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.
A rapid, specific, and sensitive method utilizing ultra-performance liquid chromatography tandem mass spectrometry was developed and validated to determine albendazole, albendazole sulfoxide, albendazole sulfone, and albendazole 2-aminosulfone in fish muscle tissue. The fish samples were extracted with ethyl acetate, then the organic phase was evaporated to dryness, and the residue was reconstituted in methanol–water solution and cleaned up by n-hexane. Reversed-phase separation of target compounds was achieved using a BEH C18 column and a gradient consisting of 0.2% (v/v) formic acid and methanol. Tandem mass spectrometry analyses were performed on a triple–quadrupole tandem mass spectrometer. In the whole procedure, the isotope-labeled internal standards were used to correct the matrix effect and variations associated with the analysis. The method was validated with respect to linearity, specificity, accuracy, and precision. The method exhibited a linear response from 0.1 to 20 ng mL-1 (r 2 > 0.9985). The limit of quantitation for albendazole (ABZ), albendazole sulfoxide (ABZSO), albendazole sulfone (ABZSO2), and albendazole 2-aminosulfone (ABZ-2-NH2SO2) was 0.1, 0.1, 0.1, and 0.2 ng g-1, respectively. The mean recoveries of ABZ, ABZSO, ABZSO2, and ABZ-2-NH2SO2 spiked at a level of 0.2–5.0 ng g-1 were 95.3–113.7%, and the relative standard deviations of intra- and inter-day measurements were less than 6.38%. The method was later successfully applied to the determination of albendazole and its three metabolites in 60 fish samples collected from local markets.  相似文献   

2.
An HPLC method was developed for the determination of albendazole (ABZ) and its metabolites, a sulfoxide (ABZSO), a sulfone (ABZSO2), and albendazole-2-aminosulfone (ABZ-2-NH2SO2), from yellow perch muscle tissue with adhering skin. The muscle tissue samples were made alkaline with potassium carbonate and extracted with ethyl acetate, followed by a series of liquid-liquid extraction steps. After solvent evaporation, the residue was reconstituted in the initial mobile phase combination of the gradient. The mobile phase consisted of a buffer, 50 mM ammonium acetate (pH = 4.0) in 10% methanol-water, and 100% acetonitrile. The gradient was from 20% acetonitrile to 85% acetonitrile. The analytes were chromatographed on an RP Luna C18(2) column and detected by fluorescence with excitation and emission wavelengths of 290 and 330 nm, respectively. The average recoveries from fortified muscle tissue for ABZ (20-100 ppb), ABZ-SO (20-200 ppb), ABZSO2 (8-100 ppb), and ABZ-2-NH2SO2 (20-100 ppb) were 85, 95, 101, and 86%, respectively, with corresponding CV values of 9, 3, 6, and 4%, respectively. Their LOQ values were 10, 10, 1, and 10 ppb, respectively. The procedure was applied to determine ABZ and its major metabolites in the incurred muscle tissue of yellow perch obtained after orally dosing the fish with ABZ.  相似文献   

3.
Two coordination polymers containing copper ions, [Cu(SO4)(pyz)(H2O)]n (1) and [Cu2(SO4)(pyz)2(H2O)2]n (2) (pyz = pyrazine), have been synthesized and characterized by single-crystal X-ray analyses. Compound 1 was synthesized by the reaction of Cu(SO4) · 5H2O with pyz (ratio = 1:2) in H2O at room temperature. The structure of 1 consists of linear chains of [Cu(pyz)(H2O)]2+, with coordinated sulfate ions bridging the chains. Compound 2 was obtained as dark red blocks from the reaction of Cu(SO4) · 5H2O and pyz (ratio = 1:2) in H2O, after heating to 180 °C in a Teflon autoclave for 48 h. The structure of 2 consists of zigzag chains of [Cu(pyz)(H2O)]+ with sulfate ions. Only the difference in the synthesis temperature, room temperature or 180 °C, determines whether Cu(II) or Cu(I) coordination polymers are formed, with the reduction of Cu(II) to Cu(I) being explained by the Gillard mechanism.  相似文献   

4.
The influence of polycaprolatone-triol (PCL-T) on the thermal degradation properties of soy protein isolate (SPI)-based films was studied by thermogravimetry and infrared spectroscopy under nitrogen atmosphere. The results showed that in the absence of PCL-T the thermal degradation began between 292 °C (pure SPI films) and ca. 264 °C (SPI/SDS films with more than 20% of SDS), and these values decreased further to the range 250-255 °C for SPI/SDS/PCL-T films. At the same time, the temperature of maximum degradation rate (Tmax) decreased from 331 °C (pure SPI film) to ca. 280 °C for SPI/SDS/PCL-T films with 39% PCL-T content. This behavior was also confirmed by the activation energy (E) values associated with the thermal degradation process. Apparently, the low thermal stability of PCL-T as compared to other film constituents, along with its plasticizer characteristics, is responsible for the decreased stability of SPI/SDS/PCL-T films. The FTIR spectra of gas products evolved during the thermal degradation indicated the formation of OH, CO2, NH3 and other saturated compounds, suggesting that the reaction mechanism involved simultaneous scission of the C(O)-O polyester bonds and C-N, C(O)-NH, C(O)-NH2 and -NH2 bonds of the protein.  相似文献   

5.
The thermal decomposition of syngenite, K2Ca(SO4)2·H2O, formed during the treatment of liquid manure has been studied by thermal gravimetric analysis, differential scanning calorimetry, high temperature X-ray diffraction (XRD) and infrared emission spectroscopy (IES). Gypsum was found as a minor impurity resulting in a minor weight loss due to dehydration around 100 °C. The main endothermic dehydration and decomposition stage of syngenite to crystalline K2Ca2(SO4)3 and amorphous K2SO4 is observed around 200 °C. The reaction involves a solid-state re-crystallisation, while water and the K2SO4 diffuse out of the existing lattice. The additional weight loss steps around 250 and 350 °C are probably due to presence of larger syngenite particles, which exhibit slower decomposition due to the slower diffusion of water and K2SO4 out of the crystal lattice. A minor endothermic sulphate loss around 450 °C is not due to the decomposition of syngenite or its products or of the gypsum impurity. The origin of this sulphate is not clear.  相似文献   

6.
建立了同时测定草鱼肉中阿苯达唑及其3种代谢物阿苯达唑亚砜、阿苯达唑砜、阿苯达唑-2-氨基砜的超高效液相色谱-串联质谱(UPLC-MS/MS)分析方法。草鱼肉样品通过碱性乙酸乙酯提取,正己烷净化,超高效液相色谱分离,串联质谱检测,氘代同位素内标定量。本方法分析时间短,在4 min内即可完成4种被分析物的液质联用分析。本方法在0.1~10μg/kg范围内各物质线性良好,线性系数在0.9991~0.9996之间。阿苯达唑、阿苯达唑亚砜、阿苯达唑砜、阿苯达唑-2-氨基砜在0.5~5.0μg/kg添加水平下的平均回收率分别为98.4%~102.6%,97.1%~103.4%,98.8%~103.7%和101.1%~104.2%检出限分别为0.2,0.1,0.2和0.1μg/kg;日内精密度小于4.77%,日间精密度小于3.03%。本方法为阿苯达唑及其代谢物的检测及监控提供了一种快速、灵敏度高、重现性好的定量分析方法。  相似文献   

7.
2-Phenylaniline reacted with Pd(OAc)2 in toluene at room temperature for 24 h in a one-to-one molar ratio and with the system PdCl2, NaCl and NaOAc in a 1 (2-phenylaniline):1 (PdCl2):2 (NaCl):1 (NaOAc) molar ratio in methanol at room temperature for one week to give the dinuclear cyclopalladated compounds (μ-X)2[Pd{κ2-N2′,C1-2-(2′-NH2C6H4)C6H4}]2 [1a (X = OAc) and 1b (X = Cl)] in high yield. Moreover, the reaction between 2-phenylaniline and Pd(OAc)2 in one-to-one molar ratio in acid acetic at 60 °C for 4 h, followed by a metathesis reaction with LiBr, allowed isolation of the dinuclear cyclopalladated compound (μ-Br)2[Pd{κ2-N2′,C1-2-(2′-NH2C6H4)C6H4}]2 (1c) in moderate yield. A parallel treatment, but using monodeuterated acetic acid (DOAc) as solvent in the cyclopalladation reaction, allowed isolation of a mixture of compounds 1c, 1cd1 [Pd{κ2-N2′,C1-2-(2′-NH2C6H4)C6H4](μ-Br)2[Pd{κ2-N2′,C1-2-(2′-NH2C6H4)-3-d-C6H3] and 1cd2 (μ-Br)2[Pd{κ2-N2′,C1-2-(2′-NH2C6H4)-3-d-C6H3}]2 in moderate yield and with a deuterium content of ca. 60%. 1a and 1b reacted with pyridine and PPh3 affording the mononuclear cyclopalladated compounds [Pd{κ2-N2′,C1-2-(2′-NH2C6H4)C6H4}(X)(L)] [2a (X = OAc, L = py), 2b (X = Cl, L = py), 3a (X = OAc, L = PPh3) and 3b (X = Cl, L = PPh3)] in a yield from moderate to high. Furthermore, 1a reacted with Na(acac) · H2O to give the mononuclear cyclopalladated compound 4 [Pd{κ2-N2′,C1-2-(2′-NH2C6H4)C6H4}(acac)] in moderate yield. 1H NMR studies in CDCl3 solution of 2a, 2b, 3a, 3b and 4 showed that 2a and 3a presented an intramolecular hydrogen bond between the acetato ligand and the amino group, and were involved in a dynamic equilibrium with water present in the CDCl3 solvent; and that the enantiomeric molecules of 2b and 4 were in a fast exchange at room temperature, while they were in a slow exchange for 2a, 3a and 3b. The X-ray crystal structures of 3b and 4 were determined. 3b crystallized in the triclinic space group with a = 9.9170(10), b = 10.4750(10), c = 12.0890(10) Å, α = 98.610(10)°, β = 94.034(10)° and γ = 99.000(10)° and 4 in the monoclinic space group P21/a with a = 11.5900(10), b = 11.2730(10), c = 12.2150(10) Å, α = 90°, β = 107.6560(10)° and γ = 90°.  相似文献   

8.
The cloud point of nonionic surfactant polyoxyethylene (23) lauryl ether (Brij-35) was determined in the presence of various inorganic electrolytes. The measurements of cloud point (CP) were carried out with UV-vis spectrophotometer instead of visual observation. CP of Brij-35 could not be measured directly because its CP is more than 100 °C. Therefore, CP values of Brij-35 were lowered by the addition of electrolytes. In this study, NaF, NaCl, NaBr, KNO3, K2CO3, K3PO4, Li2SO4, Na2SO4, K2SO4, (NH4)2SO4, CuSO4, ZnSO4, CoSO4 were used as electrolytes. Linear lines which were drawn with CP values were extrapolated to zero electrolyte concentration. The real CP value of Brij-35 which is merely listed as >100 °C in the literature was found as 118.5 ± 0.5 °C for all samples. Furthermore, the effects of the nature of the cation and the anion and the valencies of the cations on CP were reported and the rate of decrease in CP with concentration is discussed.  相似文献   

9.
Affinity capillary electrophoresis (ACE) and quantum mechanical density functional theory (DFT) calculations have been employed for investigation of non-covalent interactions between macrocyclic ligand, benzo-18-crown-6-ether (B18C6) and ammonium cation, NH4+. Firstly, by means of ACE, the strength of the B18C6-NH4+ complex in mixed binary hydro-organic solvent system, methanol–water (50/50, v/v), was determined from the dependence of effective electrophoretic mobility of B18C6 (corrected to reference temperature 25 °C and constant ionic strength, 10 mM) on the concentration of ammonium ion in the background electrolyte (BGE) using non-linear regression analysis. The logarithmic form of the apparent binding (stability) constant (log Kb) of B18C6-NH4+ complex in the above binary solvent system was found to be equal to log Kb = 1.63 ± 0.10. Secondly, the structural characteristics of B18C6-NH4+ complex were described by quantum mechanical density functional theory (DFT) calculations. According to these calculations, in the energetically most favoured form of the B18C6-NH4+ complex, three strong hydrogen bonds are formed between central ammonium ion and B18C6 ligand, one of them is directed to aryl-O-alkyl (Ar–O–CH2) ethereal oxygen and the other two hydrogen bonds are oriented to alkyl-O-alkyl (CH2–O–CH2) ethereal oxygen atoms of the macrocyclic crown ligand.  相似文献   

10.
This work is to explore a new route to synthesize functional polyesters bearing pendant hydroxyl groups. The approach is via biocatalyzed direct polycondensation. l-Malic acid, adipic acid and 1,8-octanediol were used as comonomers and lipase Novozym 435 as a biocatalyst. 1H NMR studies on the structure of the products indicated that Novozym 435 was strictly selective for esterification of l-malic acid carboxyl groups while leaving the hydroxyl groups unchanged. The influences of the monomer feeding ratio, reaction temperature, and reaction time on the molecular weight of the products were investigated. By varying l-malic acid feed ratio in the total monomers from 0 to 20 mol%, the molecular weight (MW) of the product changed from 9.5 kilo Dalton (kD) to 4.7 kD while reaction was held at 70 °C for 48 h. The maximum MW could reach 7.4 kD at 80 °C when varying temperature between 70 and 90 °C if l-malic acid is 20 mol% and reaction time is 48 h. At 75 °C the MW increased from 5.2 kD to 6.6 kD when reaction time was elongated from 48 h to 72 h. However, little change in MW was observed at 80 and 85 °C when the reaction time was above 48 h. Thermal property of the copolyesters was studied by differential scanning calorimetry (DSC). Increasing the l-malic acid content in copolyesters resulted in melting temperature depression.  相似文献   

11.
Lukasz Kraszkiewicz 《Tetrahedron》2004,60(41):9113-9119
Two ‘model’ deactivated arenes, benzoic acid and nitrobenzene, were effectively monoiodinated within 1 h at 25-30 °C, with strongly electrophilic I+ reagents, prior prepared from diiodine and various oxidants (CrO3, KMnO4, active MnO2, HIO3, NaIO3, or NaIO4) in 90% (v/v) concd sulfuric acid (ca. 75 mol% H2SO4). Next, an I2/NaIO3/90% (v/v) concd H2SO4 exemplary system was used to effectively mono- or diiodinate a number of deactivated arenes. All former papers dealing with the direct iodination of deactivated arenes are briefly reviewed.  相似文献   

12.
Belaz KR  Cass QB  Oliveira RV 《Talanta》2008,76(1):146-153
The development and validation of a fully automated achiral-chiral high performance liquid chromatography (HPLC) method for the simultaneous determination of albendazole metabolites: enantiomers of albendazole sulphoxide (ABZ-SO), albendazole sulphone (ABZ-SO(2)) and albendazole 2-aminosulphone (ABZ-SO(2)NH(2)) in bovine plasma are described. This method involves an octyl restricted access media bovine serum albumin column (C(8)-RAM-BSA) (50 mm x 4.6 mm I.D.) for sample clean-up, followed by enantioselective analysis on a column containing an amylose tris(3,5-dimethylphenylcarbamate) stationary phase (150 mm x 4.6 mm I.D.). The chromatographic separations of all target compounds were performed at 30 degrees C using a mobile phase composed of phosphate buffer (10 mmol L(-1); pH 7.5):acetonitrile (60:40, v/v), flow rate of 0.5 mL min(-1) and fluorescence detection at 290 nm and 320 nm, excitation and emission, respectively. The influence of different organic modifiers and chiral selector of the stationary phase on enantioseparation of ABZ-SO was investigated. The method developed was fully validated. The calibration curves were linear in the concentration range of 40.00-1280 ng mL(-1) for each albendazole sulphoxide enantiomer, 10.0-320 ng mL(-1) for albendazole sulphone and 20.0-320 ng mL(-1) for albendazole 2-aminosulphone. The inter- and intra-day precision ranged from 0.760% to 7.79% relative standard deviation (R.S.D.), and the accuracy ranged 101% from 114% of the nominal values while the transfer efficiency was in the range of 84.4-103%. The method showed good linearity, precision, accuracy, sensitivity and selectivity allowing it to be appropriate for further pharmacokinetics and metabolism studies of albendazole.  相似文献   

13.
Titanate nanoribbons were prepared via a hydrothermal treatment of rutile-type TiO2 powders in a 10 M NaOH solution at 200 °C for 48 h. The as-prepared titanate nanoribbons were then hydrothermally post-treated at 150 °C for 12-36 h. The titanate nanoribbons before and after hydrothermal post-treatment were characterized with FESEM, XRD, TEM, UV-VIS and nitrogen adsorption-desorption isotherms. The results showed that the hydrothermal post-treatment not only promoted the phase transformation from titanate to anatase TiO2, but also was beneficial to the removal of Na+ ions remained in the titanate nanoribbons. After hydrothermal post-treatment, the TiO2 samples retained the one-dimensional structure feature of the titanate nanoribbons and showed an obvious increase in the specific surface area and the pore volume.  相似文献   

14.
We apply in-situ synchrotron X-ray diffraction to study the transformation of calcium monosulfoaluminate 14-hydrate Ca4Al2O6(SO4)·14H2O [monosulfate-14] to hydrogarnet Ca3Al2(OH)12 on the saturated water vapor pressure curve up to 250 °C. We use an aqueous slurry of synthetic ettringite Ca6Al2(SO4)3(OH)12·26H2O as the starting material; on heating, this decomposes at about 115 °C to form monosulfate-14 and bassanite CaSO4·0.5H2O. Above 170 °C monosulfate-14 diffraction peaks slowly diminish in intensity, perhaps as a result of loss of crystallinity and the formation of an X-ray amorphous meta-monosulfate. Hydrogarnet nucleates only at temperatures above 210 °C. Bassanite transforms to β-anhydrite (insoluble anhydrite) at about 230 °C and this transformation is accompanied by a second burst of hydrogarnet growth. The transformation pathway is more complex than previously thought. The mapping of the transformation pathway shows the value of rapid in-situ time-resolved synchrotron diffraction.  相似文献   

15.
Sr2SbMnO6 (SSM) powders were successfully synthesized at reasonably low temperatures via molten-salt synthesis (MSS) method using eutectic composition of 0.635 Li2SO4-0.365 Na2SO4 (flux). High-temperature cubic phase SSM was stabilized at room temperature by calcining the as-synthesized powders at 900 °C/10 h. The phase formation and morphology of these powders were characterized via X-ray powder diffraction and scanning electron microscopy, respectively. The SSM phase formation associated with ∼60 nm sized crystallites was also confirmed by transmission electron microscopy. The activation energy associated with the particle growth was found to be 95±5 kJ mol−1. The dielectric constant of the tetragonal phase of the ceramic (fabricated using this cubic phase powder) with and without the flux (sulphates) has been monitored as a function of frequency (100 Hz-1 MHz) at room temperature. Internal barrier layer capacitance (IBLC) model was invoked to rationalize the dielectric properties.  相似文献   

16.
In this study we present results of the conductivity and resistance to thermooxidative and condensation reactions of a highly phosphonated poly(pentafluorostyrene) (PWN2010) and of its blends with poly(benzimidazole)s (PBI). This polymer, which combines both: (i) a high degree of phosphonation (above 90%) and (ii) a relatively high acidity (pKa (–PO3H2 ↔ –PO3H) ∼ 0.5) due to the fluorine neighbors, is designed for low humidity operating fuel cell. This was confirmed by the conductivity measurements for PWN2010 reaching σ = 5 × 10−4 S cm−1 at 150 °C in dry N2 and σ = 1 × 10−3 S cm−1 at 150 °C (λ = 0.75). Furthermore, this polymer showed only 48% of anhydride formation when annealing it at T = 250 °C for 5 h and only 2% weight loss during a 96 h Fenton test. These properties combined with the ability of the PWN2010 to form homogeneous blends with polybenzimidazoles resulting in stable and flexible polymer films, makes PWN2010 a very promising candidate as a polymer electrolyte for intermediate- and high-temperature fuel cell applications.  相似文献   

17.
Hydroxy-amino-diphosphonates HO-Cn-NH2, with 2 ? n ? 11, have been successfully synthesized via the Kabachnick-Field reaction at 70 °C with high yields. These hydroxy compounds are then reacted with methacryloyl chloride to lead to novel amino-diphosphonate methacrylates MACnNP2 (with 2 ? n ? 11). These highly pure methacrylate monomers were obtained with yields higher than 75%. Radical copolymerizations of MACnNP2 (with 2 ? n ? 11) with MMA have been conducted and the r1 values (related to MACnNP2) are in the range of 1.1-1.3, and r2 values (related to MMA) about 0.8; this shows that the diphosphonate groups are statistically bonded to the methacrylic backbone.  相似文献   

18.
The reactivity of the dimeric cyclopalladated compounds derived from biphenyl-2-ylamine (μ-X)22-N2′,C1-1-Pd-2-{(2′-NH2C6H4)C6H4}]2 [X = OAc (1), X = Cl (2)] towards unsaturated organic molecules is reported. Compound 1 reacted with carbon monoxide and tbutyl isocyanide producing phenanthridin-6(5H)-one and N-tert-butylphenanthridin-6-amine in 63% and 88% yield, respectively. Compound 2 reacted separately with diphenylacetylene and 3-hexyne, affording the mononuclear organopalladium compounds [κ2-N2″,C12-C2,C3- 1-Pd{(R-CC-R)2-2′-(2″-NH2C6H4)C6H4}Cl] [R = Ph (5), R = Et (6)] in 50-60% yield, which derived from the insertion of two alkyne molecules into the C-Pd σ bonds of 2. The crystal structure of compounds 5 and 6 has been determined. Compound 5 crystallized in the monoclinic space group P21/n with a = 13.3290(10) Å, b = 10.6610(10) Å and c = 22.3930(10) Å and β = 100.2690(10)°. Compound 6 crystallized in the triclinic space group with a = 7.271(7) Å, b = 10.038(3) Å and c = 16.012(5) Å, and α = 106.79(3)°, β = 96.25(4)° and γ = 99.62(4)°. The crystal structures of 5 and 6 have short intermolecular Pd-Cl?H-N-Pd non-conventional hydrogen bonds, which associated the molecules in chains in the first case and in dimers in the second.  相似文献   

19.
Anatase-type TiO2 doped with 4.7 and 12.4 mol% ZrO2 that were directly precipitated as nanometer-sized particles from acidic precursor solutions of TiOSO4 and Zr(SO4)2 by simultaneous hydrolysis under hydrothermal conditions at 200°C, showed higher photocatalytic activity than pure anatase-type TiO2 for the decomposition of methylene blue. The crystallite growth and the phase transformation from anatase-type to rutile-type structure caused by heating at high temperature were retarded by doping ZrO2 into TiO2. The anatase-type TiO2 doped with ZrO2 showed high phase stability and maintained anatase-type structure even after heating at 1000°C for 1 h.  相似文献   

20.
The compound (NpO2)2(SO4)(H2O)4 was synthesized by evaporation of a Np5+ sulfate solution. The crystal structure was determined using single crystal X-ray diffraction and refined to an R1=0.0310. (NpO2)2(SO4)(H2O)4 crystallizes in triclinic space group P-1, a=8.1102(7) Å, b=8.7506(7) Å, c=16.234(1) Å, α=90.242(2)°, β=92.855(2)°, γ=113.067(2)°, V=1058.3(2) Å3, and Z=2. The structure contains neptunyl pentagonal bipyramids that share vertices through cation-cation interactions to form a sheet or cationic net. The sheet is decorated on each side by vertex sharing with sulfate tetrahedra, and adjacent sheets are linked together through hydrogen bonding. A graphical representation of (NpO2)2(SO4)(H2O)4 was constructed to facilitate the structural comparison to similar Np5+ compounds. The prevalence of the cationic nets in neptunyl sulfate compounds related to the overall stability of the structure is also discussed.  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