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.
Isothermal vapor–liquid equilibrium data at 333.15 K are measured for the binary system ethanol + 2,2,4-trimethylpentane and for ternary system di-methyl carbonate (DMC) + ethanol + 2,2,4-trimethylpentane by using headspace gas chromatography. The experimental binary and ternary vapor–liquid equilibrium data were correlated with different activity coefficient models. Excess volume and deviations in molar refractivity data are also reported for the binary systems DMC + ethanol and DMC + 2,2,4-trimethylpentane and the ternary system DMC + ethanol + 2,2,4-trimethylpentane at 298.15 K. These data were correlated with the Redlich-Kister equation for the binary systems and the Cibulka equation for the ternary system, respectively. The ternary excess volume and deviations in molar refractivity data were also compared with estimated values from the binary contribution models of Tsao–Smith, Kohler, Rastogi and Radojkovi?.  相似文献   

2.
A gas chromatography–mass spectrometric (GC–MS) method has been established for the determination of hydrazine in drinking water and surface water. This method is based on the derivatization of hydrazine with ortho-phthalaldehyde (OPA) in water. The following optimum reaction conditions were established: reagent dosage, 40 mg mL−1 of OPA; pH 2; reaction for 20 min at 70 °C. The organic derivative was extracted with methylene chloride and then measured by GC–MS. Under the established condition, the detection and the quantification limits were 0.002 μg L−1 and 0.007 μg L−1 by using 5.0-mL of surface water or drinking water, respectively. The calibration curve showed good linearity with r2 = 0.9991 (for working range of 0.05–100 μg L−1) and the accuracy was in a range of 95–106%, and the precision of the assay was less than 13% in water. Hydrazine was detected in a concentration range of 0.05–0.14 μg L−1 in 2 samples of 10 raw drinking water samples and in a concentration range of 0.09–0.55 μg L−1 in 4 samples of 10 treated drinking water samples.  相似文献   

3.
A novel method based on the strategy of N-phosphorylation labeling is described for quantification of twenty natural amino acids in human serum by reversed-phase liquid chromatography–electrospray tandem mass spectrometry (RP-LC/ESI-MS). The derivatization reaction was easily performed in one-pot reaction under mild conditions within 30 min. The reaction mixture was then evaporated to dryness, redissolved, desalted by C18 SPE. The twenty N-phosphoryl amino acids were separated on an RP-C18 column within 20 min by isocratic elution (0.1% formic acid–acetonitrile, v/v 7:3). At the same time, multiple reaction monitoring (MRM) MS enabled quantitation of twenty natural amino with the LOD of 0.0005–0.15 μM and LOQ of 0.0020–0.5 μM in human serum. The linear range was from 0.025 to 25 μM (except Cys and Trp) with R > 0.99. The recovery range was determined to be 85.5–117.4% with the relative standard deviation (RSD) in the range of 1.3–13.9%. All twenty amino acids were successfully detected in human serum samples with the concentration from 5.7 to 577.9 μM, which indicates potential of the developed method for determination of amino acids in complex biological samples, hence for screening of amino acid metabolite related diseases.  相似文献   

4.
The vapor–liquid equilibria for methanol + glycerol and ethanol + glycerol systems were measured by a flow method at 493–573 K. The pressure conditions focused in this work were 3.03–11.02 MPa for methanol + glycerol system and 2.27–8.78 MPa for ethanol + glycerol system. The mole fractions of alcohol in vapor phase are close to unity at the pressures below 7.0 MPa for both systems. The pressures of liquid saturated lines of the liquid phase for methanol + glycerol and ethanol + glycerol systems are higher than that for the mixtures containing alcohol and biodiesel compound, methyl laurate or ethyl laurate.  相似文献   

5.
Vapor–liquid equilibrium data for the binary systems of carbon monoxide (CO) + diethyl carbonate (DEC) and carbon monoxide + ethyl acetate (EA) were measured at temperatures of 293.2 K, 313.2 K and 333.2 K and the elevated pressures up to 12.00 MPa. The measurements were carried out in a cylindrical autoclave with a moveable piston and an observation window. The experimental data were correlated using the Peng–Robisom (PR) equation of state (EOS) and Peng–Robinson–Stryjek–Vera (PRSV) equation of state with the two-parameter van der Waals II or Panagiotopoulos–Reid mixing rule. The interaction parameters were obtained while correlating. The comparison between calculation results and experimental data indicated that the method of PRSV equation of state with van der Waals II produced the better correlated results.  相似文献   

6.
p-Nitrophenyl organophosphates (OPs) including paraoxon, parathion and methyl parathion, etc, are highly poisonous OPs, for which sensitive and rapid detection method is most needed. In this work, an ultrasensitive electrochemical sensor for the determination of p-nitrophenyl OPs was developed based on ordered mesoporous carbons (OMCs) modified glassy carbon electrode (GCE) (OMCs/GCE). The electrochemical behavior and reaction mechanism of p-nitrophenyl OPs at OMCs/GCE was elaborated by taking paraoxon as an example. Experimental conditions such as buffer pH, preconcentration potential and time were optimized. By using differential pulse voltammetry, the current response of the sensor at −0.085 V was linear with concentration within 0.01–1.00 μM and 1.00–20 μM paraoxon. Similar linear ranges of 0.015–0.5 μM and 0.5–10 μM were found for parathion, and 0.01–0.5 μM and 0.5–10 μM for methyl parathion. The low limits of detection were evaluated to be 1.9 nM for paraoxon, 3.4 nM for parathion and 2.1 nM for methyl parathion (S/N = 3). Common interfering species had no interference to the detection of p-nitrophenyl OPs. The sensor can be applicable to real samples measurement. Therefore, a simple, sensitive, reproducible and cost-effective electrochemical sensor was proposed for the fast direct determination of trace p-nitrophenyl OPs at low potential without deoxygenization.  相似文献   

7.
A flow-type method was adopted to measure the vapor–liquid equilibria for methanol + methyl laurate and methanol + methyl myristate systems at 493–543 K, near the critical temperature of methanol (Tc = 512.64 K), and 2.16–8.49 MPa. The effect of temperature and fatty acid methyl esters to the phase behavior was discussed. The mole fractions of methanol in liquid phase are almost the same for both systems. In vapor phase, the mole fractions of methanol are very close to unity at all temperatures. The present vapor–liquid equilibrium data were correlated by PRASOG. A binary parameter was introduced to the combining rule of size parameter. The binary parameters of methanol + fatty acid methyl ester systems were determined by fitting the present experimental data. The correlated results are in good agreement with the experimental data. The vapor–liquid equilibria for methanol + methyl laurate + glycerol and methanol + methyl myristate + glycerol ternary systems were also predicted using the methanol + fatty acid methyl ester binary parameters. The mole fractions of methanol in vapor phase are around unity even if glycerol is included in the systems.  相似文献   

8.
A new, fast and efficient multiple reaction monitoring (MRM) high-performance liquid chromatography–tandem mass spectrometry (HPLC–MS/MS) method for the determination of cyclopiazonic acid (CPA) in mixed feed, wheat, peanuts and rice is presented. The analytical methodology involves sample extraction with an alkaline methanol–water mixture, defatting with hexane and quantification using HPLC–MS/MS without further treatment of sample extracts. Reversed-phase liquid chromatography using a C18 stationary phase coupled to negative mode electrospray triple quadrupole tandem mass spectrometry was applied. The limit of detection was 5 μg/kg while the limit of quantification was 20 μg/kg in the matrices investigated. The detector response was found to be linear over the range 25–250 μg/kg in feed and 25–500 μg/kg in wheat, peanuts and rice. The mean overall recoveries (n = 18) of CPA varied from 79% to 114% in the range of concentrations studied over a period of 4 months. Mean recoveries (n = 3 or 6) of CPA in wheat, peanuts and rice varied from 70% to 111%, 77% to 116% and 69% to 92%, respectively. The method was successfully applied to the analysis of feed and rice samples artificially infected with the fungal strain Penicillium commune, where the toxin was found at different levels.  相似文献   

9.
The new polynuclear heterometal alkoxide clusters Ln2Na8(OCH2CF3)14(THF)6 (Ln = Sm 1, Y 2, Yb 3) have been synthesized by the reaction of anhydrous LnCl3 with 7 equiv. of NaOCH2CF3 in 68–75% yields. Crystal structural analysis revealed clusters 13 are isomorphous composed of two cubanes and a double open cubane, with one face of an Ln1Na2O4 open cubane capped by an additional Ln1O2 layer. Clusters 13 show extremely high activity for the polymerization of ε-caprolactone (ε-CL) and trimethylene carbonate (TMC). The reactivity is much higher than those found for the monometallic alkoxides lanthanide complexes previously reported. The dependence of catalytic activity on lanthanide metals is observed: Yb ≈ Y < Sm for ε-CL and Yb < Y < Sm for TMC. The polymers obtained with these clusters all show a unimodal molecular weight distribution with moderate molecular weight distributions (Mw/Mn = 1.4–1.7), indicating that clusters 13 can really be used as single-component catalysts. The bimetallic cooperation and the coordination–insertion mechanism were proposed.  相似文献   

10.
An ion-pairing reversed-phase liquid chromatography–mass spectrometry (IP-RP-LC–MS) was developed for the determination of nucleotides, nucleosides and their transformation products in Cordyceps. Perfluorinated carboxylic acid, namely pentadecafluorooctanoic acid (PDFOA, 0.25 mM), was used as volatile ion-paring agent and a reversed-phase column (Agilent ZORBAX SB-Aq column) was used for the separation of three nucleotides namely uridine-5′-monophosphate (UMP, 0.638–10.200 μg/mL), adenosine-5′-monophosphate (AMP, 0.24–7.80 μg/mL) and guanosine-5′-monophosphate (GMP, 0.42–13.50 μg/mL), seven nucleosides including adenosine (0.55–8.85 μg/mL), guanosine (0.42–6.75 μg/mL), uridine (0.33–10.50 μg/mL), inosine (0.21–6.60 μg/mL), cytidine (0.48–15.30 μg/mL), thymidine (0.20–6.30 μg/mL) and cordycepin (0.09–1.50 μg/mL), as well as six nucleobases, adenine (0.22–6.90 μg/mL), guanine (0.26–4.20 μg/mL), uracil (0.38–12.15 μg/mL), hypoxanthine (0.13–4.20 μg/mL), cytosine (0.39–12.45 μg/mL) and thymine (0.26–8.25 μg/mL) with 5-chlorocytosine arabinoside as the internal standard. The overall LODs and LOQs were between 0.01–0.16 μg/mL and 0.04–0.41 μg/mL for the 16 analytes, respectively. The contents of 16 investigated compounds in natural and cultured Cordyceps were also determined and compared after validation of the developed IP-RP-LC-MS method. The transformations of nucleotides and nucleosides in Cordyceps were evaluated based on the quantification of the investigated compounds in three extracts, including boiling water extraction (BWE), 24 h ambient temperature water immersion (ATWE) and 56 h ATWE extracts. Two transformation pathways including UMP → uridine → uracil and GMP → guanosine → guanine were proposed in both natural Cordyceps sinensis and cultured Cordyceps militaris. The pathway of AMP → adenosine → inosine → hypoxanthine was proposed in natural C. sinensis, while AMP → adenosine → adenine in cultured C. militaris. However, the transformation of nucleotides and nucleosides was not found in commercial cultured C. sinensis.  相似文献   

11.
The new method presented in this article achieved the goal of capturing Salmonella typhimurium via immunoreaction and rapid in situ detection of the CdSe/ZnS quantum dots (QDs) labeled S. typhimurium by self-assembly light-emitting diode-induced fluorescence detection (LIF) microsystem on a specially designed multichannel microfluidic chip. CdSe/ZnS QDs were used as fluorescent markers improving detection sensitivity. The microfluidic chip developed in this study was composed of 12 sample channels, 3 mixing zones, and 6 immune reaction zones, which also acted as fluorescence detection zones. QDs–IgG–primary antibody complexes were generated by mixing CdSe/ZnS QDs conjugated secondary antibody (QDs–IgG) and S. typhimurium antibody (primary antibody) in mixing zones. Then, the complexes went into immune reaction zones to label previously captured S. typhimurium in the sandwich mode. The capture rate of S. typhimurium in each detection zone was up to 70%. The enriched QDs-labeled S. typhimurium was detected using a self-assembly LIF microsystem. A good linear relationship was obtained in the range from 3.7 × 10 to 3.7 × 105 cfu mL−1 using the equation I = 0.1739 log (C) − 0.1889 with R2 = 0.9907, and the detection limit was down to 37 cfu mL−1. The proposed method of online immunolabeling with QDs for in situ fluorescence detection on the designed multichannel microfluidic chip had been successfully used to detect S. typhimurium in pork sample, and it has shown potential advantages in practice.  相似文献   

12.
Quantitation of trace levels of domoic acid (DA) in seawater samples usually requires labour-intensive protocols involving chemical derivatization with 9-fluorenylmethylchloroformate and liquid chromatography with fluorescence detection (FMOC–LC–FLD). Procedures based on LC–MS have been published, but time-consuming and costly solid-phase extraction pre-concentration steps are required to achieve suitable detection limits. This paper describes an alternative, simple and inexpensive LC method with ultraviolet detection (LC–UVD) for the routine analysis of trace levels of DA in seawater without the use of sample pre-concentration or derivatization steps. Qualitative confirmation of DA identity in dubious samples can be achieved by mass spectrometry (LC–MS) using the same chromatographic conditions. Addition of an ion-pairing/acidifying agent (0.15% trifluoroacetic acid) to sample extracts and the use of a gradient elution permitted the direct analysis of large sample volumes (100 μl), resulting in both high selectivity and sensitivity (limit of detection = 42 pg ml−1 by LC–UVD and 15 pg ml−1 by LC–MS). Same-day precision varied between 0.4 and 5%, depending on the detection method and DA concentration. Mean recoveries of spiked DA in seawater by LC–UVD were 98.8% at 0.1–10 ng ml−1 and 99.8% at 50–1000 ng ml−1. LC–UVD exhibited strong correlation with FMOC–LC–FLD during inter-laboratory analysis of Pseudo-nitzschia multiseries cultures containing 60–2000 ng DA ml−1 (r2 > 0.99), but more variable results were obtained by LC–MS (r2 = 0.85). This new technique was used to confirm the presence of trace DA levels in low-toxicity Pseudo-nitzschia spp. isolates (0.2–1.6 ng ml−1) and in whole-water field samples (0.3–5.8 ng ml−1), even in the absence of detectable Pseudo-nitzschia spp. cells in the water column.  相似文献   

13.
The isolation of the retinal isomers from all-trans-retinal was performed by flash countercurrent chromatography. In each separation, isomerization reaction solution of 200 mg all-trans-retinal could be loaded on a 1200 mL of high-speed countercurrent chromatographic column with 5 mm bore, eluted by a mobile phase flow rate of 25 mL/min, resulting in 63 mg of 11-cis-retinal, 24 mg of 13-cis-retinal and 26 mg of 9-cis-retinal with purities more than 95%. n-Hexane–acetonitrile (3:1) was used as the solvent system which possesses the advantages of simplicity, re-use of the solvent and multiple injections. This method could be used to prepare 13-cis-retinal, 11-cis-retinal and 9-cis-retinal for the photoisomerization investigation, such as the effect of 11-cis-retinal in the visual system.  相似文献   

14.
The mechanism of the NHC-catalyzed intermolecular Stetter reaction between benzaldehyde and cyclopropene has been investigated using the PCM-M062X/6-311++G(3df,2p)//M062X/6-31+G(d,p) level of DFT. Compared to the direct reaction, a substantial reduction in the activation free energy by 10.6–14.4 kcal/mol is observed when the reaction is performed in the presence of water, 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD), and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU). The bases promote the proton transfer step of the reaction to yield the Breslow intermediate. An early concerted transition state has been located for the stereocontrolling C–C bond formation step (ΔG# = 26.6 kcal/mol) which is used to explain the diastereomeric ratio observed in the experiment.  相似文献   

15.
Graphene oxide (GO) was functionalized with a N-heterocyclic carbene (NHC) precursor, 3-(3-aminopropyl)-1-methylimidazolium bromide ([APMIm][Br]) for the immobilization of palladium catalyst. The GO-supported NHC precursor (IMGO) formed a stable complex with Pd(OAc)2 (GO–NHC–Pd), which showed excellent catalytic activity and fast reaction kinetics in the aqueous-phase Suzuki reaction of aryl bromides and chlorides at relatively mild conditions (1 h at 50 °C). The GO–NHC–Pd catalyst was reused several times without any loss of its catalytic activity in the Suzuki reaction of aryl bromide.  相似文献   

16.
The synthesis of dihydrochromeno[4,3-b]pyrazolo[4,3-e]pyridin-6(7H)-ones was achieved through one-pot three-component reaction from 4-hydroxycoumarin, aldehydes, and 3-amino-5-methyl-pyrazole in acetonitrile using 5 mol % TBATB as the catalyst under reflux condition. The product formation is through tandem Knoevenagel–Michael reaction followed by concomitant cyclization. Simple reaction procedure, shorter reaction time, good yields, avoidance of aqueous work-up, and column chromatographic separation are some of the salient features of the present protocol.  相似文献   

17.
A new dinuclear compound, [Cu2(bpdo)2Br4], (in which bpdo = 2,2′-Bipyridine-1,1′-dioxide), has been synthesized and fully characterized, including the X-ray and the magnetic susceptibility. Each copper(II) ion in the dinuclear compound has a distorted square pyramidal geometry with the basal plane formed by two oxygen atoms of two ligand molecules which are bridging between the Cu ions with Cu–O distances of 2.021(2) and 2.039(2) Å and two bromide atoms with Cu–Br distances of 2.3577(6) and 2.3665(7) Å. The fifth position is occupied by a non bridging oxygen atom of a ligand with a Cu–O distance of 2.197(2) Å. The distance between the Cu ions is 3.334 Å, while the Cu–O–Cu angle is 110.37(9)°. The magnetic susceptibility measurements (from 5 to 350 K) agree with a very strong antiferromagnetic interaction with a large singlet–triplet splitting (J) of −905 cm−1. At high T (above 250 K) a triplet powder EPR is observed.  相似文献   

18.
Molecular dynamics simulations of monoclinic (P21/b) hydroxy- and chlorapatite were undertaken in the range 498 K < T < 1298 K, and for pressures up to 7.5 GPa. The all-atom Born–Huggins–Mayer force field, that had been previously used to successfully describe the room temperature isotherms of both compounds, was also used in this work. The isothermal sets of pV data generated by simulation were each fitted to the three-parameter form of the isothermal Parsafar and Mason equation of state (EoS) with an accuracy better than 0.07%. Taking the temperature dependence of the coefficients into account, it was found that the MD data are satisfactorily reproduced by the universal EoS. The isothermal compressibility coefficient dependence with pressure can be described by a linear relation.  相似文献   

19.
An all-glass, dynamic recirculating still equipped with an ultrasonic homogenizer has been used to determine vapor–liquid (VLE) and vapor–liquid–liquid (VLLE) equilibria. Consistent data have been obtained for the ternary water + ethanol + p-xylene system at 101.3 kPa for temperatures in the range of 351.16–365.40 K. Experimental results have been used to check the accuracy of the UNIFAC, UNIQUAC and NRTL models in the liquid–liquid region of importance in the dehydration of ethanol by azeotropic distillation.  相似文献   

20.
In this study, 26-membered macrocyclic carbonate, cyclobis(decamethylene carbonate) [(DMC)2] was attempted to undergo ring-opening polymerization by lipase catalysis in toluene. Novozym-435 exhibited even higher catalytic activity towards (DMC)2 polymerization compared with SnOCt2 while high molecular weight (Mn) of 5.4 × 104 and yield of 99% was still achieved at ultra-low enzyme/substrate (E/S) weight ratio of 1/200. 1H NMR spectra demonstrated the existence of terminal hydroxyl group. Solid phase polymerization in the absence of toluene unexpectedly took place at the temperature lower than (DMC)2’s melting point of 110 °C. Compared with solvent-free case, the addition of toluene solvent resulted in marked increase in reaction rate. As to the polymerization during 48 h with the E/S weight ratio of 1/100, a region existed at around toluene/carbonate (vol/wt, ml/g) ratio of 1∼2 where the polymerizations gave optimal results in terms of both higher molecular weight and monomer conversion. It was found that much higher molecular weight polymers may be obtained by decreasing enzyme concentrations. Plots of ln{[M]0:[M]t} versus reaction time were in linear agreement, indicating no chain termination, and monomer consumption follows a first-order rate law. The Novozym-435 catalyzed polymerization of (DMC)2 in toluene presented pseudo-living characteristic. Compared with 6-membered trimethylene carbonate, much lower reaction activity of large-sized (DMC)2 is observed, which is opposite to the result concerning the enzymatic polymerization of lactones with different ring-size.  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