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 390 毫秒
1.
Sesquiterpenes are constituents of a variety of essential oils that are used in flavorings, perfumes, personal care, and cleaning products. Two sesquiterpenes that are commonly used as indoor fragrances are valencene and farnesol. Knowing the reaction rate constants of these chemicals with ozone (O3) and nitrate radical () is an important factor in determining their fate indoors. In this study, the bimolecular rate constants of , , , and were measured using the relative rate technique at 297 ± 3 K and 1 atm total pressure. Using the rate constants reported here and measured/modeled indoor concentrations of O3 and (20 ppb and 1 ppt, respectively), pseudo–first‐order‐rate lifetimes , , , and were determined.  相似文献   

2.
Accurate prediction of the temperature response of the velocity v of a biochemical reaction has wide applications in cell biology, reaction design, and biomass yield enhancement. Here, we introduce a simple but comprehensive mechanistic approach that uses thermodynamics and biochemical kinetics to describe and link the reaction rate and Michaelis–Menten constants (kT and T) with the biomass yield and mortality rate (T and δT) as explicit functions of . The temperature control is exerted by catabolic enthalpy at low temperatures and catabolic entropy at high temperatures, whereas changes in cell and enzyme–substrate heat capacity shift the anabolic electron use efficiency eA and the maximum reaction velocity vmax. We show that cells have optimal growth when the catabolic (differential) free energy of activation decreases the cell free energy harvest required to duplicate their internal structures as long as electrons for anabolism are available. With the described approach, we accurately predicted observed glucose fermentation and ammonium nitrification dynamics across a wide temperature range with a minimal number of thermodynamics parameters, and we highlight how kinetic parameters are linked to each other using first principles.  相似文献   

3.
The gas‐phase eliminations of benzyl chloroformate (475–523 K, 31–103 Torr) and neopentyl chloroformate (563–622 K, 37–70 Torr), in a deactivated static reaction vessel, and in the presence of a free radical suppressor, are homogeneous, unimolecular, and follow a first‐order rate law. The rate coefficients are expressed by the following Arrhenius equations: Benzyl chloroformate Neopentyl chloroformate Formation of neopentyl chloride: Formation of 2‐methylbutenes: The derived kinetic and thermodynamic parameters for benzyl chloroformate decomposition indicate the reaction proceeds through a concerted four‐membered cyclic transition state to give benzyl chloride and CO2 gas. Neopentyl chloroformate undergoes a parallel reaction, where neopentyl chloride formation may arise from a polar‐concerted four‐membered cyclic transition state, whereas the mixture of olefins, 2‐methyl‐2‐butene, and 2‐methyl‐1‐butene appears to be produced from a carbene intermediate. This intermediate seems to be originated from a concerted five‐membered cyclic transition state of the neopentyl substrate.  相似文献   

4.
Multi‐Species Multi‐Channel (MSMC) is an ab initio parallel program to calculate thermodynamic quantities (e.g., , , , and , time‐dependent species profiles, and rate coefficients as functions of temperature and pressure for complex chemical reaction systems, which consist of multiple stable species and multiple reaction channels interconnecting them. Thermodynamic properties of the species involved are calculated using statistical mechanics with molecular information from electronic structure calculations. Temperature‐ and pressure‐dependent behaviors are rigorously characterized within the eigenpair master equation/Rice–Ramsperger–Kassel–Marcus (ME/RRKM) framework. Corrections, e.g., for hindered internal rotation and tunneling treatment, are included. With the implementation of an ultra‐high precision package and rigorous matrix setup, MSMC is able to correctly mimic real behaviors of different types of chemical systems. Different eigenpair‐based approaches to extract phenomenological/macroscopic rate coefficients are implemented for different applications. Moreover, a friendly and platform‐independent graphical‐user‐interface (GUI) is provided to facilitate the use of MSMC and the pre‐/postcalculation data visualization/analysis on the fly. The program can be freely downloaded at https://sites.google.com/site/msmccode/ .  相似文献   

5.
A detailed chemical kinetic model for oxidation of CH3OH at high pressure and intermediate temperatures has been developed and validated experimentally. Ab initio calculations and Rice–Ramsperger–Kassel–Marcus/transition state theory (RRKM/TST) analysis were used to obtain rate coefficients for , , , and . The experiments, involving CH3OH/O2 mixtures diluted in N2, were carried out in a high‐pressure flow reactor at 600–900 K and 20–100 bar, varying the reaction stoichiometry from very lean to fuel‐rich conditions. Under the conditions studied, the onset temperature for methanol oxidation was not dependent on the stoichiometry, whereas increasing pressure shifted the ignition temperature toward lower values. Model predictions of the present experimental results, as well as rapid compression machine data from the literature, were generally satisfactory. The governing reaction pathways have been outlined based on calculations with the kinetic model. Unlike what has been observed for unsaturated hydrocarbons, the oxidation pathways for CH3OH under the investigated conditions were very similar to those prevailing at higher temperatures and lower pressures. At the high pressures, the modeling predictions for onset of reaction were particularly sensitive to the reaction.  相似文献   

6.
Dimethylamine (DMA) ignition delay times and OH time histories during the oxidation process were investigated behind reflected shock waves. The ignition delay time measurements cover the temperature range of 1181–1498 K, with pressures near 0.9, 1.5, and 2.8 atm, and equivalence ratios of 0.5, 1, and 2, in 4% oxygen/argon. The ignition delay time data feature low scatter and can be correlated to a single expression with 2 ~ 0.99: τign = 7.30 × 10?4 ?0.68 Φ0.45 exp(18,265/), where τign is in μs, in atm, and in K. OH time histories were measured using laser absorption of the R1(5) line of the A‐X(0,0) transition near 306.7 nm, in stoichiometric mixtures of 500 ppm DMA/O2/argon. The mechanism developed by Li et al. was used initially to simulate the measured DMA ignition delay times and the OH time histories. The Li et al. mechanism was then updated by adding the DMA unimolecular decomposition channel: DMA = CH3NH + CH3, with the reaction rate constant estimated by analogy to dimethyl ether decomposition, previously investigated by Cook et al. The reactions of DMA + OH were also updated based on recent work in our laboratory. The simulation results using the modified Li et al. mechanism are in good agreement with both the ignition delay times and OH time‐history data.  相似文献   

7.
We report shock tube studies of the kinetics of H atom addition to cyclopentene and modeling of the subsequent decomposition of cyclopentyl. Hydrogen atoms were generated with thermal precursors in dilute mixtures of cyclopentene and a reference compound in argon. Addition of H to the double bond leads to a cyclopentyl radical that rapidly ring opens and decomposes to ethene and allyl radical. The process was monitored by postshock gas chromatographic analysis of ethene and rate constants determined relative to H atom displacement of methyl from 1,3,5‐trimethylbenzene (135TMB). At 863–1167 K and 160–370 kPa, we find and, with , we obtain Using experimental values of about 3:1 for the ratio of C─C to C─H beta scission in cyclopentyl radicals and a corresponding transition‐state‐theory/Rice‐Ramsberger‐Kassel‐Marcus (TST/RRKM) model, the high‐pressure rate expression for addition of H to cyclopentene at 863–1167 K is derived as Combined with literature results from lower temperatures and a fitted TST model, the rate expression between 298 and 2000 K is determined as Results are compared with related systems. Near 1000 K, our data require a minimum value of 1.5 for branching between beta C─C and C─H scission in cyclopentyl radicals to maintain established trends in H addition rates. This conflicts with current computed values and those used in existing kinetics models of cyclopentane combustion. We additionally report and discuss minor observed channels in the decomposition of cyclopentene, including formation of 1,4‐pentadiene, (E/Z)‐1,3‐pentadiene, 1,3‐butadiene, and the direct elimination of H2 from cyclopentene to give cyclopentadiene.  相似文献   

8.
The rates of reactions of N‐substituted benzyl amines with benzyl bromide were measured using a conductivity technique in methanol medium. The reaction followed a total second‐order path. The end product of the reaction is identified as dibenzyl alkyl amine (C6H5CH2N(R)CH2C6H5). The rates increased with a decrease in the electron‐donating capacity or with an increase in the Taft σ* value of electron‐donating alkyl substituents (R) such as t‐butyl (σ* = ?0.3), i‐propyl (σ* = ?0.19), n‐butyl (σ* = ?0.13), and ethyl (σ* = ?0.1) on nitrogen of the amine until the Taft σ* value becomes zero for the methyl group ( = 0.00), and then the rates decreased with an increase in the electron‐withdrawing capacity or with an increase in the Taft σ* value of electron‐withdrawing substituents (R) such H and C6H5 ( = 0.49 and = 0.6). The locus of the Taft polar free energy relationship has a maximum near the point for N‐methyl benzyl amine, showing that there is a sharp change in the rate‐determining step. A mechanism involving formation of an SN2‐type transition state between the amine nucleophiles and the benzyl bromide and its subsequent decomposition is proposed. Activation parameters were calculated and are discussed.  相似文献   

9.
In this paper, copper composite anticorrosion and antifouling coatings were prepared by a cold spray technique. Polarization experiments of the coatings were performed by rotating ring‐disk electrode technology at high potential. The electrochemical reaction mechanisms were proposed, and corresponding polarization kinetics models were built. Experimental results show that the copper and cuprous oxide formed corrosion microcells in the coatings, and the cuprous oxide did not alter the electrochemical reaction process of copper. In the high potential region (about 0.2–0.8 V), a CuCl film formed on the surface of the coatings was not damaged or broken down. The film played a role in corrosion protection. The currents in the high potential region increased relative to the limiting current 1. Because in the high potential region, was produced by the dissolution of the CuCl film and was oxidized to Cu2+. In addition to being oxidized to Cu2+, has the other two destinations, which were deposition as a CuCl film and diffusion to the solution bulk. The three processes were in parallel competition relations. In the limiting current 2 region, oxidation of was dominant. A rate‐controlling step of electrochemical dissolution of the coatings in the high potential region was the diffusion processes of and Cu2+. The electrochemical polarization kinetic models based on the reaction mechanisms established in this research accorded well to the experimental results. It demonstrated the rationality of polarization kinetics models and reaction mechanisms.  相似文献   

10.
The heterogeneous ozonolysis of sodium oleate aerosols in an aerosol flow tube is reported under different relative humidity (RH%) conditions. Submicron sodium oleate particles were exposed to a known ozone concentration and the consumption of sodium oleate was monitored by infrared spectroscopy. When the experimental results are treated as a surface‐mediated reaction (i.e., following a Langmuir–Hinshelwood type mechanism), the following parameters are obtained: at low RH%, = (3 ± 1) × 10?16 cm3 molecule?1 and = (0.046 ± 0.006) s?1; at high RH%, = (6 ± 2) × 10?16 cm3 molecule?1 and = (0.21 ± 0.05) s?1. From these pseudo–first‐order coefficients, the reactive uptake coefficients for dry and aqueous sodium oleate aerosols are calculated as (1.5 ? 0.5) × 10?7 and (1.7 ? 0.7) × 10?6, respectively. Hydrated oleate aerosols display both an increase in the ozone trapping ability and an increase in the effective rate reaction at the droplet surface compared to dry aerosol surfaces. These observations may provide an explanation for some of the variability observed between lab studies of dry ozonolysis and real‐world, atmospheric lifetimes of oleic acid–related species.  相似文献   

11.
We have investigated the kinetics for the reaction CH3O? + NO2 in N2 bath gas. The rate constants are well‐fit by the Troe expression over the temperature (250–335 K) and pressure range (30–700 Torr) investigated. The termolecular rate constant is given by cm6 molecule?2 s?1, and the rate constant at the high‐pressure limit is given by cm3 molecule?1 s?1. We also studied the kinetics of the reaction of CD3O? + NO2 as a function of temperature and pressure under similar conditions as those for CH3O? + NO2. The resulting low‐ and high‐pressure limiting rate constants are cm6 molecule?2 s?1 and cm3 molecule?1 s?1, respectively. The rate constants for the two isotopologues track each other closely as the high‐pressure limit is approached. The present results agree with most previous results at 295 K over a range of pressures, but there is substantial disagreement about the temperature dependence.  相似文献   

12.
The HO2 yield in the reaction of peroxy radicals with OH radicals has been determined experimentally at 50 Torr helium by measuring simultaneously OH and HO2 concentration time profiles, following the photolysis of XeF2 in the presence of different hydrocarbons and O2. The following yields have been obtained:  = (0.90 ± 0.1),  = (0.75 ± 0.15),  = (0.41 ± 0.08), and  = (0.15 ± 0.03). The clear decrease in HO2 yield with increasing size of the alkyl moiety can be explained by an increased stabilization of the trioxide adduct, ROOOH. This has been confirmed by ab initio and Rice–Ramsperger–Kassel–Marcus master equation calculations. Extrapolation of the experimental results to atmospheric conditions shows that the stabilized adduct, ROOOH, is the nearly exclusive product of the reaction between OH radicals and peroxy radicals containing more than three C‐atoms. The fate and possible impact of these species is completely unexplored so far.  相似文献   

13.
The kinetics of the Diels–Alder (DA) reactions of 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione 1 , trans‐diethyl azodicarboxylate 2 , and tetracyanoethene 3 with 1,3‐cyclohexadiene 4 , cycloheptatriene 5 , 1,3‐cycloheptadiene 6 , cyclooctatetraene 7 , and 1,3‐cyclooctadiene 8 in a range of temperatures and pressures has been studied. Values of the enthalpy, entropy, and volume of activation, as well as the enthalpy and volume of reaction have been obtained. Observed reaction rates of 5+1 and 7+1 have been compared with the known rate of norcaradiene 17 formation in the equilibrium , and that of bicyclo[4,2,0]‐octa‐2,4,7‐triene 20 in the equilibrium . The kinetic data show that the rate of formation of 17 from 5 is much greater than the loss rate of dienophile 1 in reaction of 5+1 . In contrast, the formation rate of tautomer 20 is less than the loss rate of dienophile 1 in reaction of 7+1 . This reflects that the consecutive reaction of 5 → 17 (+ 1 ) → 15 is possible whereas the consecutive reaction of 7 → 20 (+ 1 ) → 22 does not occur as the only way.  相似文献   

14.
The combination of relative rate method with discharge flow and mass spectrometry (RR/DF/MS) technique was employed to determine the rate constant for the gas‐phase reaction of hydroxyl radicals (OH) with naphthalene at 240?340 K and a total pressure of 1–3 Torr. At 298 K, the rate constant was measured to be cm3 molecule?1 s?1, which is in good agreement with reported literature values determined using different techniques. The reaction of OH with naphthalene was found to be essentially independent of pressure in a range of 1?3 Torr at both 298 and 340 K. At 240–340 K, the rate constant of this reaction was found to be negatively dependent on temperature, with an Arrhenius expression of k1(T) cm3 molecule?1 s?1 and k1(T) cm3 molecule?1 s?1 using 1,4‐dioxane and styrene as the reference compounds, respectively. The atmospheric lifetime of naphthalene was estimated to be 9.6 h using the rate constant of naphthalene + OH determined at 277 K in the present work.  相似文献   

15.
The triazines 3‐(2‐pyridyl)‐5,6‐diphenyl‐1,2,4‐triazine (PDT), 3‐(4‐phenyl‐2‐pyridyl)‐5,6‐diphenyl‐1,2,4‐triazine (PPDT), and 1,10‐phenanthroline (phen) were coordinated to the Fe2+ ion to form (1) , (2) , , (3) and (4) . The complexes were synthesized and characterized by mass spectroscopy and elemental analysis. The rate of substitution of these complexes by 2,2′‐bipyridine (bpy), 1,10‐phenanthroline (phen), and 2,2′,6,2″‐terpyridine (terpy) was studied in a sodium acetate–acetic acid buffers over the range 3.6–5.6 at 25, 35, and 45°C under pseudo–first‐order conditions. The reactions are first order with respect to the concentration of the complexes. The reaction rates increase with increasing [bpy/phen/terpy] and pH, whereas ionic strength has no influence on the rate of reaction. Plots of k obs versus [bpy/phen/terpy] and 1/[H+] are linear with positive slopes and significant y‐intercepts. This indicates that the reactions proceed by both dissociative as well as associative pathways for which the associative pathway predominates the substitution kinetics. Observed temperature‐depended rate constants at the three temperatures at which substitution reactions were studied together with the protonation constants of the substituting ligands (phen, bpy, terpy) were used to evaluate the specific rate constants (k 1 and k 2) and thermodynamic parameters (Ea , ΔH #, ΔS #, and ΔG #). The reactivity order of the four complexes depends on the phenyl groups present on the triazine (PDT/PPDT) molecule. The π‐electrons on phenyl rings stabilizes the charge on the metal center by inductive donation of electrons toward the metal center resulting in a decrease in reactivity of the complex, and the order is 1 < 2 < 3 < 4 . The rate of substitution is also influenced by the basicity of the incoming ligand (bpy/phen/terpy), and it decreased in the order: phen > terpy > bpy. Higher rate constants, low Ea values, and more negative entropy of activation (−ΔS #) values were observed for the associative path, revealing that substitution reactions at the octahedral iron(II) complexes by bpy, phen, and terpy occur predominantly by the associative mechanism. Density functional theory calculations support the interpretations.  相似文献   

16.
Emissions of biogenic volatile organic compounds are higher than those from anthropogenic sources. In this work, we studied the kinetics of the reaction of three unsaturated aldehydes (trans‐2‐pentenal, trans‐2‐hexenal, and 2‐methyl‐2‐pentenal) with ozone in a rigid atmospheric simulation chamber coupled to an FTIR spectrometer at four different temperatures (273, 298, 333, and 353 K). Reaction rate constants (× 10−18 cm3 molecule−1 s−1) at 298 K are 1.24 ± 0.06 for trans‐2‐pentenal (t‐2P), 1.37 ± 0.03 for trans‐2‐hexenal (t‐2H), and 1.58 ± 0.20) for 2‐methyl‐2‐pentenal (2M2P). The following Arrhenius expressions were deduced (cm3 molecule−1 s−1): The obtained data are presented and compared to those reported in the literature at room temperature, as well as to homologous alkenes. The atmospheric lifetimes with respect to ozone, derived from this study, are estimated to vary between 7 and 10 days.  相似文献   

17.
In this work, we have measured the rate coefficients of the reactions of isopropyl (propan‐2‐yl), sec‐butyl (butan‐2‐yl), and tert‐butyl (2‐methylpropan‐2‐yl) radicals with molecular chlorine as a function of temperature (190–480 K). The experiments were done in a tubular laminar flow reactor coupled to a photoionization quadrupole mass spectrometer employing a gas‐discharge lamp for ionization. The radicals were homogeneously produced in the reactor by photolyzing suitable precursor molecules with 193‐nm pulsed exciplex laser radiation. The bimolecular rate coefficients were obtained by monitoring the radical decay signals in real time under pseudo–first‐order conditions. The rate coefficients of all three reactions showed negative temperature dependence. The bath gas used in the experiments was helium, and the rate coefficients appeared to be independent of the helium concentrations employed ([2.4–14] × 1016 cm?3) for all three reactions. The rate coefficients of the reactions can be approximated in the studied temperature range by the following parameterizations: We estimate that the overall uncertainties of the measured rate coefficients are ±20%. We were able to observe 2‐chloropropane (i‐C3H7Cl) product for the i‐C3H7 + Cl2 reaction. No products were observed for the other two reactions, and the reasons for this are briefly discussed in the text.  相似文献   

18.
This article critically compares the efficacy of three algorithms, namely Alternating Least‐squares Multi Curve Resolution (ALS‐MCR), Hard Modeling Alternating Least‐squares (HM‐ALS), and classical Hard Modeling Multi Curve Resolution (HM‐MCR) in finding the true values of rate constants associated with a kinetic model. Simulated experiments on the simple system () indicate that soft‐modeling ALS‐MRC methodology, which is subject only to linear constraints, does not ensure that experimental responses are correctly deconvolved, thus preventing further calculations to determine the true rate constants. Inclusion of the kinetic model in the ALS scheme, which gives rise to the HM‐ALS methodology, was found to yield a correct assessment of the rate coefficients but had a large computational cost. Numerical experiments employing a more complex model () were also carried out, mainly to evaluate strategies for performing efficient searches on multidimensional multimodal least‐squares surfaces using HM‐ALS and HM‐MCR. This study again revealed the efficiency and reliability of classical HM‐MCR methods. Results from simulations were corroborated by analysis of data from an experimental study of chromate reduction by hydrogen peroxide; the mechanism of which is similar in complexity to those considered in simulations. The present work suggests that HM‐MCR algorithms implementing a multiminimum search strategy are the method of choice for analyzing two‐dimensional kinetic data.  相似文献   

19.
The kinetics of the ring‐opening polymerization (ROP) of ε‐caprolactone (ε‐CL) initiated by soluble aluminum tri‐sec‐butoxide (Al(OsecBu)3) has been investigated by the differential scanning calorimetry (DSC). The DSC polymerizations were carried out under nonisothermal and isothermal conditions to obtain three‐arms poly(ε‐caprolactone) (PCL). From nonisothermal DSC, the polymerization rate (dα/dt) increased with increasing heating rates. The values of Ea were determined from Kissinger ( kJ mol?1), Friedman (31.0 – 63.0 kJ mol?1), and Starink (64.0 – 71.0 kJ mol?1) methods. From isothermal DSC, the dα/dt and the apparent rate constant (kapp) increased with increasing polymerization temperatures. The ROP of ε‐CL initiated by Al(OsecBu)3 occurred via the coordination insertion mechanism. The number average molecular weight () and percent yield of the synthesized PCL was enhanced by increasing polymerization temperature. The synthesized PCL with of 2.4 × 104 was obtained using a molar ratio of monomer to Al‐O active center ([M]/[Al‐O]) of 400 at 150ºC for 24 h. Al(OsecBu)3 is one of the promising initiator due to its solubility, low transesterification reaction, and high efficiency in ε‐CL polymerization.  相似文献   

20.
The rate constant of the comparably slow bimolecular NCN radical reaction NCN + O2 has been measured for the first time under combustion relevant conditions using the shock tube method. The thermal decomposition of cyanogen azide (NCN3) served as a clean high‐temperature source of NCN radicals. NCN concentration–time profiles have been detected by narrow‐bandwidth laser absorption at cm?1. The experiments behind incident shock waves have been performed with up to 17% O2 in the reaction gas mixture. At such high O2 mole fractions, it was necessary to take O2 relaxation into account that caused a gradual decrease of the temperature during the experiment. Moreover, following fast decomposition of NCN3 and collision‐induced intersystem crossing of the initially formed singlet NCN to its triplet ground state, an unexpected and slow additional formation of triplet NCN has been observed on a 100‐μs timescale. This delayed NCN formation was attributed to a fast recombination of 1NCN with O2 forming a 3NCNOO adduct acting as a reservoir species for NCN. Rate constant data for the reaction NCN + O2 have been measured at temperatures between 1674 and 2308 K. They are best represented by the Arrhenius expression . No pressure dependence has been observed at pressures between 216 and 706 mbar.  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