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.
《Analytical letters》2012,45(8):872-882
A new, rapid, and simple method is proposed for the determination of the pesticide napropamide by photo-induced chemiluminescence detection coupled with a flow injection analysis (FIA) system. The emission was obtained by oxidation with periodate in basic medium, of the photoproducts generated on-line by UV irradiation (254 nm) of napropamide in acidic SDS (sodium dodecyl sulfate) medium. The flow method, in combination with the solid phase extraction (SPE) performed off-line with C18 cartridges, allowed the determination of this pesticide over the 0.8–14.0 µ gL?1 range, with a limit of detection of 0.3 µ gL?1. The relative standard deviation (n = 9) at 2.5 µ gL?1 level was 4.3% for the combined FIA-SPE system. After testing the influence of several potential interfering compounds, including ions and other pesticides, the method was successfully applied to the determination of napropamide in spiked water samples with recoveries between 96–103%.  相似文献   

2.
The nitroarene products of the gas-phase reactions of acenaphthylene, acenaphthene, phenanthrene, and anthracene-d10 with N2O5 and the OH radical (in the presence of NOx) are reported. The calculated atmospheric lifetimes of these polycyclic aromatic hydrocarbons (PAH), as well as those of naphthalene, 1- and 2-methylnaphthalene, biphenyl, fluoranthene, pyrene, and acephenanthrylene, show that reaction with the OH radical is the dominant loss process for these PAH, with the exception of acenaphthylene, acenaphthene, and acephenanthrylene which contain an external cyclopenta-fused ring. For these latter PAH, reaction with the NO3 radical, and for acenaphthylene and acephenanthrylene reaction with O3, are also expected to be important atmospheric loss processes. The nitroarenes observed as products of the atmospherically-important gas-phase reactions of the PAH in environmental chamber studies are compared with the nitroarenes measured in ambient air samples collected in California. It is concluded that although nitroarenes are formed in low yields (?5%) from the OH radical-initiated reactions of the PAH, atmospheric formation of nitroarenes may contribute significantly to ambient nitroarene concentrations.  相似文献   

3.
A dispersive liquid‐liquid microextraction method based on the dispersion of 1,2‐dichlorobenzene as an extraction solvent into an aqueous phase in the presence of ethanol as a dispersive solvent for the preconcentration of Co2+ and Ni2+ ions is discussed. 1‐Nitroso 2‐naphtol was used as a chelating agent prior to the extraction and the preconcentrated analyte was determined by flame atomic absorption spectrometry. The effect of various experimental parameters including the extraction and dispersive solvent type and volume, pH, amount of the chelating agent, etc. on the microextraction and complex formation was investigated for finding the optimum conditions. The enhancement factors were about 61.9 and 51.8, the calibration graphs were linear in the range of 10‐150 μgL?1 and 10‐250 μgL?1 with detection limits of 2.42 μgL?1 and 1.59 μgL?1, and RSD (n = 5) of 3.08% and 2.17% for cobalt and nickel, respectively. The method was successfully applied to the determination of Co and Ni in water and vitamin B12.  相似文献   

4.
Benzo(a)pyrene (BaP) has been recognized as a marker for the detection of carcinogenic polycyclic aromatic hydrocarbons. In this work, a novel monolithic solid-phase extraction (SPE) sorbent based on graphene oxide nanoparticles (GO) in starch-based cryogel composite (GO-Cry) was successfully prepared for BaP analysis. Rice flour and tapioca starch (gel precursors) were gelatinized in limewater (cross-linker) under alkaline conditions before addition of GO (filler) that can increase the ability to extract BaP up to 2.6-fold. BaP analysis had a linear range of 10 to 1000 µgL−1 with good linearity (R2 = 0.9971) and high sensitivity (4.1 ± 0.1 a.u./(µgL−1)). The limit of detection and limit of quantification were 4.21 ± 0.06 and 14.04 ± 0.19 µgL−1, respectively, with excellent precision (0.17 to 2.45%RSD). The accuracy in terms of recovery from spiked samples was in the range of 84 to 110% with no significant difference to a C18 cartridge. GO-Cry can be reproducibly prepared with 2.8%RSD from 4 lots and can be reused at least 10 times, which not only helps reduce the analysis costs (~0.41USD per analysis), but also reduces the resultant waste to the environment.  相似文献   

5.
Abstract

A facile and efficient method is described for the determination of trace quantities of triazine herbicides, terbutryn, prometryn and ametryn in water. The procedure involved preconcentration of water samples by sorption on chromatographic grade silica gel particles with chemically modified surface, being covalently bonded with a nonofunctional C8H17 group. This was followed by solvent desorption with 2-propanol. The determinative step was achieved by capillary gas chromatography on Supelcowax-10 fused silica column using a nitrogen-phosphorus detector. The limit of detection was 0.1 μg-10 μgL?1.  相似文献   

6.
In the present work, we report the successful synthesis and characterization of six (two new) fullerene mono- and di-pyrene derivatives based on C60 and C70 fullerenes. The synthesized compounds were characterized by spectral methods (ESI-MS, 1H-NMR, 13C-NMR, UV-Vis, FT-IR, photoluminescence and photocurrent spectroscopy). The energy of HOMO and LUMO levels and the band gaps were determined from cyclic voltammetry and compared with the theoretical values calculated according to the DFT/B3LYP/6-31G(d) and DFT/PBE/6-311G(d,p) approach for fully optimized molecular structures at the DFT/B3LYP/6-31G(d) level. Efficiency of solar cells made of PTB7: C60 and C70 fullerene pyrene derivatives were analyzed based on the determined energy levels of the HOMO and LUMO orbitals of the derivatives as well as the extensive spectral results of fullerene derivatives and their mixtures with PTB7. As a result, we found that the electronic and spectral properties, on which the efficiency of a photovoltaic cell is believed to depend, slightly changes with the number and type of pyrene substituents on the fullerene core. The efficiency of constructed solar cells largely depends on the homogeneity of the photovoltaic layer, which, in turn, is a derivative of the solubility of fullerene derivatives in the solvent used to apply these layers by spincoating.  相似文献   

7.
The influence of temperature (77–230 K) on the fate of pyrene (Py) radical ions and Py excited states in irradiated poly(dimethylsiloxane) (PDMS) doped with Py is described. At 77 K, the Py radical ions seem to be stable, whereas the Py excited states [fluorescence (λ = 395 nm) and phosphorescence (λ = 575–650 nm)] are generated via tunneling charge transfer. In the range of the glass‐transition temperature (Tg = 152–153 K), the Py radical ions start to decay, taking part in a recombination process and leading to the Py monomer and Py excimer fluorescence (λ = 475 nm). The wavelength‐selected radiothermoluminescence (WS RTL) observed at approximately 395, 475, and 600 nm has helped us to identify the Tg range (152–153 K). The absorption maximum at approximately 404 nm, found in the temperature range under consideration, is thought to represent PyH?, cyclohexadienyl‐type radicals produced as a result of the reaction of Py?? with protonated PDMS macromolecules. With the initial‐rise method of evaluating the activation energy (Ea) with the WS RTL peaks observed in the Tg range, Ea values of 123–151 kJ mol?1 have been found. Such high Ea values can be explained by the contribution of energy connected to the molecular relaxation of the matrix in the Tg range. The well‐known Williams–Landel–Ferry equation, with universal constants C1 = 17.4 and C2 = 12.7, has been successfully applied to the interpretation of old pulse‐radiolysis/viscosity data found for crosslinked PDMS doped with Py. The mechanisms involved in these phenomena are discussed. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 6125–6133, 2004  相似文献   

8.
The fluorescence of acenaphthylene follows a Stern-Volmer relationship in air-saturated ether giving a self-quenching constant KF=0.12M?1, indicating that the syn photodimer originates from the singlet state of acenaphthylene. By comparison, 1-cyanoacenaphthylene undergoes a more efficient self-quenching (KF= 2.8 M?1) in air-saturated ether. No excimer fluorescence was detected for 1-cyanoacenaphthylene nor the parent compound in solution. The triplet state of 1-cyanoacenaphthylene, obtained by sensitization or induced by heavy atom solvent (EtI), was shown to generate exclusively the head-to-head anti photodimer in a high chemical yield. Regiospecificity and stereospecificity observed in this reaction indicates the influence of the acenaphthylenic ring and the cyano group in stabilizing the diradicaloïd transition state.  相似文献   

9.
The products of the reaction between the electrophilic alkenylxenonium cation [1-Xe+–C6F9] and the halide anions I?, Br?, Cl? and F? depend on the hardness of the halide anion. With the soft halides I? and Br? Xe(II) is formally displaced by halogen as well in basic MeCN as in superacidic (AHF1), whereas with hard fluoride and chloride no reaction takes place in AHF. In MeCN F? initiates the formation of alkenyl radicals, which abstract hydrogen from the solvent, whereas Cl? exhibits borderline character: RH and RCl formation. Possible reaction paths are discussed. The reactivity of the arylxenonium cation [C6F5Xe]+ in AHF toward halide ions is reported and the relative electrophilicity of the cations [C6F5Xe]+ and [1-Xe+–C6F9] is determined by the competitive reaction with Cl?. In addition the synthesis of cyclohexene 1-CF3–C6F9 from C6F5CF3 and XeF2 is performed and its electrophilicity is compared with that of the aromatic compound C6F5CF3.  相似文献   

10.
Polyterephthalamides of high molecular weight (ηinh up to 1.9) were obtained by the direct polycondensation reaction of terephthalic acid and aromatic diamines in the presence of poly(ethylene oxide) (PEO) with triphenyl phosphite in a N-methylpyrrolidone (NMP)–pyridine solution that contained lithium chloride. The molecular weights of the polymers produced varied with the amount and molecular weight of PEO, which showed maximum values when PEO with a molecular weight of 2.0 × 104?5.0 × 105 was used in a concentration of about 0.5 wt % in the solvent. The polycondensation reaction was significantly affected by the level of pyridine in a mixed solvent of NMP and pyridine and by the concentration of the lithium chloride added.  相似文献   

11.
Acridine dyes were found to exhibit a low light fastness in oxygen-containing aqueous solutions because of photooxidation with the participation of molecular oxygen. The values of the quantum yield for the photooxidation of Acridine Yellow, Acridine Orange, and proflavin acetate at their initial concentration 2 × 10?5 mol/l were determined to be 1.8 × 10?3, 1.5 × 10?3, and 0.8 × 10?3, respectively. It was concluded that the photooxidation of these acridine dyes in dilute solutions results in the formation of primary photoproducts of peroxide nature. For proflav in acetate, as an example, it was demonstrated that the increase of the substrate concentration to 6 × 10?4 mol/l results in a fourfold increase in the quantum yield, a behavior indicative of a change in the photodecomposition mechanism. The rate of the interaction between exited and unexcited proflavin molecules, the process responsible for the photodecomposition of this compound, was found to be ~4 × 107 l/(mol s). The quantum yield for the photooxidation of the proflavin base was estimated to exceed that for the photooxidation of the proflavin monocation by more than an order of magnitude.  相似文献   

12.
The use of N‐methylpyrrolidone (NMP) as a co‐solvent in ferric salt catalyzed cross‐coupling reactions is crucial for achieving the highly selective, preparative scale formation of cross‐coupled product in reactions utilizing alkyl Grignard reagents. Despite the critical importance of NMP, the molecular level effect of NMP on in situ formed and reactive iron species that enables effective catalysis remains undefined. Herein, we report the isolation and characterization of a novel trimethyliron(II) ferrate species, [Mg(NMP)6][FeMe3]2 ( 1 ), which forms as the major iron species in situ in reactions of Fe(acac)3 and MeMgBr under catalytically relevant conditions where NMP is employed as a co‐solvent. Importantly, combined GC analysis and 57Fe Mössbauer spectroscopic studies identified 1 as a highly reactive iron species for the selective formation generating cross‐coupled product. These studies demonstrate that NMP does not directly interact with iron as a ligand in catalysis but, alternatively, interacts with the magnesium cations to preferentially stabilize the formation of 1 over [Fe8Me12]? cluster generation, which occurs in the absence of NMP.  相似文献   

13.
Reduction of VCl3(THF)3 (THF is tetrahydrofuran) and NbCl4(THF)2 by alkali metal pyrene radical anion salts in THF affords the paramagnetic sandwich complexes bis[(1,2,3,3a,10a,10b‐η)‐pyrene]vanadium(0), [V(C16H10)2], and bis[(1,2,3,3a,10a,10b‐η)‐pyrene]niobium(0), [Nb(C16H10)2]. Treatment of tris(naphthalene)titanate(2−) with pyrene provides the isoelectronic titanium species, isolated as an (18‐crown‐6)potassium salt, namely catena‐poly[[(18‐crown‐6)potassium]‐μ‐[(1,2‐η:1,2,3,3a,10a,10b‐η)‐pyrene]‐titanate(−I)‐μ‐[(1,2,3,3a,10a,10b‐η:6,7‐η)‐pyrene]], {[K(C12H24O6)][Ti(C16H10)2]}n. The first two compounds have very similar packing, with neighboring molecules arranged orthogonally to one another, such that aromatic donor–acceptor interactions are likely responsible for the specific arrangement. The asymmetric unit contains a half‐occupancy metal center η6‐coordinated to one pyrene ligand, with the full M(pyrene)2 molecule generated by a crystallographic inversion center. In the titanium compound, the cations and anions are in alternating contact throughout the crystal structure, in one‐dimensional chains along the [101] direction. As in the other two compounds, the asymmetric unit contains a half‐occupancy Ti atom η6‐coordinated to one pyrene ligand. Additionally, the asymmetric unit contains one half of an (18‐crown‐6)potassium cation, located on a crystallographic inversion center coincident with the K atom. The full formula units are generated by those inversion centers. In all three structures, the pyrene ligands are eclipsed and sandwich the metals in one of two inversion‐related sites. These species are of interest as the first isolable homoleptic pyrene transition metal complexes to be described in the scientific literature.  相似文献   

14.
Poly-α-methylstyrene (PαMS) was degraded in CHCl3 and CCl4 solution by flash photolysis (λ = 265 nm). The degradation, as detected by light scattering measurements, is caused by the attack on PαMS by solvent radicals, assumed to be formed mainly by energy transfer processes. The direct effect did not lead to detectable main chain cleavage as evidenced by experiments with PαMS dissolved in dioxane or methylene chloride. The time dependence of the decrease of the light scattering intensity (LSI) after the flash was measured. The observed first order lifetime τ(LSI) corresponds to the lifetime of lateral macroradicals P′ that decompose by main chain cleavage (k = 3.5 × 102 sec?1). Molecular oxygen reacts with the lateral macroradicals with k = (5.5 ± 0.5) 105 M?1 sec?1. Only a minor portion of the product of this reaction (PO2′) decays by main chain scission. Thus O2 inhibits main chain scission significantly. By addition of cyclohexane and ethyl mercaptan, the main chain cleavage is inhibited. τ(LSI) was not affected by these compounds in the concentration range investigated ([C6H12]: up to 8.4 M; [C2H5SH]: up to 3 × 10?3 M), indicating that they reacted with solvent radicals which otherwise attack the polymer forming lateral macroradicals.  相似文献   

15.
The photocatalytic oxidation of diclofenac, metoprolol, estrone and chloramphenicol was tested in the tube reactor using different commercially available TiO2. The photocatalysts were characterized using BET, XRD and SEM. The studied photocatalysts differed in SBET, pore volume and rutile presence. It was observed that generally anatase TiO2 possessed the highest activity in the photocatalytic oxidation of diclofenac, chloramphenicol and estrone. The presence of rutile enhanced the photooxidation of metoprolol. In case of the other pollutants, however, rutile diminished the photooxidation efficiency. The most effective in the reduction of the COD parameter of treated water was anatase with 21 nm crystals. The photooxidation of all studied pollutants can be described by the pseudo-first order kinetics with the values ranging from 0.46 × 10?2 min?1 in case of estrone removal over Tytanpol (Z.A. Police, Poland) to 1.87 × 10?2 min?1 for the removal of chloramphenicol over TiO2 21 nm (Sigma-Aldrich). The highest initial reaction rates were obtained for metoprolol removal over TiO2 21 nm (Sigma-Aldrich) 1.9 × 10?6 mol dm3 min?1 being three times higher than that determined for estrone photocatalytic oxidation over TiO2 (Sigma-Aldrich).  相似文献   

16.
A cyclophane is reported incorporating two units of a heptagon‐containing extended polycyclic aromatic hydrocarbon (PAH) analogue of the hexa‐peri‐hexabenzocoronene (HBC) moiety (hept‐HBC). This cyclophane represents a new class of macrocyclic structures that incorporate for the first time seven‐membered rings within extended PAH frameworks. The saddle curvature of the hept‐HBC macrocycle units induced by the presence of the nonhexagonal ring along with the flexible alkyl linkers generate a cavity with shape complementarity and appropriate size to enable π interactions with fullerenes. Therefore, the cyclophane forms host–guest complexes with C60 and C70 with estimated binding constants of Ka=420±2 m ?1 and Ka=(6.49±0.23)×103 m ?1, respectively. As a result, the macrocycle can selectively bind C70 in the presence of an excess of a mixture of C60 and C70.  相似文献   

17.
Polycyclic aromatic hydrocarbons (PAHs) are regarded as key intermediates in the molecular growth process that forms soot from incomplete fossil fuel combustion. Although heavily researched, the reaction mechanisms for PAH formation have only been investigated through bulk experiments; therefore, current models remain conjectural. We report the first observation of a directed synthesis of a PAH under single‐collision conditions. By using a crossed‐molecular‐beam apparatus, phenyl radicals react with C3H4 isomers, methylacetylene and allene, to form indene at collision energies of 45 kJ mol?1. The reaction dynamics supported by theoretical calculations show that both isomers decay through the same collision complex, are indirect, have long lifetimes, and form indene in high yields. Through the use of deuterium‐substituted reactants, we were able to identify the reaction pathway to indene.  相似文献   

18.
The kinetics and stoichiometry of the reaction between C70 fullerene and ozone have been studied. The reaction obeys a bimolecular rate law. The stoichiometric coefficients of the reaction are 1: 12 to 1: 22, depending on reaction conditions. The rate constant at 22°C is 5 × 104 l mol?1 s?1 for the first stage of fullerene conversion and (0.8–0.6) × 104 l mol?1 s?1 for the subsequent stages. Since the stages differ in terms of reaction rate, the original C70 molecules are first involved in the reaction, whereas, at the subsequent stages, all molecules are involved with equal probabilities, irrespective of the number of preceding reaction events in which they have participated.  相似文献   

19.
A copolymer of cholesteryl 6‐(methacryloyloxy)hexanoate and a small amount of (1‐pyrenylmethyl) 6‐(methacryloyloxy)hexanoate (Py‐C5‐MA) was prepared by free radical copolymerization. A copolymer of 1‐eicosanylmethacrylate and a small amount of Py‐C5‐MA was also prepared as a reference copolymer. A wide‐angle X‐ray diffraction pattern for an n‐hexane solution of the cholesterol(Chol)‐containing copolymer showed a peak corresponding to a spacing of 5.3 Å. In n‐hexane, the hydrodynamic radius (Rh) for the Chol‐containing copolymer (Mw = 7.8 × 104) was 8.1 nm, while that of the eicosanyl‐containing copolymer (Mw = 4.9 × 104) was 9.6 nm, Rh for the former being smaller than that for the latter, although Mw for the former was higher than that of the latter. 1H‐NMR spectra of the Chol‐containing polymer in n‐hexane‐d14 indicated a strong restriction of local motions of pendant Chol groups. Fluorescence spectra of the Chol‐containing copolymer in n‐hexane indicated that each pyrene group was isolated from others. In n‐hexane/benzene mixed solutions of the Chol‐containing polymer, the ratio of the intensity of the excimer relative to the monomer emission decreased with increasing the ratio of n‐hexane in the mixed solvent. Electron transfer from N,N‐dimethylaniline to singlet‐excited pyrene chromophores was suppressed in the Chol‐containing copolymer in n‐hexane. The pyrene chromophores exhibited a long triplet lifetime in n‐hexane. These observations led us to conclude that Chol groups formed stacks in n‐hexane, and that the pyrene chromophores were trapped in the Chol stacks, leading to the “protection” of pyrene from the bulk phase. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 47–58, 1999  相似文献   

20.
Herein, a novel broken case‐like carbon‐doped g‐C3N4 photocatalyst was obtained via a facile one‐pot pyrolysis and cost‐effective method using glyoxal‐modified melamine as a precursor. The obtained carbon/g‐C3N4 photocatalyst showed remarkable enhanced photocatalytic activity in the degradation of gaseous benzene compared with that of pristine g‐C3N4 under visible light. The pseudo‐first‐order rate constant for gaseous benzene degradation on carbon/g‐C3N4 was 0.186 hr?1, 5.81 times as large as that of pristine g‐C3N4. Furthermore, a possible photocatalytic mechanism for the improved photocatalytic performance over carbon/g‐C3N4 nanocomposites was proposed.  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