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.
Leachate samples from a sanitary landfill of Araraquara city and composting usine of Vila Leopoldina, São Paulo, Brazil were lyophilized to remove the water content. TG/DTG curves at different heating rates were recorded. The second step of the thermal decomposition of leachate from the Araraquara landfill (CB1), from the composting usine from Vila Leopoldina (CB2) from the organic phase extracted (FO) and aqueous phase (FA) were all kinetically evaluated using the non-isothermal method.By Flynn-Wall isoconversional method the following values were obtained: E=234±3.65 kJ mol?1 and logA=29.7±0.58 min?1 for CB1; E=129±1.66 kJ mol?1 and logA=11.8±0.10 min?1 for CB2; E=51.6±1.35 kJ mol?1 and logA=6.09±0.09 min?1 for FO and E=76.91±6.33 kJ mol?1 and logA=8.88±0.7 min?1 for FA with 95% confidence level. Applying the procedures of Málek and Koga, SB kinetic model (?esták-Berggren) is the most appropriate to describe the decomposition of CB1, CB2, FO and FA.  相似文献   

2.
A study of the strong N?X????O?N+ (X=I, Br) halogen bonding interactions reports 2×27 donor×acceptor complexes of N‐halosaccharins and pyridine N‐oxides (PyNO). DFT calculations were used to investigate the X???O halogen bond (XB) interaction energies in 54 complexes. A simplified computationally fast electrostatic model was developed for predicting the X???O XBs. The XB interaction energies vary from ?47.5 to ?120.3 kJ mol?1; the strongest N?I????O?N+ XBs approaching those of 3‐center‐4‐electron [N?I?N]+ halogen‐bonded systems (ca. 160 kJ mol?1). 1H NMR association constants (KXB) determined in CDCl3 and [D6]acetone vary from 2.0×100 to >108 m ?1 and correlate well with the calculated donor×acceptor complexation enthalpies found between ?38.4 and ?77.5 kJ mol?1. In X‐ray crystal structures, the N‐iodosaccharin‐PyNO complexes manifest short interaction ratios (RXB) between 0.65–0.67 for the N?I????O?N+ halogen bond.  相似文献   

3.
The rate of the fastest ene reaction between 4-phenyl-1,2,4-triazoline-3,5-dione (1) and 2,3-dimethyl-2-butene (2) is studied by means of stopped flow in solutions of benzene (k 2 = 55.6 ± 0.5 and 90.5 ± 1.3 L mol?1 s?1 at 23.3 and 40°C) and 1,2-dichloroethane (335 ± 9 L mol?1 s?1 at 23.5°C). The enthalpy of reaction (?139.2 ± 0.6 kJ/mol in toluene and ?150.2 ± 1.4 kJ/mol in 1,2-dichloroethane) and the enthalpy (20.0 ± 0.5 kJ/mol) and entropy (144 ± 2 J mol?1 K?1) of activation are determined. A clear correlation is observed between the reaction rate and ionization potential in a series of ene reactions of 4-phenyl-1,2,4-tri-azoline-3,5-dione with acyclic alkenes.  相似文献   

4.
An extracellular, endo-??-1,4-xylanase was purified to homogeneity from the culture filtrate of the filamentous fungus Penicillium occitanis Pol6, grown on oat spelt xylan. The purified enzyme (PoXyn2) showed a single band on SDS?CPAGE with an apparent molecular weight of 30?kDa. The xylanase activity was optimal at pH?3.0 and 65?°C. The specific activity measured for oat spelt xylan was 2,368?U?mg?1. The apparent K m and V max values were 8.33?mg?ml?1 and 58.82???mol?min?1?ml?1, respectively, as measured on oat spelt xylan. Thin-layer chromatography experiments revealed that purified PoXyn2 degrades xylan in an endo-fashion releasing xylobiose as main end product. The genomic DNA and cDNA encoding this protein were cloned and sequenced. This PoXyn2 presents an open reading frame of 962?bp, not interrupted by any introns and encoding for a mature protein of 320 amino acids and 29.88?kDa.  相似文献   

5.
The formation of weakly bound molecular complexes between dimethyl ether (DME) and the trifluoromethyl halides CF3Cl, CF3Br and CF3I dissolved in liquid argon and in liquid krypton is investigated, using Raman and FTIR spectroscopy. For all halides evidence is found for the formation of C? X???O halogen‐bonded 1:1 complexes. At higher concentrations of CF3Br, a weak absorption due to a 1:2 complex is also observed. Using spectra recorded at temperatures between 87 and 125 K, the complexation enthalpies for the complexes are determined to be ?6.8(3) kJ mol?1 (DME?CF3Cl), ?10.2(1) kJ mol?1 (DME?CF3Br), ?15.5(1) kJ mol?1 (DME?CF3I), and ?17.8(5) kJ mol?1 [DME(?CF3Br)2]. Structural and spectral information on the complexes is obtained from ab initio calculations at the MP2/ 6‐311++G(d,p) and MP2/6‐311++G(d,p)+LanL2DZ* levels. By applying Monte Carlo free energy perturbation calculations to account for the solvent influences, and statistical thermodynamics to estimate the zero‐point vibrational and thermal influences, the ab initio complexation energies are converted into complexation enthalpies for the solutions in liquid argon. The resulting values are compared with the experimental data deduced from the cryosolutions.  相似文献   

6.
The vulcanization of rubber by sulfur is a large‐scale industrial process that is only poorly understood, especially the role of zinc oxide, which is added as an activator. We used the highly symmetrical cluster Zn4O4 (Td) as a model species to study the thermodynamics of the initial interaction of various vulcanization‐related molecules with ZnO by DFT methods, mostly at the B3LYP/6‐31+G* level. The interaction energy of Lewis bases with Zn4O4 increases in the following order: CO62H43H62S2<1,4‐C5H82O2S3N?CH3COO?. The corresponding binding energies range from ?57 to ?262 kJ mol?1. However, Brønsted acids react with the Zn4O4 cluster with proton transfer from the ligand molecule to one of the oxygen atoms of Zn4O4, and these reactions are all strongly exothermic [binding energies [kJ mol?1] in parentheses: H2O (?183), MeOH (?171), H2S (?245), MeSH (?230), C3H6 (?121), and CH3COOH (?255)]. The important vulcanization accelerator mercaptobenzothiazole (C7H5NS2, MBT) containing several donor sites reacts with the Zn4O4 cluster with proton transfer from the NH group to one of the oxygen atoms of ZnO, and in addition the exocyclic thiono sulfur atom and the nitrogen atom coordinate to one and the same zinc atom, resulting in a binding energy of ?247 kJ mol?1. A second isomer of [(MBT)Zn4O4] with a strong O? H???N hydrogen bond rather than a Zn? N bond is only slightly less stable (binding energy ?243 kJ mol?1). The NH form of free MBT is 36 kJ mol?1 more stable than the tautomeric SH form, while the sulfurized MBT derivative benzothiazolyl hydrodisulfide C7H5NS3 (BtSSH) is most stable with the connectivity >CSSH.  相似文献   

7.
A newly isolated bacterial strain, Bacillus sp. MX47, was actively producing extracellular xylanase only in xylan-containing medium. The xylanase was purified from the culture broth by two chromatographic steps. The xylanase had an apparent molecular weight of 26.4?kDa with an NH2-terminal sequence (Gln-Gly-Gly-Asn-Phe) distinct from that of reported proteins, implying it is a novel enzyme. The optimum pH and temperature for xylanase activity were 8.0 and 40?°C, respectively. The enzyme activity was severely inhibited by many divalent metal ions and EDTA at 5?mM. The xylanase was highly specific to beechwood and oat spelt xylan, however, not active on carboxymethyl cellulose (CMC), avicel, pectin, and starch. Analysis of the xylan hydrolysis products by Bacillus sp. MX47 xylanase indicated that it is an endo-??-1,4-xylanase. It hydrolyzed xylan to xylobiose as the end product. The K m and V max values toward beechwood xylan were 3.24?mg?ml?1 and 58.21???mol?min?1?mg?1 protein, respectively.  相似文献   

8.
Vent sizing package 2 (VSP2) was used to measure the thermal hazard and runaway characteristics of 18650 lithium-ion batteries, which were manufactured by Sanyo Electric Co., Ltd. Runaway reaction behaviors of these batteries were obtained: 50% state of charge (SOC), and 100% SOC. The tests evaluated the thermal hazard characteristics, such as initial exothermic temperature (T 0), self-heating rate (dT?dt ?1), pressure-rise rate (dP?dt ?1), pressure temperature profiles, maximum temperature, and pressure which were observed by adiabatic calorimetric methodology via VSP2 using customized test cells. The safety assessment of lithium-ion cells proved to be an important subject. The maximum self-heating rate (dT?dt ?1)max and the largest pressure-rise rate (dP?dt ?1)max of Sanyo 18650 lithium-ion battery of 100% SOC were measured to be 37,468.8???C?min?1 and 10,845.6?psi?min?1, respectively, and the maximum temperature was 733.1???C. Therefore, a runaway reaction is extremely serious when a lithium-ion battery is exothermic at 100% SOC. This result also demonstrated that the thermal VSP2 is an alternative method of thermal hazard assessment for battery safety research. Finally, self-reactive ratings on thermal hazards of 18650 lithium-ion batteries were studied and elucidated to a deeper extent.  相似文献   

9.
This paper reports the computed O2 binding to heme, which for the first time explains experimental enthalpies for this process of central importance to bioinorganic chemistry. All four spin states along the relaxed Fe? O2‐binding curves were optimized using the full heme system with dispersion, thermodynamic, and scalar‐relativistic corrections, applying several density functionals. When including all these physical terms, the experimental enthalpy of O2 binding (?59 kJ mol?1) is closely reproduced by TPSSh‐D3 (?66 kJ mol?1). Dispersion changes the potential energy surfaces and leads to the correct electronic singlet and heptet states for bound and dissociated O2. The experimental activation enthalpy of dissociation (~82 kJ mol?1) was also accurately computed (~75 kJ mol?1) with an actual barrier height of ~60 kJ mol?1 plus a vibrational component of ~10 and ~5 kJ mol?1 due to the spin‐forbidden nature of the process, explaining the experimentally observed difference of ~20 kJ mol?1 in enthalpies of binding and activation. Most importantly, the work shows how the nearly degenerate singlet and triplet states increase crossover probability up to ~0.5 and accelerate binding by ~100 times, explaining why the spin‐forbidden binding of O2 to heme, so fundamental to higher life forms, is fast and reversible.  相似文献   

10.
Insight into the key factors driving the competition of halogen and hydrogen bonds is obtained by studying the affinity of the Lewis bases trimethylamine (TMA), dimethyl ether (DME), and methyl fluoride (MF) towards difluoroiodomethane (CHF2I). Analysis of the infrared and Raman spectra of solutions in liquid krypton containing mixtures of TMA and CHF2I and of DME and CHF2I reveals that for these Lewis bases hydrogen and halogen‐bonded complexes appear simultaneously. In contrast, only a hydrogen‐bonded complex is formed for the mixtures of CHF2I and MF. The complexation enthalpies for the C?H ??? Y hydrogen‐bonded complexes with TMA, DME, and MF are determined to be ?14.7(2), ?10.5(5) and ?5.1(6) kJ mol?1, respectively. The values for the C?I ??? Y halogen‐bonded isomers are ?19.0(3) kJ mol?1 for TMA and ?9.9(8) kJ mol?1 for DME. Generalization of the observed trends suggests that, at least for the bases studied here, softer Lewis bases such as TMA favor halogen bonding, whereas harder bases such as MF show a substantial preference for hydrogen bonding.  相似文献   

11.
In order to understand the mobility of uranium it is very important to know about its sorption kinetics and the thermodynamics behind the sorption process on soil. In the present study the sorption kinetics of uranium was studied in soil and the influence parameters to the sorption process, such as initial uranium concentration, pH, contact time and temperature were investigated. Distribution coefficient of uranium on soil was measured by laboratory batch method. Experimental isotherms evaluated from the distribution coefficients were fit to Langmuir, Freundlich and Dubinin?CRadushkevich (D?CR) models. The sorption energy for uranium from the D?CR adsorption isotherm was calculated to be 7.07?kJ?mol?1.The values of ??H and ??S were calculated to be 37.33?kJ?mol?1 and 162?J?K?1?mol?1, respectively. ??G at 30?°C was estimated to be ?11.76?kJ?mol?1. From sorption kinetics of uranium the reaction rate was calculated to be 1.6?×?10?3?min?1.  相似文献   

12.
The heat of reaction for SnJ2 (c)+J2 (c)+4045 CS2 (l)=[SnJ4; 4045 CS2] (sol) has been determined to be (?41.12±0.55) kJ mol?1, [(?9.83±0.13) kcal mol?1] by isoperibol solution calorimetry. Combining this result with the heat of formation of SnJ4 in CS2 determined in a previous investigation11 the value (?153.9±1.40) kJ mol?1, [(?36.9±0.33) kcal mol?1] has been derived for the heat of formation, ΔH f ι (SnJ2;c; 298.15 K), of tin diiodide.  相似文献   

13.
The reversible dimerisation of o-phenylenedioxydimethylsilane (2,2-dimethyl-1,3,2-benzodioxasilole) has been studied by 1H NMR spectroscopy. The kinetics of this reaction can be described quantitatively by a bimolecular 10-ring formulation reaction and a monomolecular backreaction. The thermodynamic and kinetic parameters are: ΔH0 = ?43 kJ mol?1; ΔS0 = ?112 J mol?1 K?1; ΔG0298 = ?9.6 kJ mol?1; ΔH3298 = 57 kJ mol?1; ΔS3298 = ?129 J mol?1 K?1; ΔG3298 = 96 kJ mol?1; Ea = 60 kJ mol?1; A = 3.17 × 106 l mol?1 s?1. Remarkable is the low activation energy of formation of the ten-membered ring, considering that two SiO bonds have to be cleaved during the reaction. Transition states and possible structures of the ten-membered heterocycle are discussed.  相似文献   

14.
Alkali niobates and tantalates are currently important lead‐free functional oxides. The formation and decomposition energetics of potassium tantalum oxide compounds (K2O?Ta2O5) were measured by high‐temperature oxide melt solution calorimetry. The enthalpies of formation from oxides of KTaO3 perovskite and defect pyrochlores with K/Ta ratio of less than 1 stoichiometry—K0.873Ta2.226O6, K1.128Ta2.175O6, and K1.291Ta2.142O6—were experimentally determined, and the values are (?203.63±2.92) kJ mol?1 for KTaO3 perovskite, and (?339.54±5.03) kJ mol?1, (?369.71±4.84) kJ mol?1, and (?364.78±4.24) kJ mol?1, respectively, for non‐stoichiometric pyrochlores. That of stoichiometric defect K2Ta2O6 pyrochlore, by extrapolation, is (?409.87±6.89) kJ mol?1. Thus, the enthalpy of the stoichiometric pyrochlore and perovskite at K/Ta=1 stoichiometry are equal in energy within experimental error. By providing data on the thermodynamic stability of each phase, this work supplies knowledge on the phase‐formation process and phase stability within the K2O?Ta2O5 system, thus assisting in the synthesis of materials with reproducible properties based on controlled processing. Additionally, the relation of stoichiometric and non‐stoichiometric pyrochlore with perovskite structure in potassium tantalum oxide system is discussed.  相似文献   

15.
The kinetics of formation of the 1?:?1 complex of chromium(III) with 1,3-propanediamine-N,N′-diacetate-N,N′-di-3-propionate (1,3-pddadp) were followed spectrophotometrically at λ max?=?557?nm. The reaction was first-order in chromium(III). Increasing the 1,3-pddadp concentration from 2.2?×?10?2 to 0.11?mol?dm?3 accelerated the reaction rate. Increasing the hydrogen ion concentration from 1.995?×?10?5 to 6.31?×?10?4 mol?dm?3 retarded the reaction rate. The reaction rate was also retarded by increasing ionic strength and dielectric constant of the reaction medium. A mechanism was suggested to account for the results obtained which involves ion-pair formation between the various reactants. Values of 22?kJ?mol?1 and ?115?J?K?1 mol?1 were obtained for the energy and the entropy of activation, respectively, which indicate an associative mechanism. The logarithm of the formation constant of the 1?:?1 complex formed was 11.3.  相似文献   

16.
The thermal decomposition of cyanogen azide (NCN3) and the subsequent collision‐induced intersystem crossing (CIISC) process of cyanonitrene (NCN) have been investigated by monitoring excited electronic state 1NCN and ground state 3NCN radicals. NCN was generated by the pyrolysis of NCN3 behind shock waves and by the photolysis of NCN3 at room temperature. Falloff rate constants of the thermal unimolecular decomposition of NCN3 in argon have been extracted from 1NCN concentration–time profiles in the temperature range 617 K <T< 927 K and at two different total densities: k(ρ ≈ 3 × 10?6 mol/cm3)/s?1=4.9 × 109 × exp (?71±14 kJ mol?1/RT) (± 30%); k(ρ ≈ 6 × 10?6 mol/cm3)/s?1=7.5 × 109 × exp (‐71±14 kJ mol?1/RT) (± 30%). In addition, high‐temperature 1NCN absorption cross sections have been determined in the temperature range 618 K <T< 1231 K and can be expressed by σ /(cm2/mol)= 1.0 × 108 ?6.3 × 104 K?1 × T (± 50%). Rate constants for the CIISC process have been measured by monitoring 3NCN in the temperature range 701 K <T< 1256 K resulting in kCIISC (ρ ≈ 1.8 ×10?6 mol/cm3)/ s?1=2.6 × 106× exp (‐36±10 kJ mol?1/RT) (± 20%), kCIISC (ρ ≈ 3.5×10?6 mol/cm3)/ s?1 = 2.0 × 106 × exp (?31±10 kJ mol?1/RT) (± 20%), kCIISC (ρ ≈ 7.0×10?6 mol/cm3)/ s?1=1.4 × 106 × exp (?25±10 kJ mol?1/RT) (± 20%). These values are in good agreement with CIISC rate constants extracted from corresponding 1NCN measurements. The observed nonlinear pressure dependences reveal a pressure saturation effect of the CIISC process. © 2012 Wiley Periodicals, Inc. Int J Chem Kinet 45: 30–40, 2013  相似文献   

17.
Restricted rotation about the naphthalenylcarbonyl bonds in the title compounds resulted in mixtures of cis and trans rotamers, the equilibrium and the rotational barriers depending on the substituents. For 2,7-dimethyl-1,8-di-(p-toluoyl)-naphthalene (1) ΔH° = 3.66 ± 0.14 kJ mol?1, ΔS° = 1.67 ± 0.63 J mol?1 K?1, ΔHct = 55.5 ± 1.3 kJ mol?1, ΔHct = 51.9 ± 1.3 kJ mol?1, ΔSct = ?41.3±4.1 J mol?1 K?1 and ΔSct = ?42.9±4.1 J mol?1 K?1. The rotation about the phenylcarbonyl bond requires ΔH = ?56.9±4.4 kJ mol?1 and ΔS = ?20.5±15.3 J mol?1 K?1 for the cis rotamer, and ΔH = 43.5Δ0.4 kJ mol?1 and ΔS =± ?22.4Δ1.3 J mol?1 K?1 for the trans rotamer. The role of electronic factors is likely to be virtually the same for both these rotamers but steric interaction between the two phenyl rings occurs in the cis rotamer only. Hence, the difference of the activation enthalpies obtained for the cis and trans rotamers, ΔΔH?1 = 13.4 kJ mol?1, provides a basis for the estimation of the role of steric factors in this rotation. For the tetracarboxylic acid 2 and its tetramethyl ester 3 the equilibrium is even more shifted towards the trans form because of enhanced steric and electrostatic interactions between the substituents in the cis form. The barriers for the rotation around the phenylcarbonyl bond and the cis-trans isomerization are lowered; an explanation for this result is presented.  相似文献   

18.
Gas-phase reactions typical of the Earth’s atmosphere have been studied for a number of partially fluorinated alcohols (PFAs). The rate constants of the reactions of CF3CH2OH, CH2FCH2OH, and CHF2CH2OH with fluorine atoms have been determined by the relative measurement method. The rate constant for CF3CH2OH has been measured in the temperature range 258–358 K (k = (3.4 ± 2.0) × 1013exp(?E/RT) cm3 mol?1 s?1, where E = ?(1.5 ± 1.3) kJ/mol). The rate constants for CH2FCH2OH and CHF2CH2OH have been determined at room temperature to be (8.3 ± 2.9) × 1013 (T = 295 K) and (6.4 ± 0.6) × 1013 (T = 296 K) cm3 mol?1 s?1, respectively. The rate constants of the reactions between dioxygen and primary radicals resulting from PFA + F reactions have been determined by the relative measurement method. The reaction between O2 and the radicals of the general formula C2H2F3O (CF3CH2? and CF3?HOH) have been investigated in the temperature range 258–358 K to obtain k = (3.8 ± 2.0) × 108exp(?E/RT) cm3 mol?1 s?1, where E = ?(10.2 ± 1.5) kJ/mol. For the reaction between O2 and the radicals of the general formula C2H4FO (? HFCH2O, CH2F?HOH, and CH2FCH2?) at T = 258–358 K, k = (1.3 ± 0.6) × 1011exp(?E/RT) cm3 mol?1 s?1, where E = ?(5.3 ± 1.4) kJ/mol. The rate constant of the reaction between O2 and the radicals with the general formula C2H3F2O (?F2CH2O, CHF2?HOH, and CHF2CH2?) at T = 300 K is k = 1.32 × 1011 cm3 mol?1 s?1. For the reaction between NO and the primary radicals with the general formula C2H2F3O (CF3CH2? and CF3?HOH), which result from the reaction CF3CH2OH + F, the rate constant at 298 K is k = 9.7 × 109 cm3 mol?1 s?1. The experiments were carried out in a flow reactor, and the reaction mixture was analyzed mass-spectrometrically. A mechanism based on the results of our studies and on the literature data has been suggested for the atmospheric degradation of PFAs.  相似文献   

19.
In an effort to probe the reaction of antibiotic hydrolysis catalyzed by B3 metallo-??-lactamase (M??L), the thermodynamic parameters of penicillin G hydrolysis catalyzed by M??L L1 from Stenotrophomonas maltophilia were determined by microcalorimetric method. The values of activation free energy ??G ?? ?? are 88.26, 89.44, 90.49, and 91.57?kJ?mol?1 at 293.15, 298.15, 303.15, and 308.15?K, respectively, activation enthalpy ??H ?? ?? is 24.02?kJ?mol?1, activation entropy ??S ?? ?? is ?219.2511?J?mol?1?K?1, apparent activation energy E is 26.5183?kJ?mol?1, and the reaction order is 1.0. The thermodynamic parameters reveal that the penicillin G hydrolysis catalyzed by M??L L1 is an exothermic and spontaneous reaction.  相似文献   

20.
The enthalpy of solvation of the gaseous chromate ion, ΔHsolv0(CrO42?) (g) is estimated on the basis of recent lattice energy studies made in this laboratory, and a charge distribution assigned to the ion. On the basis of the assigned charge of ?0.57 proton units to the oxygen atoms of the CrO42? unit, the total lattice potential energies are found to be: Upot(Na2CrO4) = 1836 kJ mol?1; Upot(K2CrO4) = 1717 kJ mol?1; Upot(Rb2CrO4) = 1645 kJ mol?1 and Upot(Cs2CrO4) = 1598 kJ mol?1. The corresponding value for ΔHsolv0(CrO42?) (g) = ?1077 kJ mol?1.  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