DFT Study of Trichloroethene Reaction with Permanganate in

Mar 7, 2011 - The mechanism of the environmentally important reaction between permanganate anion and trichloroethene (TCE) has been studied theoretica...
8 downloads 14 Views 1MB Size
ARTICLE pubs.acs.org/est

DFT Study of Trichloroethene Reaction with Permanganate in Aqueous Solution Pawez Adamczyk, Agnieszka Dybala-Defratyka, and Piotr Paneth* Institute of Applied Radiation Chemistry, Faculty of Chemistry, Technical University of Lodz, Zeromskiego 116, 90-924 Lodz, Poland

bS Supporting Information ABSTRACT: The mechanism of the environmentally important reaction between permanganate anion and trichloroethene (TCE) has been studied theoretically using modern DFT functional. It has been shown that IEFPCM/M05-2X/aug-cc-pVDZ theory level yields activation parameters and carbon isotopic fractionation factor in excellent agreement with the experimental data. Obtained results indicate that this reaction proceeds via the 3 þ 2 mechanism with a very early transition state, in which the new C-O bonds are formed only in about 20%. An alternative, stepwise mechanism that involves initial formation of a single new C-O bond and a C-Mn bond, followed by rearrangement to the permanganate-TCE adduct, has been found to be more energetically demanding and in disagreement with the experimental isotopic fractionation.

’ INTRODUCTION Chloroethylenes, such as trichloroethylene (TCE), are among the most important environmental contaminants.1 TCE, which is commonly used as an industrial solvent2 in cleaning machinery, is widely released into subsurface groundwater by a significant number of companies in the EU and the USA,3,4 which makes it the most frequently detected environmental pollutant in water.5 Suspicion of TCE being a human carcinogen6 and also other possible effects on human health such as an impact on the female reproductive functions7 or possible association with congenital heart defects8-12 have made degradation of TCE extremely important. Additionally, in groundwater TCE can undergo an anaerobic enzymatic reductive dechlorination to cis-dichloroethylene (cDCE), which is also considered a possible human carcinogen,6 and vinyl chloride (VC), which is a known human carcinogen.13 Degradation of TCE as well as other chlorinated aliphatic compounds can be carried out in several ways, including chemical and biological approaches, which can be divided into two groups - reductive dehalogenation14-17 and oxidation, which can be further divided into pure chemical oxidation18 and photochemical oxidation.19,20 Chemical oxidation of groundwater subsurface pollutants is usually performed by injections of a chemical oxidant such as permanganate, hydroxyl peroxide, ozone, or persulfate, alone or in combination with other adjuvants.14-16,18-24 Among various treatments of TCE, chemical oxidation with permanganate is most common due to its effectiveness, rapidity, selectivity (in contrast to e.g. hydroxyl peroxide), wide pH range, and relatively low cost.24 The overall mechanism of permanganate oxidation of ethylene and some of its derivatives is well studied25,26 although some controversy regarding the mechanism in the case of TCE remains (see Discussion for details).18 Most frequently it is considered r 2011 American Chemical Society

to be a complex mechanism with the first step, formation of the cyclic hypomanganate ester via the 3 þ 2 electrocyclic addition of permanganate to the carbon-carbon double bond (illustrated by the upper part of Figure 1), being ratedetermining. Environmental importance of the degradation of TCE prompted us to use this reaction as a model for finding a level of theory adequate for predicting reactivity of environmentally important dechlorinations. Herein we report Density Functional Theory (DFT) calculations of the mechanism of the reaction between TCE and permanganate (Figure 1) carried out with the aim of finding a robust, widely applicable method for modeling dehalogenation processes. As the reference points we used experimental data on the Gibbs free energy of activation for this reaction18,27 and carbon isotopic fractionation.28,29 After establishing the adequate theory level we have applied it to learn details of the mechanism of TCE transformation.

’ METHODS AND RESULTS Calculations were carried out using two versions of the Gaussian package (G09, rev.A.0230 and Gaussian03-MN-GFM31) and Jaguar ver.7.6 program of the Schrodinger package32 with various combinations of B3LYP,33-35 CAM-B3LYP,36 MPW1K,37 M05,38 M052X,39,40 M06,40,41 M06-2X,40,41 M06-L,42 M08-HX,43 M08SO,43 LC-wPBE,44-48 LC-TPSS,48,49 MP2,50 and B2PLYP51 levels of theory and IEFPCM52 with UA053 and UFF54 atom radii, PBF,55-57 and SMD58 continuum solvent models of water. The majority of calculations were carried out using the 6-31G,59-61 Received: September 24, 2010 Accepted: February 22, 2011 Revised: February 16, 2011 Published: March 07, 2011 3006

dx.doi.org/10.1021/es103251u | Environ. Sci. Technol. 2011, 45, 3006–3011

Environmental Science & Technology

ARTICLE

Figure 1. Mechanism of the 3 þ 2 (top) and 2 þ 2 (bottom) paths of electrocyclic addition of permanganate anion to TCE.

basis set augmented with a single set of diffuse functions,62 a set of d functions on nonhydrogenic atoms, and a set of p functions on hydrogen atoms;63 6-31þG(d,p), however 6-311þþG (d,p),64-66 def2-TZVPP,67 aug-cc-pVDZ,68-71 and aug-cc-pVTZ72 basis sets were also tested. Geometry optimizations were carried out using default convergence criteria. Vibrational analysis was performed to confirm that the obtained structures are stationary points on the potential energy surface and correspond to either a local minimum (3n-6 real normal modes of vibrations) or a transition state (exactly one imaginary frequency), to calculate free energies of activation (after inclusion of thermal and ZPE contributions), and to calculate isotopic fractionation factors. In all cases full optimizations were carried out. IRC73 calculations confirmed that the transition state structures correspond to the reaction presented in Figure 1. Carbon 13C isotopic fractionation factors, ε (%), were obtained, based on the rule of geometric mean,74 from the kinetic isotope effects calculated for both carbon atoms using complete Bigeleisen equation75 as implemented in the ISOEFF program.76 The obtained results of activation free energies and isotopic fractionations with different combinations of functionals, solvent models, and basis sets are presented in Table S1 in the Supporting Information. Figure 2 illustrates the transition state structure and provides the atom numbering scheme. Details of geometric data of the obtained transition state structures, which include values of bond lengths, the dihedral angle defining the twist of the oxygen atom O8 toward the hydrogen atom of the permanganate anion, and the imaginary frequencies are provided in Tables S2 and S3 of the Supporting Information.

’ DISCUSSION One of the most rapidly developing techniques in studies of environmentally important reactions is compound specific isotopic analysis (CSIA).77-79 The reaction between the permanganate anion and TCE has been studied by carbon CSIA at different reaction conditions. In the first report28 an experiment with a significant headspace was described. This experimental condition allowed for partition of TCE between solution and the gas phase, leading to a too low value of the carbon fractionation factor of -21.4 %. Later studies29 were carried out with no headspace and two different conditions with excess and limited permanganate supply. Similar values in both experiments have been obtained. We have adopted the more precise value

Figure 2. Transition state structure of the 3 þ 2 reaction mechanism.

of -25.1 ( 0.4 % for comparison with the values calculated theoretically. Since no temperature has been indicated we have assumed that the experiments were carried out at room temperature and used it in calculations of the kinetic isotope effects, KIE, that are related to isotopic fractionation factors, ε, by eq 1.80 ε ¼ ð1=KIE - 1Þ 3 1000

ð1Þ

CSIA analysis averages isotopic composition of atoms of a particular element present in the studied compound and since TCE contains two carbon atoms, their averaged theoretical ε are given in Table S1 (see the Supporting Information). The higher the position in this table of the results obtained at different theory levels, the smaller their absolute deviation from the experimental data. Out of the studied theory levels only those reported in the first three rows are within the reported confidence interval. It should be kept in mind though that theoretical values calculated from the Bigeleisen equations are obtained within the conventional transition state theory with harmonic normal modes and rigid rotor approximations. Typically, frequencies are scaled to account for anharmonicity. In terms of isotope effects, frequency scaling leads to lowering of their theoretical values; however, tunneling correction, which was shown to be important even for heavy atoms,81 tends to increase their values and these two effects nearly cancel out. For example, the value of -24.93% obtained from the unscaled frequencies at the PCM/B3LYP/6-31þG(d, p) level changes to -24.45% when the scaling factor82 for this theory level is used but reaches -24.96 % when tunneling is also included via the one-dimensional Wigner correction.83 We consider theoretical results that are within 1% (2SD) of the experimental value as acceptable; such results were obtained for theory levels reported in the first six rows of Table S1. Because of the importance of the studied reaction in bioremediation, its energetics has also been studied in detail. The values of the enthalpy and entropy of activation reported in 3007

dx.doi.org/10.1021/es103251u |Environ. Sci. Technol. 2011, 45, 3006–3011

Environmental Science & Technology literature18 allow to estimate the Gibbs free energy of activation to be equal 10.3 kcal/mol. Since no estimation of errors has been provided, we consider the value of 0.9 kcal/mol reported in earlier studies of this reaction energetics84 as a fair estimation of the standard deviation. Only six values obtained theoretically are within the experimental uncertainty ((2SD). It should be noted, however, that several factors hidden in technical details of the calculations may influence the results. The energy in solution may be calculated with the solute’s electrostatic potential selfconsistent with the solvent reaction field85 or not. Furthermore, Wheeler and Houk showed recently86 that meta-GGA functionals are prone to grid errors. For example, changing the grid for DFT calculations from fine (default) to ultrafine increases ΔG‡ by about 0.2-0.3 kcal/mol and lowers ε by about 1%. Out of the six theory levels that yielded acceptable carbon isotopic fractionation factors, only three (PCM/M05-2X/augcc-pVDZ, PCM/M05-2X/aug-cc-pVTZ, and PCM/M08-SO/631þG(d,p)) predicted correctly the free energy of activation. Out of these, wide availability of the M05-2X functional in programs for quantum-mechanical calculations compared to M08-SO, and the cost of calculations, we have selected the PCM/M05-2X/aug-cc-pVDZ level for interpreting the mechanism of the reaction between TCE and permanganate. Before we move to detailed analysis of the mechanism, in the following paragraphs we comment on the performance of other theory levels tested in this study. Apart from several popular DFT methods we have tested three post-Hartree-Fock levels; CCSD, MP2, and B2PLYP. Both MP-based methods, MP2 and B2PLYP, gave wrong imaginary frequencies of 1346i and 1063i cm-1, respectively (see Table S2 in the Supporting Information), heavily overestimating the contribution of the hydrogen atom to this vibration. Poor performance of the MP2 level for reactions considered herein has been evidenced earlier in the literature.26 Furthermore, B2PLYP (and M05) predicted formation of the two C-O bonds not to be concerted. At the B2PLYP level the formation of the C1-O7 bond is significantly more advanced in the transition state than that of the C2-O8 bond, the difference in length is nearly 0.24 Å. The corresponding difference at the M05 level of theory approaches 0.12 Å also indicating lack of synchronicity in the formation of these new bonds. Other theory levels give the difference between these bonds not exceeding 0.04 Å. CCSD, on the other hand, heavily underestimates isotopic fractionation factor (see Table S1). It should be kept in mind that these theory levels converge slowly with basis sets and those used here might be too small. The use of large basis sets (CBS-type) with these levels, on the other hand, makes their applicability limited due to the computational requirements.87 Good performance of the PBF solvent model in predicting activation parameters has been noticed. Results of free energies of activation are collected in Table S1. In several cases these values are substantially larger than the experimental value. This is the case when a DFT method yielded a hydrogen-bonded complex of the reactants, as illustrated by the structure presented on the right in Figure S1 in the Supporting Information, rather than the proximity complex that leads to reaction. B3LYP, CAM-B3LYP, MPW1K, LC-WPBE, and LCTPSS functionals optimized structures of the reactants were hydrogen-bonded structures. The same result was obtained with the M05-2X functional when calculations did not involve any solvent model. We have evaluated that the hydrogen bonded complex is stabilized over the proximity complex by about 3 kcal/ mol in the case of B3LYP calculations for which both types of

ARTICLE

complexes were obtained. Typical proximity complex obtained in the remaining calculations is illustrated by the structure on the left side of Figure S1. Both structures in this figure are shown with the plane containing TCE atoms perpendicular to the plane of the page. It is interesting to note that M05-2X outperforms newer functionals from the same family and also M05 and M06 methods that are recommended for systems containing transition metals. We have, however, arrived at a similar conclusion earlier when studying spectral properties of iridium complexes.88 We now move our discussion to the mechanism of the reaction between TCE and permanganate. The chemical nature of the reaction is illustrated by Figure 1 as it emerges from the calculations at the selected PCM/M05-2X/aug-cc-pVDZ theory level. The reaction proceeds with a moderate activation barrier of less than 10 kcal/mol and a very large heat of reaction, with the adduct being more stable than the reactants by more than 100 kcal/mol. This huge exothermicity causes that the transition state is very early; formation of the two C-O bonds is practically concerted with the bonds lengths equal to 2.17 and 2.15 Å that correspond to the Wiberg bond orders89 of the forming C-O bonds of 0.22 and 0.24 for C1-O7 and C2-O8, respectively. At the same time partial atomic charges obtained from the NPA analysis90 indicate that there is practically no charge transfer from the permanganate anion to TCE in the transition state. The dihedral angle, Φ, defined by atoms Mn-O7-C1-C reflects asymmetry of the transition state induced by the presence of one hydrogen atom in the TCE molecule. This angle changes from the value of 42.4 degrees in the proximity complex to 19.5 degrees in the transition state and -32.6 degrees in the product. It has been postulated in the literature18 that the reaction between TCE and permanganate proceeds via the 2 þ 2 attack as illustrated in the bottom part of Figure 1 based on the assumption that this arrangement better explains interactions of electron-rich permanganate with carbon-carbon double bond91 and the predictions of a lower barrier92 than that of the 3 þ 2 reaction. We have studied this reaction path on the selected theory level and have found that it in fact corresponds to a stepwise formation of the C-O bonds. In agreement with literature, the transition state corresponding to rearrangement to the TCE-permanganate adduct is lower in energy than the one obtained for the concerted 3 þ 2 mechanism by about 11 kcal/ mol. This transition state structure is illustrated by the structure TS2 on the right side of Figure S2 in the Supporting Information. The imaginary frequency of 243i cm-1 that characterizes this transition state corresponds to the formation of the C2-O8 bond. Further calculation using IRC73 protocol disclosed, however, that this step is preceded by an initial formation of the C1O7 bond and that the transition state TS1 of this step (given on the left side of Figure S2) corresponds to the free energy of activation of 27.2 kcal/mol, much higher than experimentally observed (see Figure 3). The imaginary frequency of 774i cm-1 corresponds to the formation of the C-O bond. Calculated carbon isotope fractionation factors for both these transition states are equal to -16.9 and -21.9, respectively, indicating that this mechanism is not supported by experimental observations. In summary, very good results for both activation parameters and isotopic fractionation factors are simultaneously obtained with the M05-2X and M08-SO functionals with DZ-quality basis sets and the PCM continuum solvent model. Since the latter is not widely available, we have used the former in studies of the reaction of permanganate with TCE. Our results support recent communications from other laboratories that dispersion 3008

dx.doi.org/10.1021/es103251u |Environ. Sci. Technol. 2011, 45, 3006–3011

Environmental Science & Technology

Figure 3. Energy diagram of the TCE-permanganate reaction paths.

corrected Truhlar’s functional M05-2X yields good results in systems containing an alkene coordinated to a transition metal93 and in E2 or SN2 reactions of X- þ CH3CH2X (X = F, Cl).94 This theory level can thus be used for studying a whole range of environmental transformation reactions. Results obtained at the PCM/M05-2X/aug-cc-pVDZ level of theory for the 3 þ 2 reaction between permanganate anion and TCE indicate that this reaction proceeds via a very early transition state in which the new C-O bonds are formed only in about 20%. The alternative mechanism involving stepwise formation of the two C-O bonds is not supported by the experimental results; calculated overall barrier is much higher and predicted carbon isotopic fractionation is much smaller than observed.

’ ASSOCIATED CONTENT

bS

Supporting Information. Tables reporting activation free energies (kcal/mol) and carbon isotopic fractionations (%), key geometrical parameters of the calculated transition state structures of the 3 þ 2 reaction at all studied levels of theory, structures of proximity complexes and transitions states of the 2 þ 2 reaction, and the Cartesian coordinates for transition states for both mechanisms. This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*Phone: þ48 42 631 3199. E-mail: [email protected].

’ ACKNOWLEDGMENT This work was supported by the EU grant isoSoil, FP7-212781 (2009-2012) and the grant from the Polish Ministry of Science and Higher Education 1130/7.PRUE/2009/7. Access to supercomputing facilities at ICM, PCCS, and Cyfronet (Poland) and MSI (U.S.A.) is gratefully acknowledged. ’ REFERENCES (1) ATSDR (Agency for Toxic Substances and Disease Registry) 2007 CERCLA Priority List of Hazardous Substances, U.S. Department of Health and Human Services, Agency for Toxic Substances and

ARTICLE

Disease Registry, Division of Toxicology and Environmental Medicine, Atlanta, GA, in cooperation with U.S. Environmental Protection Agency, http://www.atsdr.cdc.gov/cercla/07list.html (accessed January 15, 2008). (2) Yamazaki-Nishida, S.; Cervera-March, S.; Nagano, K.; Anderson, M. A.; Hori, K. An Experimental and Theoretical Study of the Reaction Mechanism of the Photoassisted Catalytic Degradation of Trichloroethylene in the Gas Phase. J. Phys. Chem. 1995, 99, 15814–15821. (3) Dyksen, J. E.; Hess, A. F. Alternatives for controlling organics in groundwater supplies. J. Am. Water Works Assoc. 1982, 74, 394–404. (4) NRC. Alternatives for Groundwater Cleanup; National Academy Press: Washington, DC, 1994. (5) Pankow, A.; Cherry, J. A. Dense chlorinated solvents and other DNAPLs in groundwater; Waterloo Press: Waterloo, Canada, 1996. (6) IARC. Trichloroethylene; International Agency for Research on Cancer: Ottawa, Canada, 1995. (7) Xu, F.; Papanayotou, I.; Putt, D. A.; Wang, J.; Lash, L. H. Role of mitochondrial dysfunction in cellular responses to S-(1,2-dichlorovinyl)-Lcysteine in primary cultures of human proximal tubular cells. Biochem. Pharmacol. 2008, 76, 552–567. (8) Goldberg, S. J.; Lebowitz, M. D.; Graver, E. J.; Hicks, S. An association of human congenital cardiac malformations and drinking water contaminants. J. Am. Coll. Cardiol. 1990, 16, 155–164. (9) Dawson, B. V.; Johnson, P. D.; Goldberg, S. J.; Ulreich, J. B. Cardiac teratogenesis of halogenated hydrocarbon-contaminated drinking water. J. Am. Coll. Cardiol. 1993, 21, 1466–1472. (10) Boyer, A. S.; Finch, W. T.; Runyan, R. B. Trichloroethylene inhibits development of embryonic heart valve precursors in vitro. Toxicol. Sci. 2000, 53, 109–117. (11) Collier, J. M.; Selmin, O.; Johnson, P. D.; Runyan, R. B. Trichloroethylene effects on gene expression during cardiac development. Birth Defects Res., Part A 2003, 67, 488–495. (12) Drake, V. J.; Koprowski, S. L.; Lough, J.; Hu, N.; Smith, S. M. Trichloroethylene exposure during cardiac valvuloseptal morphogenesis alters cushion formation and cardiac hemodynamics in the avian embryo. Environ. Health Perspect. 2006, 114, 842–847. (13) International Agency for Research on Cancer (IARC); Vinyl chloride; IARC Monogr. Eval. Carcinog. Risks Hum., Suppl. 7, 1987, 373. (14) Whittaker, M.; Monroe, D.; Oh, D. J.; Anderson, S.; 2008, Trichloroethylene pathway map. University of Minnesota biocatalysis/ biodegradation database: http://umbbd.msi.umn.edu/tce/tce_map. html (accessed January 15, 2008). (15) Hara, J.; Ito, H.; Suto, K.; Inoue, C.; Chida, T. Kinetics of trichloroethene dechlorination with iron powder. Water Res. 2005, 39, 1165–1173. (16) Lim, D.-H.; Lastoskie, Ch. M. Density Functional Theory Studies on the Relative Reactivity of Chloroethenes on Zerovalent Iron. Environ. Sci. Technol. 2009, 43, 5443–5448. (17) Bylaska, E. J.; Dupuis, M.; Tratnyek, P. G. One-ElectronTransfer Reactions of Polychlorinated Ethylenes: Concerted and Stepwise Cleavages. J. Phys. Chem. A 2008, 112, 3712–3721. (18) Yan, Y. E.; Schwartz, F. W. Kinetics and Mechanisms for TCE Oxidation by Permanganate. Environ. Sci. Technol. 2000, 34, 2535–2541. (19) Hironen, A.; Tuhkanen, T.; Kalliokoski, P. Formation of chlorinated acetic acids during UV/H2O2-oxidation of ground water contaminated with chlorinated ethylenes. Chemosphere 1996, 32 1091–1102. (20) Li, K.; Stefan, M. I.; Crittenden, J. C. UV Photolysis of Trichloroethylene: Product Study and Kinetic Modeling. Environ. Sci. Technol. 2004, 38, 6685–6693. (21) Pham, T. H.; Kitsuneduka, M.; Hara, J.; Suto, K.; Inoue, C. Trichloroethylene Transformation by Natural Mineral Pyrite: The Deciding Role of Oxygen. Environ. Sci. Technol. 2008, 42, 7470–7475. (22) Pham, T. H.; Kitsuneduka, M.; Hara, J.; Suto, K.; Inoue, C. Trichloroethylene Transformation in Aerobic Pyrite Suspension: Pathways and Kinetic Modeling. Environ. Sci. Technol. 2009, 43, 6744–6749. (23) Waldemer, R. H.; Tratnyek, P. G.; Johnson, R. L.; Nurmi, J. T. Oxidation of Chlorinated Ethenes by Heat-Activated Persulfate: Kinetics and Products. Environ. Sci. Technol. 2007, 41, 1010–1015. 3009

dx.doi.org/10.1021/es103251u |Environ. Sci. Technol. 2011, 45, 3006–3011

Environmental Science & Technology (24) Technical and Regulatory Guidance for In Situ Chemical Oxidation of Contaminated Soil and Groundwater, 2nd ed.; Interstate Technology and Regulatory Council (ITRC): Technical/Regulatory Guideline, 2005. (25) Houk, K. N.; Strassner, T. Establishing the (3 þ 2) mechanism for the permanganate oxidation of alkenes by theory and kinetic isotope effects. J. Org. Chem. 1999, 64, 800–802. (26) Wiberg, K. B.; Wang, Y.-G.; Sklenak, S.; Deutsch, C.; Trucks, G. Permanganate Oxidation of Alkenes. Substituent and Solvent Effects. Difficulties with MP2 Calculations. J. Am. Chem. Soc. 2006, 128 11537–11544. (27) Huang, K.-Ch; Hoag, G. E.; Chheda, P.; Woody, B. A.; Dobbs, G. M. Kinetics and mechanism of oxidation of tetrachloroethylene with permanganate. Chemosphere 2002, 46, 815–25. (28) Poulson, S. R.; Naraoka, H. Carbon Isotope Fractionation during Permanganate Oxidation of Chlorinated Ethylenes (cDCE, TCE, PCE). Environ. Sci. Technol. 2002, 36, 3270–3274. (29) Hunkeler, D.; Aravena, R.; Parker, B. L.; Cherry, J. A.; Diao, X. Monitoring Oxidation of Chlorinated Ethenes by Permanganate in Groundwater Using Stable Isotopes: Laboratory and Field Studies. Environ. Sci. Technol. 2003, 37, 798–804. (30) Frisch, M. J. et al. . Gaussian 09, Revision A.02; Gaussian, Inc.: Wallingford, CT, 2009. (31) Zhao, Y.; Truhlar, D. G.;MN-GFM: Minnesota Gaussian Functional Module, Version 3.0, University of Minnesota, Minneapolis, MN, 2007. (32) Jaguar, version 7.6, Schr€odinger, LLC: New York, NY, 2007. (33) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. (34) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. (35) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Chem. Phys. 1994, 98, 11623–11627. (36) Yanai, T.; Tew, D. P.; Handy, N. C. A new hybrid exchangecorrelation functional using the Coulomb-attenuating method (CAMB3LYP). Chem. Phys. Lett. 2004, 393, 51–57. (37) Lynch, B. J.; Fast, P. L.; Harris, M.; Truhlar, D. G. Adiabatic Connection for Kinetics. J. Phys. Chem. A 2000, 104, 4811–4815. (38) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Exchange-correlation functional with broad accuracy for metallic and nonmetallic compounds, kinetics, and noncovalent interactions. J. Chem. Phys. 2005, 123 161103–161106. (39) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364–382. (40) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. (41) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. (42) Zhao, Y.; Truhlar, D. G. A new local density functional for maingroup thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101–194119. (43) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Exploring the Limit of Accuracy of the Global Hybrid Meta Density Functional for MainGroup Thermochemistry, Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2008, 4, 1849–1868. (44) Tawada, Y.; Tsuneda, T.; Yanagisawa, S.; Yanai, T.; Hirao, K. A long-range-corrected time-dependent density functional theory. J. Chem. Phys. 2004, 120, 8425–8433. (45) Vydrov, O. A.; Scuseria, G. E. Assessment of a long-range corrected hybrid functional. J. Chem. Phys. 2006, 125, 234109–234118.

ARTICLE

(46) Vydrov, O. A.; Heyd, J.; Krukau, A.; Scuseria, G. E. Importance of short-range versus long-range Hartree-Fock exchange for the performance of hybrid density functionals. J. Chem. Phys. 2006, 125, 074106–074101. (47) Vydrov, O. A.; Scuseria, G. E.; Perdew, J. P. Tests of functionals for systems with fractional electron number. J. Chem. Phys. 2007, 126, 154109–154118. (48) Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K. A long-range correction scheme for generalized-gradient-approximation exchange functionals. J. Chem. Phys. 2001, 115, 3540–3544. (49) Tao, J. M.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401–146404. (50) Møller, C.; Plesset, M. S. Note on the approximation treatment for many-electron systems. Phys. Rev. 1934, 46, 618–622. (51) Grimme, S. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 2006, 124, 34108–34123. (52) Miertus, S.; Scrocco, E.; Tomasi Electrostatic interaction of a solute with a continuum. A direct utilization of ab initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55 117–129. (53) Barone, V.; Cossi, M.; Tomasi, J. A new definition of cavities for the computation of solvation free energies by the polarizable continuum model. J. Chem. Phys. 1997, 107, 3210–3221. (54) Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard, W. A., III; Skiff, W. M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114 10024–10035. (55) Tannor, D. J.; Marten, B.; Murphy, R. B.; Friesner, R. A.; Sitkoff, D.; Nicholls, A.; Ringnalda, M. N.; Goddard, W. A.,, III; Honig, B. Accurate First Principles Calculation of Molecular Charge Distributions and Solvation Energies from Ab Initio Quantum Mechanics and Continuum Dielectric Theory. J. Am. Chem. Soc. 1994, 116, 11875–11882. (56) Marten, B.; Kim, K.; Cortis, C.; Friesner, R. A.; Murphy, R. B.; Ringnalda, M. N.; Sitkoff, D.; Honig, B. A New Model For Calculation of Solvation Free Energies: Correction of Self-Consistent Reaction Field Continuum Dielectric Theory for Short Range Hydrogen-Bonding Effects. J. Phys. Chem. 1996, 100, 11775–11788. (57) Friedrichs, M.; Zhou, R. H.; Edinger, S. R.; Friesner, R. A. Poisson-Boltzmann Analytical Gradients for Molecular Modeling Calculations. J. Phys. Chem. B 1999, 103, 3057–3061. (58) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. (59) Hariharan, P. C.; Pople, J. A. Influence of Polarization Functions on MO Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213–222. (60) Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-consistent molecular-orbital methods. IX. Extended Gaussian-type basis for molecularorbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724–728. (61) Francl, M. M.; Pietro, W. J.; Hehre, W. J.; Binkley, J. S.; Gordon, M. S.; DeFrees, D. J.; Pople, J. A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements J. Chem. Phys. 1982, 77, 3654–3665. (62) Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. v R. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21 þ G basis set for first-row elements, lithium to fluorine. J. Comput. Chem. 1983, 4, 294–301. (63) Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-consistent molecular orbital methods. 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. (64) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Selfconsistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. (65) McLean, A. D.; Chandler, G. S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11-18. J. Chem. Phys. 1980, 72, 5639–5648. 3010

dx.doi.org/10.1021/es103251u |Environ. Sci. Technol. 2011, 45, 3006–3011

Environmental Science & Technology (66) Raghavachari, K.; Trucks, G. W. Highly correlated systems. Excitation energies of first row transition metals Sc-Cu. J. Chem. Phys. 1989, 91, 1062–1065. (67) F. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Phys. 2005, 7, 3297–3305. (68) Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. (69) Woon, D. E.; Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. (70) Peterson, K. A.; Woon, D. E.; Dunning, T. H., Jr. Benchmark calculations with correlated molecular wave functions. IV. The classical barrier height of the H þ H2 f H2 þ H reaction. J. Chem. Phys. 1994, 100, 7410–7415. (71) Wilson, A. K.; van Mourik, T.; Dunning, T. H., Jr. Gaussian basis sets for use in correlated molecular calculations. VI. Sextuple-zeta correlation-consistent sets for boron through neon. J. Mol. Struct. (Theochem) 1996, 388, 339–349. (72) Balabanov, N. B.; Peterson, K. A. Systematically convergent basis sets for transition metals. I. All-electron correlation consistent basis sets for the 3d elements Sc-Zn. J. Chem. Phys. 2005, 123 064107–064121. (73) Fukui, K. Formulation of the reaction coordinate. J. Phys. Chem. 1970, 74, 4161–4163. (74) Bigeleisen, J. Statistical mechanics of isotope systems with small quantum corrections. General considerations and the rule of the geometric mean. J. Chem. Phys. 1955, 23, 2264–2267. (75) Bigeleisen, J. The relative reaction velocities of isotopic molecules. J. Chem. Phys. 1949, 17, 675–678. (76) Anisimov, V.; Paneth, P. ISOEFF98. A program for studies of isotope effects using Hssian modifications. J. Math. Chem. 1999, 26, 75–86. (77) Koester, C. J.; Simonich, S. L.; Esser, B. K. Environmental Analysis. Anal. Chem. 2003, 75, 2813–2829. (78) Elsner, M.; Zwank, L.; Hunkeler, D.; Schwarzenbach, R. P. A New Concept Linking Observable Stable Isotope Fractionation to Transformation Pathways of Organic Pollutants. Environ. Sci. Technol. 2005, 39, 6896–6916. (79) Dybala-Defratyka, A.; Szatkowski, L.; Kaminski, R.; Wujec, M.; Siwek, A.; Paneth, P. Kinetic Isotope Effects on Dehalogenations at an Aromatic Carbon. Environ. Sci. Technol. 2008, 42, 7744–7750. (80) Hofstetter, T. H.; Schwarzenbach, R. P.; Bernasconi, S. M. Assessing Transformation Processes of Organic Compounds Using Stable Isotope Fractionation. Environ. Sci. Technol. 2008, 42, 7737–7743. (81) Meyer, M. P.; DelMonte, A. J.; Singleton, D. A. Reinvestigation of the Isotope Effects for the Claisen and Aromatic Claisen Rearrangements: The Nature of the Claisen Transition States. J. Am. Chem. Soc. 1999, 121, 10865–10874. (82) Scott, A. P.; Radom, L. Harmonic Vibrational Frequencies: An Evaluation of Hartree-Fock, Moller-Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors. J. Phys. Chem. 1996, 100, 16502–16513. (83) Wigner, E. Crossing of potential thresholds in chemical reaction. Z. Phys. Chem. B 1932, 19, 203–216. (84) Huang, K.-C.; Hoag, G. E.; Chheda, P.; Woody, B. A.; Dobbs, G. M. Kinetics and mechanism of oxidation of tetrachloroethylene wit permanganate. Chemosphere 2002, 46, 815–825. (85) Improta, R.; Barone, V.; Scalmani, G.; Frisch, M. J. A statespecific polarizable continuum model time dependent density functional theory method for excited state calculations in solution. J. Chem. Phys. 2006, 125, 054103–054112. (86) Wheeler, S. E.; Houk, K. N. Integration Grid Errors for MetaGGA-Predicted Reaction Energies: Origin of Grid Errors for the M06 Suite of Functionals. J. Chem. Theory Comput. 2010, 6, 395–404. (87) Mladek, A.; Sponer, J. E.; Jurecka, P.; Banas, P.; Otyepka, M.; Svozil, D.; Sponer, J. Conformational Energies of DNA Sugar-Phosphate Backbone: Reference QM Calculations and a Comparison with Density

ARTICLE

Functional Theory and Molecular Mechanics. J. Chem. Theory Comput. 2010, ASAP. (88) Swiderek, K.; Paneth, P. Modeling excitation properties of iridium complexes. J. Phys. Org. Chem. 2009, 22, 845–856. (89) Wiberg, K. A. Application of the Pople-Santry-Segal complete neglect of differential overlap method to the cyclopropyl-carbinyl and cyclobutyl cation and to bicyclobutane. Tetrahedron 1968, 24, 1083–1096. (90) Martin, F.; Zipse, H. Charge distribution in the water molecule a comparison of methods. J. Comput. Chem. 2005, 26, 97–105. (91) Sharpless, K. B.; Teranishi, A. Y.; Backvall, J. E. Chromyl chloride oxidations of olefins. Possible role of organometallic intermediates in the oxidations of olefins by oxo transition metal species. J. Am. Chem. Soc. 1977, 99, 3120–3128. (92) Rappe, A. K.; Goddard, W. A.,, III. Olefin metathesis - a mechanistic study of high-valent Group VI catalysts. J. Am. Chem. Soc. 1982, 104, 448–456. (93) Flener-Lovitt, C.; Woon, D. E.; Dunning, T. H., Jr.; Girolami, G. S. A DFT and ab Initio Benchmarking Study of Metal-Alkane Interactions and the Activation of Carbon-Hydrogen Bonds. J. Phys. Chem. A 2010, 114, 1843–1851. (94) Bento, A. P.; Sola, M.; Bickelhaupt, F. M. E2 and SN2 Reactions of X- þ CH3CH2X (X = F, Cl); an ab Initio and DFT Benchmark Study. J. Chem. Theory Comput. 2008, 4, 929–940.

3011

dx.doi.org/10.1021/es103251u |Environ. Sci. Technol. 2011, 45, 3006–3011