Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 11 de 11
Filter
Add more filters










Publication year range
1.
J Phys Chem A ; 120(37): 7320-6, 2016 Sep 22.
Article in English | MEDLINE | ID: mdl-27556743

ABSTRACT

FTIR smog chamber techniques were used to study the kinetics of the gas-phase reactions of Cl atoms, OD radicals, and O3 with the five-membered ring-structured compounds tetrahydrofuran (C4H8O, THF), 2-methyltetrahydrofuran (CH3C4H7O, 2-MTHF), 2,5-dimethyltetrahydrofuran ((CH3)2C4H5O, 2,5-DMTHF), and furan (C4H4O). The rate coefficients determined using relative rate methods were kTHF+Cl = (1.96 ± 0.24) × 10(-10), kTHF+OD = (1.81 ± 0.27) × 10(-11), kTHF+O3 = (6.41 ± 2.90) × 10(-21), k2-MTHF+Cl = (2.65 ± 0.43) × 10(-10), k2-MTHF+OD = (2.41 ± 0.51) × 10(-11), k2-MTHF+O3 = (1.87 ± 0.82) × 10(-20), k2,5-DMTHF+OD = (4.56 ± 0.68) × 10(-11), k2,5-DMTHF+Cl = (2.84 ± 0.34) × 10(-10), k2,5-DMTHF+O3 = (4.58 ± 2.18), kfuran+Cl = (2.39 ± 0.27) × 10(-10), and kfuran+O3 = (2.60 ± 0.31) × 10(-18) molecules cm(-3) s(-1). Rate coefficients of the reactions with ozone were also determined using the absolute rate method under pseudo-first-order conditions. OD radicals, in place of OH radicals, were produced from CD3ONO to avoid spectral overlap of isopropyl and methyl nitrite with the reactants. The kinetics of OD radical reactions are expected to resemble the kinetics of OH radical reactions, and the rate coefficients of the reactions with OD radicals were used to calculate the atmospheric lifetimes with respect to reactions with OH radicals. The lifetimes of THF, 2-MTHF, and 2,5-DMTHF are approximately 15, 12, and 6 h, respectively.

2.
J Phys Chem A ; 119(11): 2562-7, 2015 Mar 19.
Article in English | MEDLINE | ID: mdl-25654214

ABSTRACT

Acetaldehyde is a hazardous pollutant found in indoor and ambient air. Acetaldehyde photolysis is pressure- and wavelength-dependent with three distinct product channels. In this study, the photolysis rates of CH3CHO, CD3CDO, and CD3CHO are studied in natural tropospheric conditions using long path FTIR spectroscopy, at the European Photoreactor Facility (EUPHORE) in Valencia, Spain. The average relative photolysis rate as an average of four experiments for the fully deuterated isotopologue is j(CH3CHO)/j(CD3CDO) = 1.75 ± 0.04, and as a result of a single experiment j(CH3CHO)/j(CD3CHO) = 1.10 ± 0.10. These results, combined with our previous determination of j(CH3CHO)/j(CH3CDO) = 1.26 ± 0.03, provide mechanistic insight into the photodissociation dynamics of the photoexcited species. Despite the extensive isotopic scrambling in photoexcited acetaldehyde that has recently been reported, the position of the substitution has a clear effect on the relative photolysis rates.

3.
Environ Sci Technol ; 48(15): 8768-76, 2014.
Article in English | MEDLINE | ID: mdl-24955878

ABSTRACT

In this article, gas-phase advanced oxidation, a new method for pollution control building on the photo-oxidation and particle formation chemistry occurring in the atmosphere, is introduced and characterized. The process uses ozone and UV-C light to produce in situ radicals to oxidize pollution, generating particles that are removed by a filter; ozone is removed using a MnO2 honeycomb catalyst. This combination of in situ processes removes a wide range of pollutants with a comparatively low specific energy input. Two proof-of-concept devices were built to test and optimize the process. The laboratory prototype was built of standard ventilation duct and could treat up to 850 m(3)/h. A portable continuous-flow prototype built in an aluminum flight case was able to treat 46 m(3)/h. Removal efficiencies of >95% were observed for propane, cyclohexane, benzene, isoprene, aerosol particle mass, and ozone for concentrations in the range of 0.4-6 ppm and exposure times up to 0.5 min. The laboratory prototype generated a OH(•) concentration derived from propane reaction of (2.5 ± 0.3) × 10(10) cm(-3) at a specific energy input of 3 kJ/m(3), and the portable device generated (4.6 ± 0.4) × 10(9) cm(-3) at 10 kJ/m(3). Based on these results, in situ gas-phase advanced oxidation is a viable control strategy for most volatile organic compounds, specifically those with a OH(•) reaction rate higher than ca. 5 × 10(-13) cm(3)/s. Gas-phase advanced oxidation is able to remove compounds that react with OH and to control ozone and total particulate mass. Secondary pollution including formaldehyde and ultrafine particles might be generated, depending on the composition of the primary pollution.


Subject(s)
Air Pollution/prevention & control , Ozone/isolation & purification , Particulate Matter/isolation & purification , Photolysis , Volatile Organic Compounds/isolation & purification , Air Filters , Atmosphere/chemistry , Benzene/isolation & purification , Butadienes/isolation & purification , Cyclohexanes/isolation & purification , Environmental Pollution , Hazardous Substances/isolation & purification , Hemiterpenes/isolation & purification , Oxidation-Reduction , Pentanes/isolation & purification , Propane/isolation & purification , Ultraviolet Rays
4.
J Phys Chem A ; 115(32): 8906-19, 2011 Aug 18.
Article in English | MEDLINE | ID: mdl-21797203

ABSTRACT

The atmospheric chemistry of two C(4)H(8)O(2) isomers (methyl propionate and ethyl acetate) was investigated. With relative rate techniques in 980 mbar of air at 293 K the following rate constants were determined: k(C(2)H(5)C(O)OCH(3) + Cl) = (1.57 ± 0.23) × 10(-11), k(C(2)H(5)C(O)OCH(3) + OH) = (9.25 ± 1.27) × 10(-13), k(CH(3)C(O)OC(2)H(5) + Cl) = (1.76 ± 0.22) × 10(-11), and k(CH(3)C(O)OC(2)H(5) + OH) = (1.54 ± 0.22) × 10(-12) cm(3) molecule(-1) s(-1). The chlorine atom initiated oxidation of methyl propionate in 930 mbar of N(2)/O(2) diluent (with, and without, NO(x)) gave methyl pyruvate, propionic acid, acetaldehyde, formic acid, and formaldehyde as products. In experiments conducted in N(2) diluent the formation of CH(3)CHClC(O)OCH(3) and CH(3)CCl(2)C(O)OCH(3) was observed. From the observed product yields we conclude that the branching ratios for reaction of chlorine atoms with the CH(3)-, -CH(2)-, and -OCH(3) groups are <49 ± 9%, 42 ± 7%, and >9 ± 2%, respectively. The chlorine atom initiated oxidation of ethyl acetate in N(2)/O(2) diluent gave acetic acid, acetic acid anhydride, acetic formic anhydride, formaldehyde, and, in the presence of NO(x), PAN. From the yield of these products we conclude that at least 41 ± 6% of the reaction of chlorine atoms with ethyl acetate occurs at the -CH(2)- group. The rate constants and branching ratios for reactions of OH radicals with methyl propionate and ethyl acetate were investigated theoretically using transition state theory. The stationary points along the oxidation pathways were optimized at the CCSD(T)/cc-pVTZ//BHandHLYP/aug-cc-pVTZ level of theory. The reaction of OH radicals with ethyl acetate was computed to occur essentially exclusively (∼99%) at the -CH(2)- group. In contrast, both methyl groups and the -CH(2)- group contribute appreciably in the reaction of OH with methyl propionate. Decomposition via the α-ester rearrangement (to give C(2)H(5)C(O)OH and a HCO radical) and reaction with O(2) (to give CH(3)CH(2)C(O)OC(O)H) are competing atmospheric fates of the alkoxy radical CH(3)CH(2)C(O)OCH(2)O. Chemical activation of CH(3)CH(2)C(O)OCH(2)O radicals formed in the reaction of the corresponding peroxy radical with NO favors the α-ester rearrangement.


Subject(s)
Acetates/chemistry , Atmosphere/chemistry , Biofuels , Chlorine/chemistry , Hydroxyl Radical/chemistry , Organic Chemicals/chemistry , Propionates/chemistry , Computer Simulation , Esters/chemistry , Kinetics , Nitric Oxide , Oxidation-Reduction , Oxygen , Spectrophotometry, Infrared
5.
J Phys Chem A ; 113(15): 3498-504, 2009 Apr 16.
Article in English | MEDLINE | ID: mdl-19298072

ABSTRACT

The photolysis rates of HCHO, DCDO, CH(3)CHO, and CH(3)CDO are studied by long-path FTIR spectroscopy in natural tropospheric conditions at the European Photoreactor Facility (EUPHORE) in Valencia, Spain. Average relative photolysis rates j(HCHO)/j(DCDO) = 3.15 +/- 0.08 and j(CH(3)CHO)/j(CH(3)CDO) = 1.26 +/- 0.03 are obtained from three days of experiments for each reaction in the period June 17 to July 7, 2006.

6.
J Phys Chem A ; 113(9): 1731-9, 2009 Mar 05.
Article in English | MEDLINE | ID: mdl-19206225

ABSTRACT

The kinetic isotope effects in the reactions of CHCl(3), CDCl(3), and (13)CHCl(3) with Cl, OH, and OD radicals have been determined in relative rate experiments at 298 +/- 1 K and atmospheric pressure monitored by long path FTIR spectroscopy. The spectra were analyzed using a nonlinear least-squares spectral fitting procedure including line data from the HITRAN database and measured infrared spectra as references. The following relative reaction rates were determined: k(CHCl(3)+Cl)/k(CDCl(3)+Cl) = 3.28 +/- 0.01, k(CHCl(3)+Cl)/k((13)CHCl(3)+Cl) = 1.000 +/- 0.003, k(CHCl(3)+OH)/k(CDCl(3)+OH) = 3.73 +/- 0.02, k(CHCl(3)+OH)/k((13)CHCl(3)+OH) = 1.023 +/- 0.002, k(CHCl(3)+OD)/k(CDCl(3)+OD) = 3.95 +/- 0.03, and k(CHCl(3)+OD)/k((13)CHCl(3)+OD) = 1.032 +/- 0.004. Larger isotope effects in the OH reactions than in the Cl reactions are opposite to the trends for CH(4) and CH(3)Cl reported in the literature. The origin of these differences was investigated using electronic structure calculations performed at the MP2/aug-cc-PVXZ (X = D, T, Q) level of theory and are compared with previously calculated values for the other methane derivatives. The Born-Oppenheimer barrier heights to H abstraction are 12.2 and 17.0 kJ mol(-1) at the CCSD(T)/aug-cc-pVTZ level of theory for OH and Cl, respectively. The reaction rate coefficients of the two elementary vapor phase reactions including the (2)H and (13)C kinetic isotope effects were calculated using improved canonical variational theory with small curvature tunneling (ICVT/SCT) and the results compared with experimental data.

7.
J Chem Phys ; 128(19): 194503, 2008 May 21.
Article in English | MEDLINE | ID: mdl-18500876

ABSTRACT

The performance of the Hartree-Fock method and the three density functionals B3LYP, PBE0, and CAM-B3LYP is compared to results based on the coupled cluster singles and doubles model in predictions of the solvatochromic effects on the vertical n-->pi* and pi-->pi* electronic excitation energies of acrolein. All electronic structure methods employed the same solvent model, which is based on the combined quantum mechanics/molecular mechanics approach together with a dynamical averaging scheme. In addition to the predicted solvatochromic effects, we have also performed spectroscopic UV measurements of acrolein in vapor phase and aqueous solution. The gas-to-aqueous solution shift of the n-->pi* excitation energy is well reproduced by using all density functional methods considered. However, the B3LYP and PBE0 functionals completely fail to describe the pi-->pi* electronic transition in solution, whereas the recent CAM-B3LYP functional performs well also in this case. The pi-->pi* excitation energy of acrolein in water solution is found to be very dependent on intermolecular induction and nonelectrostatic interactions. The computed excitation energies of acrolein in vacuum and solution compare well to experimental data.

8.
J Phys Chem A ; 111(45): 11506-13, 2007 Nov 15.
Article in English | MEDLINE | ID: mdl-17958339

ABSTRACT

UV (240-370 nm) and IR (3200-1500 cm(-1)) absorption cross-sections of HCHO, HCDO, and DCDO in a bath gas of N(2) at atmospheric pressure and 296 K are reported from simultaneous measurements in the two spectral regions. Cross-sections were placed on an absolute scale through quantitative conversion of formaldehyde to CO and HCOOH by titration with Br atoms, also monitored by FTIR. The integrated UV absorption cross-sections of HCHO, HCDO, and DCDO are equal to within the experimental uncertainty.

9.
J Chem Ecol ; 30(7): 1431-43, 2004 Jul.
Article in English | MEDLINE | ID: mdl-15503529

ABSTRACT

We used video recording of the movement pattern of Protaphorura armata (Collembola) to test whether its avoidance of the odor of dead conspecifics extends to related species. P. armata was repelled by the odor of dead individuals of Onychiurus scotarius and Onychiurus circulans, but not by live individuals. Free palmitic, oleic, and linoleic acids were present in extracts of the three repellent species, but only free palmitic acid was detected in extracts of a fourth nonrepellent species, Folsomia candida. Synthetic palmitic acid was attractive to P. armata, linoleic acid was repellent, and oleic acid gave no response. o. scotarius and O. circulans also contained 2,3-dimethoxy-pyrido[2,3-b]pyrazine, known as a defense substance. We discuss the role of free fatty acids in predator avoidance, conspecifics attraction, and food recognition in P. armata.


Subject(s)
Arthropods/physiology , Fatty Acids, Nonesterified/analysis , Pheromones/analysis , Animals , Arthropods/chemistry , Gas Chromatography-Mass Spectrometry , Insect Repellents/analysis , Linoleic Acid/analysis , Locomotion , Odorants , Oleic Acid/analysis , Palmitic Acid/analysis , Time Factors
10.
J Theor Biol ; 231(4): 497-513, 2004 Dec 21.
Article in English | MEDLINE | ID: mdl-15488527

ABSTRACT

We use a linear autoregressive model to describe the movement of a soil-living insect, Protaphorura armata (Collembola). Models of this kind can be viewed as extensions of a random walk, but unlike a correlated random walk, in which the speed and turning angles are independent, our model identifies and expresses the correlations between the turning angles and a variable speed. Our model uses data in x- and y-coordinates rather than in polar coordinates, which is useful for situations in which the resolution of the observations is limited. The movement of the insect was characterized by (i) looping behaviour due to autocorrelation and cross correlation in the velocity process and (ii) occurrence of periods of inactivity, which we describe with a Poisson random effects model. We also introduce obstacles to the environment to add structural heterogeneity to the movement process. We compare aspects such as loop shape, inter-loop time, holding angles at obstacles, net squared displacement, number, and duration of inactive periods between observed and predicted movement. The comparison demonstrates that our approach is relevant as a starting-point to predict behaviourally complex moving, e.g. systematic searching, in a heterogeneous landscape.


Subject(s)
Insecta/physiology , Linear Models , Movement/physiology , Animals , Environment , Models, Biological , Soil
11.
J Theor Biol ; 231(2): 299-306, 2004 Nov 21.
Article in English | MEDLINE | ID: mdl-15380394

ABSTRACT

Analysis of small-scale movement patterns of animals we may help to understand and predict movement at a larger scale, such as dispersal, which is a key parameter in spatial population dynamics. We have chosen to study the movement of a soil-dwelling Collembola, Protaphorura armata, in an experimental system consisting of a clay surface with or without physical obstacles. A combination of video recordings, descriptive statistics, and walking simulations was used to evaluate the movement pattern. Individuals were found to link periods of irregular walk with those of looping in a homogeneous environment as well as in one structured to heterogeneity by physical obstacles. The number of loops varied between 0 and 44 per hour from one individual to another and some individuals preferred to make loops by turning right and others by turning left. P. armata spent less time at the boundary of small obstacles compared to large, presumably because of a lower probability to track the steepness of the curvature as the individual walks along a highly curved surface. Food deprived P. armata had a more winding movement and made more circular loops than those that were well fed. The observed looping behaviour is interpreted in the context of systematic search strategies and compared with similar movement patterns found in other species.


Subject(s)
Insecta/physiology , Animals , Models, Biological , Movement/physiology , Population Dynamics , Soil
SELECTION OF CITATIONS
SEARCH DETAIL
...