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 non-isothermal decomposition of nickel acetate tetrahydrate in air was studied using thermogravimetry (TG)–DTG, differential scanning calorimetry (DSC) and XRD techniques. The decomposition occurs in two major steps and the final product is NiO. The dependence of mass loss on heating rates in TG measurements imply that the dehydration and hydrolysis concur at temperature below 240 °C; the apparent activation energies calculated by Flynn, Wall and Ozawa (FWO) isoconversional method indicate the existence of a consecutive process. The kinetics of the first major decomposition step (below 240 °C) was obtained with multivariate non-linear regression of four measurements at different heating rates. According to the kinetics results from non-linear regression, the dehydration reaction (F1 type with an activation energy E of 167.7 kJ/mol) goes first. After the loss of almost half of water, the retained water and acetate are linked to each other by hydrogen bonding, so dehydration and hydrolysis concur. The pathway with a lower E is related to the hydrolysis process and the other is corresponding to the dehydration process. The simulations of reactants at different heating rates were used to verify the correctness of the reaction model. With the kinetics results, the dehydration mechanism was discussed for the first time.  相似文献   

2.
A comparison of the thermal decomposition and combustion characteristics of raw and scoured cottons has demonstrated a mechanistic link caused by the presence of inorganic salts in raw cotton, which enhances resistance to heat and flame. Thermogravimetry, differential thermogravimetry, and microscale combustion calorimetry were used to examine the thermal decomposition kinetics and thermal stability of cotton. During pyrolysis, both raw cotton nonwoven and woven fabrics exhibited a slower decomposition with a larger initial weight loss and produced a greater char yield, as compared to the fabrics after scouring, which removes most inorganic components from cotton. The activation energy (E a ) values, calculated using the Kissinger method, the Flynn–Wall–Ozawa method, and the modified Coats–Redfern method, were consistently determined to be smaller for raw cotton than for scoured cotton. The analyses of cotton fabrics heated at elevated temperatures by 13C CP/MAS NMR and ATR-FTIR showed that trace quantities of inorganic components promoted the formations of oxygenated moieties at low temperatures and aliphatic intermediate char. In the combustion, raw cotton exhibited a much smaller heat release capacity and a smaller total heat release than scoured cotton, indicating enhanced thermal stability when the inorganic components are intact.  相似文献   

3.
Ammonium niobium oxalate was prepared and characterized by elemental analysis, XRD and FTIR spectroscopy analysis, which confirmed that the molecular formula of the complex is NH4(NbO(C2O4)2(H2O)2)(H2O)3. Dynamic TG analysis under air was used to investigate the thermal decomposition process of synthetic ammonium niobium oxalate. It shows that the thermal decomposition occurs in three stages and the corresponding apparent activation energies were calculated with the Ozawa–Flynn–Wall and the Friedman methods. The most probable kinetic models of the first two steps decomposition of the complex have been estimated by Coats–Redfern integral and the Achar–Bridly–Sharp differential methods.  相似文献   

4.
An intercalated nanocomposite of the organically modified montmorillonite Cloisite C25A and a degradable poly(ester amide) based on glycolic acid and 6-aminohexanoic acid units (poly(glc-alt-amh)) was prepared using a twin-screw co-rotating extruder. The non-isothermal degradation kinetics was investigated by thermogravimetric analysis (TG and DTG) in the temperature range of 50–600 °C at five heating rates (2, 5, 10, 20 and 40 °C/min) and compared with the neat polymer. Significant differences were found since the nanocomposite showed three degradation steps instead of the two decomposition processes detected in the pristine sample. The onset mass loss temperature decreased in the nanocomposite due to the presence of the organo-modifier compound, but the presence of the silicate layers significantly decreased the degradation rate at the last stages of decomposition. Kinetic analysis was performed using the Kissinger method and the isoconversional (Kissinger–Akahira–Sunose and Friedman) methods. The true kinetic triplets (E, A, f(α)) were determined for the first two steps of degradation through the Coats–Redfern and the Invariant Kinetic Parameters methods. The results clearly indicated that the presence of the organo-modified clay modified the mechanisms of degradation.  相似文献   

5.
Thermogravimetric (TG) data of oil sand obtained at Engineering Research Center of Oil Shale Comprehensive Utilization were studied to evaluate the kinetic parameters for Indonesian oil sand samples. Experiments were carried out at heating rates of 5, 15, and 25 °C min?1 in nitrogen, 10, 20, and 50 °C min?1 in oxygen atmosphere, respectively. The extent of char combustion was found out by relating TG data for pyrolysis and combustion with the ultimate analysis. Due to distinct behavior of oil shale during pyrolysis, TG curves were divided into three separate events: moisture release, devolatilization, and evolution of fixed carbon/char, where for each event, kinetic parameters, based on Arrhenius theory, were calculated. Coats–Redfern method, Flynn–Wall–Ozawa method, and distributed activation energy model method have been used to determine the activation energies of degradation. The methods are compared with regard to their characteristics and the ease of interpretation of the thermal kinetics. Activation energies of the samples were determined by three different methods and the results are discussed.  相似文献   

6.
Metal complexes of gliclazide (GLZ; HL) drug are prepared and characterized based on elemental analyses, IR, diffused reflectance, magnetic moment, molar conductance and thermal analyses (TG and DTG) technique. From the elemental analyses data, the complexes are proposed to have the general formulae [M(HL)Cl3(H2O)]·3H2O (M = Cr(III) and Fe(III)), [M(HL)Cl2(H2O)2yH2O (M = Co(III), Ni(II) and Cu(II), y = 0–2) and [M(HL)Cl2yH2O (M = Mn(II) and Zn(II), y = 0–1). The molar conductance data reveal that all the metal chelates are non-electrolytes. IR spectra show that GLZ is coordinated to the metal ions in a neutral bidentate manner with ON donor sites of the amide-O and sulphonamide-OH. From the magnetic and solid reflectance spectra, it is found that the geometrical structures of these complexes are octahedral (Cr(III), Fe(III), Co(II), Ni(II) and Cu(II)) and tetrahedral (Mn(II) and Zn(II)). The thermal behaviour of these chelates is studied using thermogravimetric analysis (TG and DTG) techniques. The results obtained show that the hydrated complexes lose water molecules of hydration followed immediately by decomposition of the anions and ligand molecules in the successive unseparate steps. The activation thermodynamic parameters are calculated using Coats–Redfern method. The GLZ drug, in comparison to its metal complexes also is screened for their biological activity against house fly, Musca domestica (Diptera – Muscidae). Dose of 5 μg/insect of gliclazide is typically applied against 3 days-old larval instar of M. domestica. Survival of pupal and adult stages has been affected by the complexes of gliclazide more than larval instars. Morphogenic abnormalities of larvae, pupae and adults are studied. On the other hand, pupation and adult emergence program is deteriorated by the effect of different chemicals.  相似文献   

7.
Polytriazole polyethylene oxide‐tetrahydrofuran (PTPET) is an energetic propellant elastomer that is prepared using glycidyl azide polymer and trifunctional alkynyl‐terminated polyethylene oxide‐tetrahydrofuran. Its thermal decomposition, determined using thermogravimetic analysis, showed two mass‐loss peaks largely related to the decomposition of azide groups and the main chain. Flynn‐Wall‐Ozawa and Kissinger‐Akahira‐Sunose methods were deployed to obtain kinetic triplet parameters of PTPET thermal decomposition by the traditional model‐free method; the Coats‐Redfern approach was used as the model‐fitting method. Kinetics analysis indicated that the mechanism of the two‐step reactions were the primary‐reaction of first order and the power‐law phase reaction of the 2/3 order. The first decomposition stage of PTPET had an activation energy (Ea) of 113 to 116 kJ/mol while the second was 196 to 210 kJ/mol. The thermal decomposition of PTPET with different heating rates and mechanisms showed good kinetic compensation effects, the gas products being further studied with TG‐FTIR.  相似文献   

8.
Thermal decomposition of compounds consisting of tetrahalogenomanganese(II) anions, [MnBrnCl4?n]2? (n = 0–4), and a tetrabutylammonium cation has been studied using the DSC, TG-FTIR, TG–MS and DTA techniques. The measurements were carried out in an argon and static air atmospheres over the temperature ranges 173–450 K (DSC) and 300–1073 K (TG). Solid products of the thermal decomposition were identified by FT-FIR spectroscopy as well as X-ray powder diffractometry.  相似文献   

9.
Struvite-K (KMgPO4·6H2O) is the main hydration product of magnesium potassium phosphate cement. Its thermal stability is critical to the properties of magnesium potassium phosphate cement. Therefore, in this study, the dehydration behavior of struvite-K was investigated at N2 atmosphere in non-isothermal condition. The process was conducted and controlled in a simultaneous TG/DTA analyzer, at heating rates of 2, 5, 10, 15, and 20 K min?1. The residual mass was always around 58.5% of the initial one, regardless of the heating rate, which corresponds to the dehydration reaction through one step, KMgPO4·6H2O → KMgPO4. The activation energy (E a) corresponding to the dehydration of struvite-K was evaluated by non-isothermal kinetic analysis based on the application of isoconversional methods (Flynn–Wall–Ozawa and Kissinger–Akahira–Sunose methods). The calculated results show that Flynn–Wall–Ozawa has slightly higher values of activation energy (E a) and correlation coefficients (R 2). Both methods have been proved to be suitable for analyzing dehydration behavior of struvite-K.  相似文献   

10.
The surface acid–base property of carboxylic multi-walled carbon nanotubes (MWNTs) is investigated by zero current potentiometry with a new electrochemical measurement system. The pH dependent interface potential variation at the interface of carboxylic MWNTs/solution is investigated by measuring zero current potential Ezcp. In the pH range of 1–11, the pH response of carboxylic MWNTs exhibits two linear relationships according to the following equations: Ezcp = 0.791–0.0535 pH (pH 1–5.1) and Ezcp = 0.643–0.0241 pH (pH 5.1–11), respectively. The intersection at pH 5.1 of two regions indicates the surface pKa value of carboxylic group terminated MWNTs.  相似文献   

11.
The vapour pressures of liquid (3-diethylaminopropylamine (3-DEPA) + n-heptane) mixtures were measured by a static method between T = (303.15 and 343.15) K at 10 K intervals. The molar excess enthalpies HE at T = 303.15 K were measured for the systems {3-DEPA + CnH2n+2 (n = 6, 7, 12)}. The molar excess Gibbs free energies GE were obtained with Barker’s method and fitted to the Redlich–Kister equation. The Wilson equation was also used. Deviations between experimental and predicted GE and HE, by using group contribution UNIFAC (Gmehling version) model, were evaluated.  相似文献   

12.
Binary (vapor + liquid) equilibrium data were measured for the {carbon dioxide + isopropoxyethanol (iC3E1)} and the {carbon dioxide + isobutoxyethanol (iC4E1)} systems at temperatures ranging from (313.15 to 333.15) K. These experiments were performed with a circulating-type apparatus with on-line gas chromatography. The experimental data correlated well with the Peng–Robinson equation of state using the Wong–Sandler mixing rules.  相似文献   

13.
《Solid State Sciences》2007,9(7):574-579
This paper reports the specific conditions used in the preparation of the dimorphic phases of HoCrO4 oxide. The scheelite form has been obtained from the room pressure stable HoCrO4–zircon heated at 823 K at 40 kbar. The structures of both the polytypes have been refined from X-ray powder diffraction data using the Rietveld method. The zircon type of HoCrO4 oxide crystallizes with tetragonal symmetry, S.G. I41/amd and lattice parameters a = 7.119(10) and c = 6.2557(5) Å; while the scheelite–HoCrO4 derivative shows tetragonal symmetry, S.G. I41/a and lattice parameters a = 5.0017(1) and c = 11.2664(2) Å. Magnetic susceptibility measurements show that both zircon and scheelite forms of HoCrO4 oxide present a very different magnetic behaviour at low temperatures. The zircon form behaves as a ferromagnet with a Curie temperature of 17.6 K, while the scheelite form is antiferromagnetic with a Néel temperature of 7.6 K. A metamagnetic transition for the scheelite form with a critical field of 1 T at 2 K has been also observed. The changes in the sign of the interactions have been also analyzed by considering the possible super-exchange mechanisms and the differences found in the Ho–O–Cr distances and bond angles in going from zircon to scheelite structural type.  相似文献   

14.
According to TG/DTA analysis, the new pyrochlore-type compound K0.88H1.12Nb2O6·1.58H2O (Fd-3 m, a = 10.645(4) Å at 300 K) loses crystallization water on heating up to 773 K, but retains –OH hydrogen. Impedance spectroscopy in non-humidified air suggests the material is promising for SOFC applications at intermediate temperatures because the bulk conductivity values reach 10−2 S/cm at 623 K.  相似文献   

15.
《Solid State Sciences》2007,9(6):491-495
A novel zinc(II) phthalate compound, [Zn{C6H4(COO)2}2]n2−[NH3–(CH2)3–NH3]n2+, 1 which contains four different phthalate moieties coordinated to the Zn(II) ion through one each of their carboxylate moieties in a η1-form and each phthalate moiety acting as a bridging unit with an overall tetrahedral geometry around the metal ion has been prepared and structurally characterized. The structure is unique in the sense that the dianionic moieties [Zn{C6H4(COO)2}2]2− form an infinite one-dimensional network composed of 14-membered cyclic units interconnected in a way that they are disposed alternatively in almost perpendicular planes. The dicationic [NH3–(CH2)3–NH3]2+ units are seen to hold these one-dimensional chains by strong coulombic and hydrogen bond interaction, resulting in an infinite two-dimensional layered structure of 1. The compound is thermally stable up to 250 °C. Above this temperature it loses one equivalent of phthalate moiety along with the diammonium unit to yield Zn(II) phthalate. The thermodynamic and kinetic parameters associated with this process could be evaluated using Coats–Redfern equation which shows the activation energy Ea for the process as 305.0 kJ mol−1, the frequency factor A as 1.49 × 1011 s−1 and the entropy change ΔS as −35.90 JK−1 mol−1. Fluorescent emission properties of 1 was studied by exciting the compound at 380 nm and also at 322 nm which were found to be the two λmax corresponding to absorptions of the molecule. Interestingly, the compound gave same type of emission spectra showing a maximum around 444 nm on exciting at these two different wavelengths, indicating that the molecule comes to the first excited state from the higher excited state by a fast non-radiative process before it exhibits singlet emission to come to the ground state.  相似文献   

16.
This work points out that electrogeneration of silica gel (SG) films on glassy carbon electrodes (GCEs) can be applied to immobilize biomolecules – hemoglobin (Hb) or glucose oxidase (GOD) or both of them in mixture – without preventing their activity. These proteins were physically entrapped in the sol–gel material in the course of the electro-assisted deposition process applied to form the thin films onto the electrode surface. SG films were prepared from a precursor solution by applying a suitable cathodic potential likely to induce a local pH increase at the electrode/solution interface, accelerating thereby polycondensation of the silica precursors with concomitant film formation. Successful immobilization of proteins was checked by various physico-chemical techniques. Both Hb and GOD were found to undergo direct electron transfer, as demonstrated by cyclic voltammetry. GCE–SG–Hb gave rise to well-defined peaks at potentials Ec = −0.29 V and Ea = −0.17 V in acetate buffer, corresponding to the FeIII/FeII redox system of heme group of the protein, while GCE–SG–GOD was characterized by the typical signals of FAD group at Ec = −0.41 V and Ea = −0.33 V in phosphate buffer. These two redox processes were also evidenced on a single voltammogram when both Hb and GOD were present together in the same SG film. Hb entrapped in the silica thin film displayed an electrocatalytic behavior towards O2 and H2O2 in solution, respectively in the mM and μM concentration ranges. Immobilized GOD kept its biocatalytic properties towards glucose. Combined use of these two proteins in mixture has proven to be promising for detection of glucose in solution via the electrochemical monitoring of oxygen consumption (decrease of the oxygen electrocatalytic signal).  相似文献   

17.
Solubilities of six pharmaceuticals, namely nadolol, atenolol, bifonazole, nimesulide, estrone, mefenamic acid at constant pH, were measured over the range of temperature from (240 to 340) K in three important for drug solvents: water, ethanol, and 1-octanol using the dynamic method and spectroscopic UV–Vis method. Dissociation constants and corresponding pKa values of the drugs were obtained with the Bates–Schwarzenbach method using UV–Vis Perkin–Elmer Lambda 35 Spectrophotometer at temperature 298.15 K in the buffer solutions. Our experimental pKa values for nadolol, bifonazole, nimesulide, and mefenamic acid are 9.3, 5.85, 7.34, and 3.88, respectively. The basic thermal properties of pure drugs i.e. fusion and glass-transition temperatures, as well as the enthalpy of fusion and the molar heat capacity at the glass-transition (at constant pressure) have been measured using the differential scanning microcalorimetry technique (DSC). Molar volumes have been calculated with the Barton group contribution method. The experimental solubility results have been correlated by means of three commonly known GE equations: the Wilson, NRTL, and UNIQUAC with the assumption that the systems studied here are simple eutectic mixtures. The activity coefficients of pharmaceuticals in saturated solutions in each correlated binary mixture were calculated from the experimental results. Prediction of solubility in water at T = 298.15 K was made by the group contribution method.  相似文献   

18.
《Fluid Phase Equilibria》2002,200(1):41-51
Vapor–liquid equilibrium (VLE) data are reported for the binary mixtures formed by octane and the branched ether 1,1-dimethylpropyl methyl ether (tert-amyl methyl ether or TAME). A Gibbs–van Ness type apparatus was used to obtain total vapor pressure measurements as a function of composition at 298.15, 308.15, 318.15 and 328.15 K. The system shows positive deviations from Raoult’s law. These VLE data are analyzed together with data previously reported for octane+TAME mixtures: VLE data at 323.15 and 423.15 K, excess enthalpy (HmE) data at 298.15 and 313.15 K and excess volume (VmE) data at 298.15 K. The UNIQUAC model, the lattice–fluid (LF) model, and the Flory theory are used to simultaneously correlate VLE and HmE data. The two latter models are then used to predict VmE data. The original UNIFAC group contribution model and the modified UNIFAC (Dortmund model) are used to predict VLE data.  相似文献   

19.
The relationship of the nucleophilicity of alkylamines to their basicity is explored with emphasis on steric hindrance to solvation. The equation n = 1.43(?Σσ1 + δEs) + 6.35, where n is the Swain-Scott nucleophilic value, σ1 is the Taft polarity value, and δEs is the Taft steric value, correlates the nucleophilic constants of 17 common alkylamines and ammonia over two powers of 10 with a correlation coefficient of 0.98. The equation n ? 1.42 δEs = 0.44(pKa + S) + 0.17, where S is the solvation constant, correlates the nucleophilicities of these amines and ammonia with their pKa values over 3 powers of 10 with a correlation coefficient of 0.99. Excessive steric hindrance and nearby functional groups cause deviations from these equations.  相似文献   

20.
The densities of (heptane + 1-chlorobutane, or 1-chloropentane, or 1-chlorohexane) were measured at the temperatures (308.15, 318.15, and 328.15) K by means of a vibrating-tube densimeter. The excess molar volumes, VmE, calculated from the density data, along with our previous data(1) determined at T=298.15 K for the same systems, provide the temperature dependence of VmE in the temperature range of 298 to 328 K. The VmE results were correlated using the fourth-order Redlich–Kister equation, with the maximum likelihood principle applied for the determination of the adjustable parameters. It was found that the deviations from ideal behaviour (both positive and negative) in the systems studied increase with increasing temperature.  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