. The magnetic properties of 1 and the solid state fluorescence spectra of 2 and 3 have also been explored.  相似文献   

13.
A Kekulé structure for a benzenoid or a fullerene $\Gamma $ is a set of edges $K$ such that each vertex of $\Gamma $ is incident with exactly one edge in $K$ , i.e. a perfect matching. All fullerenes admit a Kekulé structure; however, this is not true for benzenoids. In this paper, we develop methods for deciding whether or not a given benzenoid admits a Kekulé structure by constructing Kekulé structures that have a high density of benzene rings. The benzene rings of the Kekulé structure $K$ are the faces in $\Gamma $ that have exactly three edges in $K$ . The Fries number of $\Gamma $ is the maximum number of benzene rings over all possible Kekulé structures for $\Gamma $ and the set of benzene rings giving the Fries number is called a Fries set. The Clar number is the maximum number of independent benzene rings over all possible Kekulé structures for $\Gamma $ and the set of benzene rings giving the Clar number is called a Clar set. Our method of constructing Kekulé structures for benzenoids generally gives good estimates for the Clar and Fries numbers, often the exact values.  相似文献   

14.
1,2,3,4,5-Pentafluorodiphenyl disulfide (1) was synthesized from C6F5SCl and C6H5SSiMe3 in quantitative yield. The homo-crystals of disulfide 1 and co-crystals of 1,1′,2,2′,3,3′,4,4′,5,5′-decafluorodiphenyl disulfide (2) with naphthalene (stoichiometry 1:2, complex 4) and diphenyl disulfide (3) with octafluoronaphthalene (stoichiometry 2:1, complex 5) were prepared followed by XRD characterization. In the crystal lattice of 1, face-to-face and face-to-edge PhH/PhF orientations of neighboring rings were observed together with face-to-edge PhF/PhF orientations. For the face-to-face PhH/PhF orientation, the large offset of PhH and PhF groups excludes their π-stacking interaction which is very non-typical of the field. The crystal lattice of 4 reveals standard π-stacking interactions of the arene-polyfluoroarene type. While in the lattice of 4 each PhF ring interacts alternating with naphthalenes, in 5 two disulfides 3 are bridged by one octafluoronaphthalene with only one of the PhH rings of each disulfide interacting with the polyfluoroarene π-system. The large offset of neighboring molecules excludes however their π-stacking interactions in complex 5. An attempt to prepare 2/3 co-crystals failed.  相似文献   

15.
《Polyhedron》2005,24(16-17):2538-2543
A series of directly meso-linked Zn(II) porphyrin dimers were prepared, and properties of their bis(π-radical cation)s were investigated. The redox potentials of the dimers bearing various meso-substituents are distributed in wide range although the potential difference (ΔE) between the first and the second oxidations were similar values. Among these dimers, di-tert-butylphenyl derivative and octyloxyphenyl derivative were chemically oxidized by the treatment with NaAuCl4 in chloroform to afford mono(π-radical cation)s and bis(π-radical cation)s. In the mono(π-radical cation), the π-radical resides one of the porphyrin rings in view of the ESR hyperfine structure. The bis(π-radical cation)s of these dimers showed fine-structured ESR spectra due to triplet species, indicating that the orthogonal alignment of π-radicals leads a stable triplet state. The plots of the triplet ESR signal intensity of the bis(π-radical cation)s versus temperature showed an anomalous feature, in which slight change at lower temperatures and drastic change at higher temperatures were observed. The spectral behavior at lower temperatures was elucidated to be derived from the existence of several kinds of frozen dimers with various dihedral angles between two porphyrin rings. An abrupt increase of triplet signal intensity around 130–160 K can be rationalized by molecular motion around the mesomeso linkage. Such phenomena are comparable with spin-crossover. It also proved that the magnetic interaction in the π-diradical is modulated by the electron-donating ability of the meso-aryl substituents.  相似文献   

16.
The structures and conformations of a 2-[bis(2-ethylphenyl)thiophosphorylhydroxymethyl]-1-ethylbenzimidazole (L) molecule in the free form (Lfr) and in the [ZnCl2L]L complex (I) are compared using the X-ray diffraction data. In complex I, one ligand molecule (Lc) is coordinated to the zinc atom in the bidentate chelating mode and the second molecule (Lsolv) accomplishes the solvating function. An isomer with the cis orientation of the O(1) atom to the N(1) atom is observed in the crystal structure of Lfr, like Lsolv in complex I, with the formation of the five-membered corrugated H-heterocycle OC2NH. Lfr and Lsolv molecules differ in the arrangement of the aromatic rings of the 2R substituent of benzimidazole. Crystalline compound L forms dimeric associates in which three conjugate H-rings are closed by a bifurcate intermolecular hydrogen bond: two five-membered ONC2H rings and one planar four-membered N2H2 ring ((N(1)…N(1A), 3.705 and H(1)…H(1A) 2.67 Å). Unlike crystals of compound I with the π-π-stacking interaction between the planar-parallel aromatic rings of benzimidazole of the Lsolv and Lc molecules, the crystals of L have no similar interaction between the Lfr molecules.  相似文献   

17.
In the present study, the molecular chain changes and structural transitions of partially hydrolyzed poly(vinyl alcohol) (PVA) having a 12 mol% acetate unit were analyzed by moving-window two-dimensional (MW2D) correlation infrared spectroscopy combined with differential scanning calorimetry and thermogravimetric analysis. The results show the glass-transition temperature (T g ) of PVA is clearly distinguished by MW2D correlation infrared spectroscopy, and the acetate groups start to be eliminated around the melting temperature, whereas the free water molecules in PVA are eliminated above T g. The correlation movements of the O–H stretching modes, including the free hydroxyl groups and the hydrogen bonds, are clearly determined using MW2D correlation infrared spectroscopy. The spectral variations in the C=O stretching region caused by the elimination of the acetate unit from polymer chains are also discussed on the basis of the results of the MW2D correlation analysis. Such results cannot be obtained by traditional infrared spectroscopy owing to the complex overlapping peaks.
Figure
The structural variations of partially hydrolyzed poly(vinyl alcohol) studied by moving-window two-dimensional correlation infrared spectroscopy  相似文献   

18.
19.
Three hydroxypropyl methylcellulose samples (HPMC1–3, DSMe?=?1.45, 1.29, and 1.36; MSHP?=?0.28, 0.46, and 0.84) were analyzed with respect to their methyl and hydroxypropyl substitution pattern in the polymer chains. Ionization yield of HPMC oligomers in electrospray ionization ion trap mass spectrometry (ESI-IT-MS) is strongly influenced by the hydroxypropyl pattern. Therefore, a sample derivatization procedure, as well as suitable measurement conditions that enable relative quantification were elaborated. Analysis was performed by negative ESI-IT-MS after per(deutero)methylation, partial depolymerization, and reductive amination with m-aminobenzoic acid. Measurement parameters like solvent, trap drive, and voltages of the ion transportation unit were studied with regard to the suitability for quantitative evaluation. Using direct infusion of the samples, strong influence of trap drive and octopole settings was observed. Optimized measurement conditions were used for the determination of the HP pattern of the permethylated samples by direct infusion. The methyl pattern was determined from the perdeuteromethylated samples by high-performance liquid chromatography–electrospray tandem mass spectrometry. For HPMC1, substituents were both found to fit the random distribution model. The other two samples showed pronounced heterogeneity which could be interpreted in more detail by extracting methyl subpatterns depending on the number of HP groups.
Figure
Determination of the substituent distribution in hydroxypropyl methylcellulose oligomers by ESI-IT-MS (negative mode) after labeling with m-aminobenzoic acid  相似文献   

20.
Novel imidazole fluorescent ionic liquids with anthracene groups (ImS-FILA) were synthesized for the first time to act as fluorescent probes. They were developed for the determination of superoxide anion radicals (O2 ?-) in an aqueous system. O2 ?- was produced by pyrogallol autoxidation. The fluorescence of ImS-FILA was quenched by superoxide anion radicals. The π-bond structure of the fluorescent molecules was oxidized and damaged. This method is very simple and sensitive. The linear range of sensitivity was 1–70 μM ImS-FILA, and the detection limit for reactive oxygen species was 0.1 μM. This method was used to detect superoxide radicals in papaya and garlic, with satisfactory results. Further work is needed to demonstrate the utility of this method in detecting reactive oxygen species in a biological aqueous system.
Figure
Reaction of fluorescent probes with O2 ?- and the fluorescence change  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