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 46 毫秒
1.
We have used the single‐pulse shock tube technique with postshock GC/MS product analysis to investigate the mechanism and kinetics of the unimolecular decomposition of isopropanol, a potential biofuel, and of its reaction with H atoms at 918‐1212 K and 183‐484 kPa. Experiments employed dilute mixtures in argon of isopropanol, a radical scavenger, and, for H‐atom studies, two different thermal precursors of H. Without an added H source, isopropanol decomposes in our studies predominantly by molecular dehydration. Added H atoms significantly augment decomposition, mainly by abstraction of the tertiary and primary hydrogens, reactions that, respectively, lead to acetone and propene as stable organic products. Traces of acetaldehyde were observed in some experiments above ≈ 1100 K and establish branching limits for minor decomposition pathways. To quantitatively account for secondary chemistry and optimize rate constants of interest, we employed the method of uncertainty minimization using polynomial chaos expansions (MUM‐PCE) to carry out a unified analysis of all datasets using a chemical model–based originally on JetSurF 2.0. We find: k(isopropanol → propene + H2O) = 10(13.87 ± 0.69) exp(?(33 099 ± 979) K/ T) s?1 at 979‐1212 K and 286‐484 kPa, with a factor of two uncertainty (2σ), including systematic errors. For H atom reactions, optimization yields: k(H + isopropanol → H2 + p‐C3H6OH) = 10(6.25 ± 0.42) T2.54 exp(?(3993 ± 1028) K /T) cm3 mol?1 s?1 and k(H + isopropanol → H2 + t‐C3H6OH) = 10(5.83 ± 0.37) T2.40 exp(?(1507 ± 957) K /T) cm3 mol?1 s?1 at 918‐1142 K and 183‐323 kPa. We compare our measured rate constants with estimates used in current combustion models and discuss how hydrocarbon functionalization with an OH group affects H abstraction rates.  相似文献   

2.
The thermal dissociation of gaseous Mo(CO)6 and W(CO)6 in an argon carrier gas, Mo(CO)6 → Mo(CO)5 + CO (1) and W(CO)6 → W(CO)5 + CO (2), is studied over temperature ranges of ∼585–685 K for (1) and ∼690−810 K for (2) at a total gas concentrations of 4 × 10−6 and 4 × 10−5 mol/cm3 by using the shock tube technique in conjunction with absorption spectrophotometry. The measured rate constants are extrapolated to the high-pressure limit by means of a newly developed procedure, with the resultant expressions for the indicated temperature ranges reading as kd1,∞(T),[s−1] = 1016.12 ± 0.68exp[(−148.8 ± 8.1 kJ/mol)/RT] and kd2,∞(T),[s−1] = 1015.93 ± 0.63exp[(−171.7 ± 8.9 kJ/mol)/RT]. Comparison of the high-pressure dissociation rate constants with the published data revealed a considerable discrepancy, a tentative explanation of which is given. Based on the obtained high-pressure dissociation rate constants and the available data on the high-pressure room-temperature rate constants for the reverse reaction of recombination, the first bond dissociation energies for these molecules are evaluated and compared with previous determinations, both theoretical and experimental. The enthalpies of formation of Mo(CO)5 and W(CO)5 are determined: ΔfH°(Mo(CO)5, g, 298.15 K) = −644.1 ± 5.6 kJ/mol and ΔfH°(W(CO)5, g, 298.15 K) = −581.9 ± 6.6 kJ/mol. Based on the enthalpies of formation of Mo(CO)5, W(CO)5, Mo(CO)6, and W(CO)6, and the published molecular parameters of these four species, their thermochemical functions are calculated and presented in the form of NASA seven-term polynomials.  相似文献   

3.
Low-temperature heat capacities of the compound Ni(C4H7O5)2·2H2O(S) have been measured with an auto- mated adiabatic calorimeter. A thermal decomposition or dehydration occurred in 350--369 K. The temperature, the enthalpy and entropy of the dehydration were determined to be (368.141 ±0.095) K, (18.809±0.088) kJ·mol ^-1 and (51.093±0.239) J·K^-1·mol^-1 respertively. The experimental values of the molar heat capacities in the temperature regions of 78-350 and 368-390 K were fitted to two polynomial equations of heat capacities (Cp,m) with the reduced temperatures (X), [X=f(T)], by a least squares method, respectively. The smoothed molar heat capacities and thermodynamic functions of the compound were calculated on the basis of the fitted polynomials. The smoothed values of the molar heat capacities and fundamental thermodynamic functions of the sample relative to the standard reference temperature 298.15 K were tabulated with an interval of 5 K.  相似文献   

4.
Three chromium(III) complexes of general formula [Cr(ox)2(pdaH)]2− (where ox = C2O4 2− and pdaH is N,O-bonded 2,3-, 2,4- or 2,5-pyridinedicarboxylic acid anion) were obtained and characterized in solution. Acid-catalysed aquation of [Cr(ox)2(pdaH)]2− gave two products: [Cr(ox)(pdaH)(H2O)2]0 (P1) and cis-[Cr(ox)2(H2O)2]2− (P2). The kinetics of these reactions were studied spectrophotometrically, within the 0.1–1.0 M HClO4 range, and the pseudo-first-order rate constants for the oxalato (k obs1) and pdaH (k obs2) ligands dissociation were calculated based on the determined pseudo-first-order rate constants (k obs) and P1:P2 molar ratio. The dependencies of the pseudo-first-order rate constants on [H+] are as follows: k obs1 = b 1[H+] and k obs2 = b 2[H+], where b 1 and b 2 are the second-order rate constants for the oxalato and pdaH ligands dissociation, respectively. Kinetic parameters were determined and the mechanism of the pdaH ligand dissociation is proposed.  相似文献   

5.
Decomposition of formic acid (HCO2H) proceeds via three unimolecular channels: dehydration, decarboxylation, and dissociation, the latter expected to be of minor contribution to the overall kinetics. In addition, despite the similar values reported for the individual activation energies for the dehydration and decarboxylation reactions, experimental works have shown that the former is dominant in the reaction mechanism. These reactions show pressure-dependent rate coefficients, and the high-pressure condition is not yet verified at atmospheric pressure. This work aims to investigate the influence of temperature and pressure on the rate coefficients. Hence, theoretical calculations at the CCSD(T)/CBS level have been performed to accurately describe the unimolecular reaction and Rice-Ramsperger-Kassel-Marcus (RRKM) rate coefficients have been calculated and integrated for the prediction of k(T,P) rate coefficients, adopting both strong and weak collision models, over the intervals 0.5-10 atm and 298-2200 K. Our results suggest that the isomerization path is important and explains the preference for the (CO + H2O) channel. Rate coefficients for the (CO2 + H2) and (CO + H2O) formations are given, in s−1, as exp(−34404/T) and exp(−33785/T), respectively. The dissociation limit of 107.29 kcal mol–1, with respect the Z-HCO2H conformer, leading to OH + HCO, via a barrierless potential curve, with rate coefficients, in s−1, expressed as kHCO+OH(T) = 1.68 × 1017 exp(−56018/T). Temperature and pressure dependence for the HCO + OH → CO2 + H2 and HCO + OH → CO + H2O reactions have also been estimated.  相似文献   

6.
Standard electrode potentials E° of Ag-AgC1 electrode in molality scale and acidityconstants of glyeine pK_1° at constant molality of NaCl (1.0 mol·kg~(-1)) in 5 and 15 mass%glucose-water mixed solvents over a range of temperatures from 278.15 to 318.15 K weredetermined from precise emf measurements.The dependence of acidity constant on temperatureis given as a function of the thermodynamic temperature T by an empirical equation, pK_1°=A_1(K/T)-A_2+A_3(T/K).The corresponding thermodynamic quantities of the first dissociationprocess of glycine were calculated and the effects of both tho solvent and the salt on themwere also discussed.  相似文献   

7.
The capture of rotationally state-selected and unselected asymmetric top polar molecules by ions is investigated. Analytical expressions (for all rotational states up to j = 2) of capture rate constants in the perturbed-rotor second-order limit are derived for application to low temperature conditions. Approximate analytical representations over wider temperature ranges are also given for rotationally unselected molecules. The capture of H2O, D2O, and HDO by arbitrary ions is chosen for demonstration of the approach. Capture rate constants for the about 60 reactions of H2O with ions listed in the UMIST 2006 data base for astrochemistry are calculated, compared with experimental data, and represented in the format kcap(T) ≈ c1 + c2(T/300 K)−1/2. The parameters c1 and c2 can be predicted in a very simple way. The approach allows one to identify capture-controlled mechanisms and/or to trace experimental artifacts. The approach applies equally well to the capture of symmetric top and linear dipole molecules by arbitrary ions.  相似文献   

8.
Using ab initio [SCS‐MP2 and CCSD(T)] and density functional theory (M062X) calculations, we have studied the geometries and energies of sulfur oxoacids H2SmO6 (m = 2–4) and their monohydrated and dihydrated clusters. When including the results from previously reported disulfuric acid (H2S2O7) cases, the gas phase acidity is ordered as H2S2O6 < H2S3O6 < H2S2O7 < H2S4O6. The intramolecular H‐bonding, which may indicate the degree of structural flexibility in this molecular series, is an important factor for the order of the gas phase acidity. All these sulfur oxoacids show dissociated (or deprotonated) geometries with only two water molecules, although the energies of the dissociated conformers are ranked differently. All of the dissociated conformers form a unique H‐bonding network structure in which the protonated first water (H3O+) is triply H‐bonded to each oxygen atom of two SO3 moieties as well as the second water, which in turn is H‐bonded to a SO3 moiety. H2S3O6 has the best molecular flexibility for adopting such an H‐bonding network structure, and thereby all the low‐lying conformers of H2S3O6(H2O)2 are dissociated. In contrast, the least flexible H2S2O6 forms such a structure with a high strain, and dissociation of H2S2O6(H2O)2 is found from the third lowest conformer. Although the gas phase acidity of H2S4O6 is the highest in this series, the lowest dissociated conformer and the lowest undissociated conformer of H2S4O6(H2O)2 are very close in energy. This is because forming the H‐bonding network structure is somewhat difficult due to the large distance between the two SO3 moieties.  相似文献   

9.
Thermodynamic properties (ΔH°f(298), S°(298) and Cp(T) from 300 to 1500 K) for reactants, adducts, transition states, and products in reactions of CH3 and C2H5 with Cl2 are calculated using CBSQ//MP2/6‐311G(d,p). Molecular structures and vibration frequencies are determined at the MP2/6‐311G(d,p), with single‐point calculations for energy at QCISD(T)/6‐311 + G(d,p), MP4(SDQ)/CbsB4, and MP2/CBSB3 levels of calculation with scaled vibration frequencies. Contributions of rotational frequencies for S°(298) and Cp(T)'s are calculated based on rotational barrier heights and moments of inertia using the method of Pitzer and Gwinn [1]. Thermodynamic parameters, ΔH°f(298), S°(298), and CP(T), are evaluated for C1 and C2 chlorocarbon molecules and radicals. These thermodynamic properties are used in evaluation and comparison of Cl2 + R· → Cl· + RCl (defined forward direction) reaction rate constants from the kinetics literature for comparison with the calculations. Data from some 20 reactions in the literature show linearity on a plot of Eafwd vs. ΔHrxn,fwd, yielding a slope of (0.38 ± 0.04) and intercept of (10.12 ± 0.81) kcal/mole. A correlation of average Arrhenius preexponential factor for Cl· + RCl → Cl2 + R· (reverse rxn) of (4.44 ± 1.58) × 1013 cm3/mol‐sec on a per‐chlorine basis is obtained with EaRev = (0.64 ± 0.04) × ΔHrxn,Rev + (9.72 ± 0.83) kcal/mole, where EaRev is 0.0 if ΔHrxn,Rev is more than 15.2 kcal/mole exothermic. Kinetic evaluations of literature data are also performed for classes of reactions. Eafwd = (0.39 ± 0.11) × ΔHrxn,fwd + (10.49 ± 2.21) kcal/mole and average Afwd = (5.89 ± 2.48) × 1012 cm3/mole‐sec for hydrocarbons: Eafwd = (0.40 ± 0.07) × ΔHrxn,fwd + (10.32 ± 1.31) kcal/mole and average Afwd = (6.89 ± 2.15) × 1011 cm3/mole‐sec for C1 chlorocarbons: Eafwd = (0.33 ± 0.08) × ΔHrxn,fwd + (9.46 ± 1.35) kcal/mole and average Afwd = (4.64 ± 2.10) × 1011 cm3/mole‐sec for C2 chlorocarbons. Calculation results on the methyl and ethyl reactions with Cl2 show agreement with the experimental data after an adjustment of +2.3 kcal/mole is made in the calculated negative Ea's. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 548–565, 2000  相似文献   

10.
11.
12.
Zusammenfassung Es wurde die systematische Untersuchung des Systems im sauren Bereich durchgeführt und die Stoffe der Zusammensetzung* La2(H2 T)3·9 H2O und H[La(H2 T)2]·3 H2O isoliert. Diese wurden röntgenographisch untersucht und ihre thermische Analyse durchgeführt. Die IR-Absorptionsbanden werden zugeordnet. Auf Grund von pH-Messungen wurden die Dissoziationskonstanten bestimmt.
A systematic investigation of the system in the acidic range was performed, and the compounds* La2(H2 T)3·9 H2O and H[La(H2 T)2]·3 H2O were isolated. X-ray powder diagrams were taken, and their thermal analysis carried out. The IR-absorption bands were assigned, and the dissociation constants, based on pH-measurements determined.


Mit 2 Abbildungen  相似文献   

13.
Zusammenfassung Auf Grund von pH-Messungen wurden die erste und zweite Dissoziationskonstante der Weinsäure und Dissoziationskonstanten der KomplexionenLnH2 T +,Ln 2H2 T 4+,Ln(H2 T)2-undLnHT bestimmt.
The first and second constant of dissociation of tartaric acid, and the constants of dissociation of the complex ionsLnH2 T +,Ln 2H2 T 4+,Ln(H2 T)2- andLnHT were determined, making use of pH-measurements.


Zugl. 30. Mitt. der Reihe: Koordinationsverbindungen mit organischen Hydroxysäuren.

Ln 3+=La3+, Ce3+.

H4 T=C4H6O6.  相似文献   

14.
The overall rate constants for H-abstraction (kH) from tetrahydrofuran and D-abstraction (kD) from fully deuterated tetrahydrofuran by chlorine atoms in the temperature range of 298-547 K were determined. In both cases, very weak negative temperature dependences of the overall rate constants were observed, described by the expressions: kH = (1.55 ± 0.13) × 10−10 exp(52 ± 28/T) cm3 molecule−1 s−1 and kD = (1.27 ± 0.25) × 10−10exp(55 ± 62/T) cm3 molecule−1 s−1. The experimental results show that the value of the kinetic isotope effect (kH/kD), amounting to 1.21 ± 0.10, is temperature independent at 298-547 K.  相似文献   

15.
Ab initio molecular orbital calculations at the G2(MP2) level have been carried out on cyclopropylsilylene C3H5SiH. Four equilibrium structures were located. Like H2Si, the ground state of C3H5SiH is singlet and the triplet is the low‐lying excited state. The singlet–triplet separation energy is 127.9 kJ/mol. The cis‐trans isomerization path of singlet cyclopropylsilylene was investigated by intrinsic reaction coordinate (IRC) calculations. The calculations show that no gauche conformers exist along the potential energy curve of the cis‐trans isomerization and the isomerization happens with a barrier of 30.1 kJ/mol. Changes (ΔH and ΔG) in thermodynamic functions, equilibrium constant K(T), and A factor and reaction rate constant k(T) in Eyring transition state theory of the cis‐trans isomerization were also calculated. © 2001 John Wiley & Sons, Inc. Int J Quantum Chem, 2001  相似文献   

16.
This article studies the reactions and mechanisms of H8Si8O12 (T8H8) molecules with n-propanol, acetone, allyl alcohol, n-butylamine, allylamine, acetic acid, and 1-octene in air, at room temperature, and without catalysts. The reaction between T8H8 and n-propanol involves both the highly polarized Si O and Si H bonds and results in cage breakage and forming Q4 and Q3 structures with  OC3H7 in the reaction product. T8H8 also reacts with acetone, and the resultant product possesses Si OCH(CH3)2. Allyl alcohol is less reactive to cause T8H8 decomposition, and the resultant product contains Si OCH2CHCH2 and Si OCH2(CH2)3CHCH2. However, it is found that basically T8H8 does not react with acetic acid and 1-octene. In the reactions of T8H8 with n-butylamine and allylamine, the resultant products contain Si NH(CH2)3CH3 and Si NHCH2CHCH2, respectively. For the reaction with T8H8, allylamine is less active than n-butylamine. Possible mechanisms for the T8H8 reactions are discussed.  相似文献   

17.
The spin-forbidden dissociation reaction of the N2O(X1Σ+) ground state has been investigated by both quantum calculations and experiments. Ab initio prediction at the CCSD(T)/CBS(TQ5)//CCSD(T)/aug-cc-pVTZ+d level of theory gave the crossing point (MSX) energy at 60.1 kcal/mol for the N2O(X1Σ+) → N2() + O(3P) transition, in good agreement with published data. The T- and P-dependent rate coefficients, k1(T,P), for the nonadiabatic thermal dissociation predicted by nonadiabatic Rice-Ramsperger-Kassel-Marcus (RRKM) calculations agree very well with literature data. The rate constants at the high- and low-pressure limits, k1 = 1011.90 exp (−61.54 kcal mol−1/RT) s−1 and k1o = 1014.97 exp(−60.05 kcal mol−1/RT) cm3 mol−1 s−1, for example, agree closely with the extrapolated results of Röhrig et al. at both pressure limits. The second-order rate constant (k1o) is also in excellent agreement with our result measured by FTIR spectrometry in the present study for the temperature range of 860-1023 K as well as with many existing high-temperature data obtained primarily by shock-wave heating up to 3340 K. Kinetic modeling of the NO product yields measured at long reaction times in the present work also allowed us to reliably estimate the rate constant for reaction (3), O + N2O → N2 + O2, based on its strong competition with the NO formation from reaction (2) which has been better established. The modeled values of k3 confirmed the previous finding by Davidson et al. with significantly smaller values of A-factor and activation energy than the accepted ones. A least-squares analysis of both sets of data gave k3 = 1012.22 ± 0.04 exp[− (11.65 ± 0.24 kcal mol−1/RT)] cm3 mol−1 s−1, covering the wide temperature range of 988-3340 K.  相似文献   

18.
For a set of 32 selected free radicals, energy minimum structures, harmonic vibrational wave numbers ωe, principal moments of inertia IA, IB, and IC, heat capacities C°p(T), entropies S°(T), thermal energy contents H°(T) ? H°(0), and standard enthalpies of formation ΔfH°(T) were calculated at the G3MP2B3 level of theory in the temperature range 200–3000 K. In this article, thermodynamic functions at T = 298.15 K are presented and compared with recent experimental values. The mean absolute deviation between calculated and experimental ΔfH°(298.15) values resulted in 3.91 kJ mol?1, which is close to the average experimental uncertainty of ± 3.55 kJ mol?1. The influence of hindered rotation on thermodynamic functions is studied for isopropyl and tert‐butyl radicals. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 550–560, 2002  相似文献   

19.
For plasma enhanced and catalytic chemical vapor deposition (PECVD and Cat‐CVD) processes using small silanes as precursors, disilanyl radical (Si2H5) is a potential reactive intermediate involved in various chemical reactions. For modeling and optimization of homogeneous a‐Si:H film growth on large‐area substrates, we have investigated the kinetics and mechanisms for the thermal decomposition of Si2H5 producing smaller silicon hydrides including SiH, SiH2, SiH3, and Si2H4, and the related reverse reactions involving these species by using ab initio molecular‐orbital calculations. The results show that the lowest energy path is the production of SiH + SiH4 that proceeds via a transition state with a barrier of 33.4 kcal/mol relative to Si2H5. Additionally, the dissociation energies for breaking the Si? Si and H? SiH2 bonds were predicted to be 53.4 and 61.4 kcal/mol, respectively. To validate the predicted enthalpies of reaction, we have evaluated the enthalpies of formation for SiH, SiH2, HSiSiH2, and Si2H4(C2h) at 0 K by using the isodesmic reactions, such as 2HSiSiH2 + 1C2H61Si2H6 + 2HCCH2 and 1Si2H4(C2h) + 1C2H61Si2H6 + 1C2H4. The results of SiH (87.2 kcal/mol), SiH2 (64.9 kcal/mol), HSiSiH2 (98.0 kcal/mol), and Si2H4 (68.9 kcal/mol) agree reasonably well previous published data. Furthermore, the rate constants for the decomposition of Si2H5 and the related bimolecular reverse reactions have been predicted and tabulated for different T, P‐conditions with variational Rice–Ramsperger–Kassel–Marcus (RRKM) theory by solving the master equation. The result indicates that the formation of SiH + SiH4 product pair is most favored in the decomposition as well as in the bimolecular reactions of SiH2 + SiH3, HSiSiH2 + H2, and Si2H4(C2h) + H under T, P‐conditions typically used in PECVD and Cat‐CVD. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
Theoretical studies are carried out on the multi-channel reactions of SiH(CH3)3 with Cl (reaction 1, R1) and Br atoms (R2) by direct dynamics method. The minimum energy path is calculated at the MP2/6-31+G(d,p) level, and energetic information is further refined by the MC-QCISD (single-point) method. The rate constants for individual reaction channels, R1a, R1b-in, R1b-out, R1c, R1d, R2a, R2b-in, R2b-out, R2c, and R2d, are calculated by the improved canonical variational transition state theory with small-curvature tunneling correction over the temperature range 200–1,500 K. The theoretical three-parameter expressions k 1 (T) = 6.30 × 10−15 T 1.36exp(704.94/T) and k 2 (T) = 9.41 × 10−26 T 4.89exp(−1,033.80/T) cm3 molecule−1 s−1 are given. Our calculations indicate that reaction channels R1c and R2c are the major channel.  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