Reaction Energetics for the Abstraction Process C2H3 + H2 → C2H4 + H

Sep 25, 2011 - Small hydrocarbons are ubiquitous in the chemistries of combustion processes and planetary atmospheres.1А5 Their reactivity, which can...
0 downloads 0 Views 1MB Size
LETTER pubs.acs.org/JPCL

Reaction Energetics for the Abstraction Process C2H3 + H2 f C2H4 + H Jay Agarwal, Justin M. Turney, and Henry F. Schaefer, III* Center for Computational Quantum Chemistry, University of Georgia, Athens, Georgia 30602, United States

bS Supporting Information ABSTRACT: The fundamentally important combustion reaction of vinyl radical with hydrogen has been studied in the laboratory by at least five experimental groups. Herein, the reaction C2H3 + H2 f C2H4 + H has been examined using focal-point analysis. Molecular energies were determined from extrapolations to the complete basis-set limit using correlation-consistent basis sets (cc-pVTZ, cc-pVQZ, and cc-pV5Z) and coupled-cluster theory with single and double excitations (CCSD), perturbative triples [CCSD(T)], full triples [CCSDT], and perturbative quadruples [CCSDT(Q)]. Reference geometries were optimized at the all-electron CCSD(T)/cc-pCVQZ level. Computed energies were also corrected for relativistic effects and the BornOppenheimer approximation. The activation energy for hydrogen abstraction is predicted to be 9.65 kcal mol1, and the overall reaction is predicted to be exothermic by 5.65 kcal mol1. Natural resonance theory (NRT) analysis was performed to verify the reaction pathway and describe bond-breaking and bond-forming events along the reaction coordinate. SECTION: Molecular Structure, Quantum Chemistry, General Theory

S

mall hydrocarbons are ubiquitous in the chemistries of combustion processes and planetary atmospheres.15 Their reactivity, which can lead to the formation of aromatic compounds or soot, is a central topic of research in both fields. In flames, soot production results from incomplete combustion, which yields fragments that combine to form larger hydrocarbon species. Similarly, many planetary atmospheres contain high concentrations of molecular hydrogen and methane.68 The photolysis of these species, usually from incident UV radiation, results in photofragments, which also combine to produce larger hydrocarbons as in combustion. Soot does not form directly from initial hydrocarbon fragments, but rather as a result of a complex network of reactions that may be modeled with kinetic data. Prior work has elucidated the reaction rates and absorption spectra of many neutral hydrocarbons,913 but the analogous radical and charged species are much more difficult to study. Whereas neutral species can be produced in large quantity, isolated, and observed, radical and charged species are often produced through photolysis, which yields a myriad of highly reactive products. Subsequent isolation of the desired species, usually though mass-selection, significantly reduces the concentration, which makes spectroscopic analysis difficult. The time-independent nature of a theoretical investigation is ideal for the study of these species, and high-accuracy computational methods have been shown to yield subchemical accuracy for the reaction enthalpies of such systems.1417 The vinyl radical is among the most important species in combustion chemistry. Herein we have investigated, using computational methods, a single pathway involving the reaction of the vinyl radical with hydrogen. Although vinyl radical readily reacts with other hydrocarbons (CH3, C2H3, C2H5, etc.) and r 2011 American Chemical Society

diatomics (O2, Cl2, etc.),18 prior work has shown that in an atmosphere of H2 vinyl radical abstracts a hydrogen atom from molecular hydrogen to yield C2H4 and H.1825 Subsequently the two products combine in a propagation step to form ethyl radical, C2H5. Callear and Smith first indirectly measured the kinetic rate for the hydrogen abstraction step in the mid-1980s. They reported a small activation energy of 5.5 kcal mol1 and a rate of 2.6  1017 cm3 molecule 1 s1 at 298 K.19 Since their work, several groups have reported kinetic rates for the hydrogen abstraction process using data from experimental and theoretical investigations. In 1986, Tsang and Hampson reported a database of kinetic rates for combustion processes, which included vinyl radical.20 Two years later, Weissman and Benson reported results from transition-state theory analysis, which estimated the rate at 5.0  1017 cm3 molecule 1 s1 at 298 K.26 In 1995, two groups reported reaction rates for this process. In a theoretical investigation using QCISD(T) and CCSD(T) theories with a 6-311+G(3df,2p) basis for single-point energies, Mebel, Morokuma, and Lin reported a rate of 2.98  1019 cm3 molecule 1 s1 at 298 K. They found the activation barrier for the forward reaction to be 10.4 kcal mol1 and the overall reaction to be exothermic by 7.1 kcal mol1 [using CCSD(T)].21 Then, Fahr, Monks, Steif, and Laufer reported a comparable rate by employing kinetic absorption spectroscopy and gas chromatographic product analysis.22 In 1996, Knyazev, Bencsura, Stoliarov, and Slagle reported the first direct experimental Received: August 18, 2011 Accepted: September 23, 2011 Published: September 25, 2011 2587

dx.doi.org/10.1021/jz201124j | J. Phys. Chem. Lett. 2011, 2, 2587–2592

The Journal of Physical Chemistry Letters

LETTER

Figure 1. Optimized geometries of the reactants, products, and transition state for the abstraction of a hydrogen atom by vinyl radical, C2H3 + H2 f C2H4 + H. Bond lengths are shown in angstroms and bond angles are shown in degrees. Geometries were optimized at the AE-CCSD(T)/cc-pCVQZ level of theory.

measurement of the abstraction process using photoionization mass spectroscopy. They found the rate to be 7.1  1018 cm3 molecule 1 s1 at 298 K.23 Much of this past work is summarized in a review by Laufer and Fahr, which discusses the kinetics for many small hydrocarbon reactions.18 More recently, there have been two theoretical studies concerning the abstraction step. In 2004, Li, L€u, and Wang employed MP4(SDQ) theory with a 6-311++G(d,p) Pople basis set to fit a rate constant using improved canonical variational transitionstate theory with a small-curvature tunneling correction. They reported a rate of 5.22  1017 cm3 molecule 1 s1 at 298 K using an activation barrier of 9.95 kcal mol1 and an exothermicity of 6.12 kcal mol1,24 and in 2006, Tautermann, Wellenzohn, and Clary found a similar rate constant of 2.1  1018 cm3 molecule 1 s1 at 298 K using an activation barrier of 10.34 kcal mol1 and an exothermicity of 4.84 kcal mol1. They employed time-independent scattering theory with a 2D potential energy surface using B3LYP/6-31G* geometries and G3GB scheme for single-point energies.25 The reported reaction rates vary from 1017 to 1020 cm3 molecule 1 s1 and represent relatively small changes in the activation energy or exothermicity. In fact, Knyazev and coworkers note that a reduction of only 1.3 kcal mol1 in the activation barrier obtained by Mebel and coworkers would account for the discrepancy between their rate constants, which differ by one unit of magnitude.23 Given that variances in the reaction energies are of the same magnitude as the energies themselves, particular care must be taken to obtain values with an accuracy of 0, indicating they are after the energetic maximum is reached. Therefore, the transition state resembles the reactants, and the overall process should be exothermic by Hammond’s postulate.30

C. Vibrational Frequencies. Normal vibrational frequencies for the optimized structures are shown in Table 1. Frequencies were computed at the same level of theory as optimization, AECCSD(T)/cc-pCVQZ. Experimental frequencies are also shown for vinyl radical and ethylene. We have included the complete spectral assignment for ground-state vinyl radical obtained by Letendre and coworkers28 in addition to a revised set of frequencies obtained by Nikow and coworkers.29 Similarly, Table 1 includes comparative general harmonic force-field data for ethylene obtained by Duncan and coworkers.13 With regard to vinyl radical, obtaining resolved experimental spectra is difficult owing to the low concentrations in which vinyl radical is produced and the inherent instability of a reactive radical intermediate. Whereas much of the prior work has resolved vibrationally hot bands for vinyl radical,3138 only recently have ground-state frequencies been obtained. The first ground-state assignment for the entire vibrational spectra was obtained by Letendre and coworkers.28 Previously, Sattelmeyer and Schaefer identified a discrepancy in the assignment of the v5 band using equation of motion coupled cluster methods.39 As a follow-up to previous work, and in an effort to obtain a more finely resolved spectra, Nikow and coworkers29 re-examined the assignments of the vibrational modes using time-resolved FTIR. As mentioned, we have also included that data in Table 1. We find good agreement between our computed frequencies and the experimentally obtained spectra but note that we have reported harmonic vibrational frequencies that cannot be directly compared with the anharmonic frequencies listed without a scaling factor. The general harmonic force-field frequencies for ethylene from Duncan and coworkers13 may be directly compared with 2589

dx.doi.org/10.1021/jz201124j |J. Phys. Chem. Lett. 2011, 2, 2587–2592

The Journal of Physical Chemistry Letters

LETTER

Table 2. Valence Focal Point Analysis of the Classical Abstraction Barrier for C2H3 + H2 f TS1a ΔEeHF

+δMP2

+δCCSD

+δCCSD(T)

cc-pVDZ

+ 20.38

 12.78

+ 1.78

 0.98

cc-pVTZ

+ 21.62

 13.69

+ 2.10

 1.28

cc-pVQZ

+ 21.72

13.99

+2.18

cc-pV5Z

+ 21.75

14.08

CBS LIMIT

[+21.76]

[14.18]

basis set

+δCCSDT

+δCCSDT(Q)

NET

+ 0.02

 0.06

[+8.36]

+ 0.06

 0.08

[+8.72]

 1.35

[+ 0.06]

[0.08]

[+8.53]

+2.21

 1.38

[+0.06]

[0.08]

[+8.48]

[+2.25]

[1.40]

[+0.06]

[0.08]

[+8.40]

ΔEa(final) = ΔEa[CBS CCSDT(Q)] + Δcore[CCSD/cc-pCVQZ] + ΔZPVE[CCSD(T)/cc-pCVQZ] + ΔDBOC[HF/cc-pVTZ] + Δrel[CCSD/cc-pCVTZ] ΔEa(final) = 8.40  0.01 + 0.83 + 0.19 + 0.24 = 9.65 kcal mol1 a

Energies shown in kilocalories per mole. Delta (δ) denotes the change in relative energy (ΔEe) with respect to the preceding level of theory.

Table 3. Valence Focal Point Analysis of the Classical Exothermicity for C2H3 + H2 f C2H4 + Ha basis set

ΔEeHF

+δ MP2

+ δCCSD

+δCCSD(T)

+δCCSDT

+δCCSDT(Q)

NET

cc-pVDZ

12.92

1.88

+ 5.44

0.40

+0.15

0.02

[9.62]

cc-pVTZ

11.79

1.63

+5.52

0.60

+0.19

0.00

[8.31]

cc-pVQZ

11.70

1.63

+ 5.47

0.63

[+0.19]

[0.00]

[8.30]

cc-pV5Z

11.70

1.59

+ 5.44

0.64

[+0.19]

[0.00]

[8.31]

[11.71]

[1.56]

[+5.40]

[0.64]

[+0.19]

[0.00]

[8.32]

CBS LIMIT

ΔE0(final) = ΔE0[CBS CCSDT(Q)] + Δcore[CCSD/cc-pCVQZ] + ΔZPVE[CCSD(T)/cc-pCVQZ] + ΔDBOC[HF/cc-pVTZ] + Δrel [CCSD/cc-pCVTZ] ΔE0(final) = 8.32  0.15 + 2.72 + 0.15 + 0.26 = 5.65 kcal mol1 a

Energies shown in kilocalories per mole. Delta (δ) denotes the change in relative energy (ΔEe) with respect to the preceding level of theory.

our theoretical results. Here we find good agreement between our results and those obtained by experiment, with our computed frequencies deviating by at most 20 cm1 from those obtained by Duncan and coworkers and on average by only 8 cm1. Our predicted ZPVE (31.96 kcal mol1) is also within 0.1 kcal mol1 of their experimental value (31.91 kcal mol1). Additionally, we predict a ZPVE of 22.95 kcal mol1 for C2H3, 30.07 kcal mol1 for C2H4, and 6.29 kcal mol1 for H2. D. Energetics. For the optimized geometries, we have computed the requisite single-point energies for focal point extrapolation to the CBS limit, as described in the Theoretical Methods section. The focal point analyses are listed in Tables 2 and 3 for the abstraction barrier and exothermicity, respectively. Both extrapolations suggest convergence to within 0.1 kcal mol1. We have also included four corrections to the computed energy, as mentioned, and these are reported in Tables 2 and 3. After corrections, we predict the activation barrier for abstraction to be 9.65 kcal mol1 and the overall reaction to be exothermic by 5.65 kcal mol1. Mebel and coworkers,21 Li and coworkers,24 and Tautermann and coworkers25 report the abstraction barrier to be 10.4, 9.95, and 10.34, respectively. Similarly, they report the reaction to be exothermic by 7.1, 6.12, and 4.84 kcal mol1, respectively. These previous results contain a correction for the zero-point vibrational energy and therefore may be directly compared with our results. We find our abstraction barrier to be comparable to but lower than the previously reported values. Aligned with our initial NRT analysis and prior work, we also find the overall reaction to be exothermic. Our value of 5.65 kcal mol1, however, is much lower than that of Mebel and coworkers, roughly 0.5 kcal mol1 lower than the value obtained by Li and coworkers and ∼0.8 kcal mol1 higher than the value obtained by Tautermann and coworkers. Given the magnitude of the energy values, small

differences yield large variations in the resulting kinetics, emphasizing our choice of a such a rigorous treatment of the reaction energies. We estimate our predictions to be accurate to within 0.1 kcal mol1.

’ THEORETICAL METHODS Reference geometries were optimized by all-electron (AE) coupled cluster theory incorporating single, double, and perturbative triple excitations [CCSD(T)]4045 using a corevalence correlation-consistent quadruple-ζ basis-set (cc-pCVQZ).46 The large corevalence cc-pCVQZ basis-set provides 288 contracted Gaussian functions for ethylene and 258 for vinyl radical. This AE-CCSD(T)/cc-pCVQZ level of theory has been shown to be highly accurate for equilibrium geometric parameters of singlereference molecules.47 The relative energies of the species involved in the reaction were obtained by the focal point approach,1417 a convergent scheme in which both electron correlation and basis set limits are systematically approached. The correlation treatment was extended as far as the recently developed CCSDT(Q) method,4850 which includes contributions from quadruple excitations analogous to the triples term in the popular CCSD(T) theory. The CCSD(T) and CCSDT(Q) methods using the cc-pVXZ (X = D, T, Q, 5) family of basis sets51 were used for computation. The resulting singlepoint energies were extrapolated to the CBS limits using a threeparameter exponential function52 for the HartreeFock energy extrapolation and a two-term X3 form53 for the correlation energy extrapolation. In the focal point extrapolations, only valence electrons were correlated; thus, an auxiliary core correlation correction (Δcore) was determined by differencing AE and frozen-core CCSD(T)/cc-pCVQZ energetics. Because the hydrogen atoms rearrange in the reaction, high accuracy was ensured by including diagonal BornOppenheimer 2590

dx.doi.org/10.1021/jz201124j |J. Phys. Chem. Lett. 2011, 2, 2587–2592

The Journal of Physical Chemistry Letters corrections54,55 (ΔDBOC) evaluated at the HF level using the cc-pVTZ basis set.51 Relativistic effects (Δrel) were also accounted for by appending mass-velocity and Darwin one-electron terms56,57 computed with the AE-CCSD/cc-pCVTZ method. All geometric structures were optimized using the Molpro58 suite of electronic structure codes. The CCSDT(Q) correction employed in the focal point analysis was obtained using the string-based quantum chemistry code (MRCC) of Kallay, as interfaced with CFOUR.49,59,60 DBOC energies and relativistic corrections were also evaluated using CFOUR.60 NRT analysis was performed using NBO 5.061

’ SUMMARY We have examined the reaction C2H3 + H2 f C2H4 + H, an intermediary process that occurs during combustion and in planetary atmospheres. Through focal-point analysis, we have obtained molecular energies for this hydrogen abstraction step that are extrapolated to the CBS limit. Additionally we have included several corrections to this energy, namely, a corecorrection to account for “freezing” of the coreelectrons in computing the single-point energy, a correction for the zeropoint vibrational energy, a relativistic correction, and a correction for the BornOppenheimer approximation. Our geometries, obtained at the rigorous CCSD(T)/cc-pCVQZ level, agree reasonably well with prior theoretical work, and our computed vibrational modes for vinyl radical and ethylene generally align with previous experimental and theoretical studies. We report a barrier height of 9.65 kcal mol1 and overall reaction enthalpy of 5.65 kcal mol1 with a suggested accuracy of 0.1 kcal mol1. ’ ASSOCIATED CONTENT

bS

Supporting Information. Complete reference 49, data for focal point extrapolations, and data for stationary points. This material is available free of charge via the Internet at http://pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected].

’ ACKNOWLEDGMENT This work was supported by the Department of Energy, Office of Basic Energy Sciences (grant no. DEFG02-97-ER14748). This research used resources of the National Energy Research Scientific Computing Center (NERSC), which is supported by the Office of Science of the U.S. Department of Energy under contract no. DE-AC02-05CH11231. ’ REFERENCES (1) Noll, K. S.; Knacke, R. F.; Tokunaga, A. T.; Lacy, J. H.; Beck, S.; Serabyn, E. The Abundances of Ethane and Acetylene in the Atmospheres of Jupiter and Saturn. Icarus 1986, 65, 257–263. (2) Hanel, R. A.; Conrath, B. J.; Flaser, F. M.; Hearth, L.; Kunde, V.; Maguire, W.; Pearl, J.; Pirraglia, J.; Samuelson, R.; Gautier, D.; Gierasch, P.; Kumar, S.; Ponnamperuma, C. Infrared Observations of the Jovian System from Voyager 2. Science 1979, 206, 952–956. (3) Hanel, R. A.; Conrath, B. J.; Flaser, F. M.; Hearth, L.; Kunde, V.; Maguire, W.; Pearl, J.; Pirraglia, J.; Samuelson, R.; Gautier, D.; Gierasch, P.; Kumar, S.; Ponnamperuma, C. Infrared Observations of the Jovian System from Voyager 1. Science 1979, 204, 972–976.

LETTER

(4) Miller, J. A.; Melius, C. F. Kinetic and Thermodynamic Issues in the Formation of Aromatic Compounds in Flames of Aliphatic Fuels. Combust. Flame 1992, 91, 21–39. (5) Bennet, C. J.; Jamieson, C. S.; Osamura, Y.; Kaiser, R. I. Laboratory Studies on the Irradiation of Methane in Interstellar, Cometary, and Solar System Ices. Astrophys. J. 2006, 653, 792–811. (6) Atreya, S. K. Atmospheres and Ionospheres of the Outer Planets and Their Sattelites; Springer-Verlag: Berlin, 1986. (7) Romani, P. N.; Atreya, S. K. Methane Photochemistry and Haze Production on Neptune. Icarus 1988, 74, 424–445. (8) Strobel, D. F. Photochemistry of the Reducing Atmosphere of Jupiter, Saturn, and Titan. Int. Rev. Phys. Chem. 1983, 3, 145–176. (9) Baker, A. D.; Baker, C.; Brundle, C. R.; Turner, D. W. The Electronic Structures of Methane, Ethane, Ethylene and Formaldehyde Studied by High-Resolution Molecular Photoelectron Spectroscopy. Int. J. Mass Spectrom. Ion Phys. 1968, 1, 285–301. (10) Gordon, M. S.; Truong, T. N.; Pople, J. A. Thermal Decomposition Pathways of Ethane. Chem. Phys. Lett. 1986, 130, 245–248. (11) Ahern, A. M.; Garrell, R. L.; Jordan, K. D. Theoretical Investigation of the Normal-Mode Vibrational Frequencies and Intensities of Ethylene. J. Phys. Chem. 1988, 92, 6228–6232. (12) Sosa, C.; Schlegel, H. B. Ab Initio Calculations on the Barrier Height for the Hydrogen Addition to Ethylene and Formaldehyde. Int. J. Quantum Chem. 1986, 29, 1001–1015. (13) Duncan, J. L.; McKean, D. C.; Mallinson, P. D. Infrared Crystal Spectra of C2H4, C2D4 and as-C2H2D2 and the General Harmonic Force Field of Ethylene. J. Mol. Spectrosc. 1973, 45, 221–246. (14) Csaszar, A. G.; Allen, W. D.; Schaefer, H. F. In Pursuit of the Ab Initio Limit Conformational Energy Prototypes. J. Chem. Phys. 1998, 108, 9751–9764. (15) Schuurman, M. S.; Muir, S. R.; Allen, W. D.; Schaefer, H. F. Toward Subchemical Accuracy in Computational Thermochemistry: Focal Point Analysis of the Heat of Formation of NCO and [H,N,C,O] Isomers. J. Chem. Phys. 2004, 120, 11586–11599. (16) Gonzales, J. M.; Pak, C.; Cox, R. S.; Allen, W. D.; Schaefer, H. F.; Csaszar, A. G.; Tarczay, G. Definitive Ab Initio Studies of Model SN2 Reactions CH3X+F (X = F, Cl, CN, OH, SH, NH2, PH2). Chem. —Eur. J. 2003, 9, 2173–2192. (17) Csaszar, A. G.; Tarczay, G.; Leininger, M. L.; Polyansky, O. L.; Tennyson, J.; Allen, W. D. In Spectroscopy from Space; Demaison, J., Sarka, K., Eds.; Kluwer: Dordrecht, The Netherlands, 2001; p 317. (18) Laufer, A. H.; Fahr, A. Reactions and Kinetics of Unsaturated C2 Hydrocarbon Radicals. Chem. Rev. 2004, 104, 2813–2831. (19) Callear, A. B.; Smith, G. B. Recurring Chains Following Addition of Atomic Hydrogen to Acetylene. J. Phys. Chem. 1986, 90, 3229–3237. (20) Tsang, W.; Hampson, R. F. Chemical Kinetic Data Base for Combustion Chemistry. Part I. Methane and Related Compounds. J. Phys. Chem. Ref. Data 1986, 15, 1087–1279. (21) Mebel, A. M.; Morokuma, K.; Lin, M. C. Ab Initio Molecular Orbital Study of Potential Energy Surface for the Reaction of C2H3 with H2 and Related Reactions. J. Chem. Phys. 1995, 103, 3440–3449. (22) Fahr, A.; Monks, P. S.; Stief, L. J.; Laufer, A. H. Experimental Determination of the Rate Constant for the Reaction of C2H3 with H2 and Implications for the Partitioning of Hydrocarbons in Atmospheres of the Outer Planets. Icarus 1995, 116, 415–422. (23) Knyazev, V. D.; Bencsura, A.; Stoliarov, S. I.; Slagle, I. R. Kinetics of the C2H3 + H2 a H + C2H4 and CH3 + H2 a H + CH4. J. Phys. Chem. 1996, 100, 11346–11354. (24) Li, Q. S.; L€u, R. H.; Wang, C. Y. Accurate Ab Initio Calculations on the Rate Constants of the Hydrogen Abstraction Reaction C2H3 + H2 f H + C2H4. J. Mol. Struct. 2004, 668, 35–39. (25) Tautermann, C. S.; Wellenzohn, B.; Clary, D. C. Rates of the Reaction C2H3 + H2 f C2H4 + H. Mol. Phys. 2006, 104, 151–158. (26) Weissman, M. A.; Benson, S. W. Rate Parameters for the Reactions of Vinyl and Butadienyl Radicals with Hydrogen and Acetylene. J. Phys. Chem. 1988, 92, 4080–4084. 2591

dx.doi.org/10.1021/jz201124j |J. Phys. Chem. Lett. 2011, 2, 2587–2592

The Journal of Physical Chemistry Letters (27) Moran, D.; Simmonett, A. C.; Leach, F. E.; Allen, W. D.; Schleyer, P. v. R.; Schaefer, H. F. Popular Theoretical Methods Predict Benzene and Arenes to be Nonplanar. J. Am. Chem. Soc. 2006, 128, 9342–9343. (28) Letendre, L.; Liu, D. K.; Pibel, C. D.; Halpern, J. B.; Dai, H. L. Vibrational Spectroscopy of a Transient Species Through TimeResolved Fourier Transform Infrared Emission Spectroscopy: The Vinyl Radical. J. Chem. Phys. 2000, 112, 9209–9212. (29) Nikow, M.; Wilhelm, M. J.; Dai, H. L. Vibrational Modes of the Vinyl and Deuterated Vinyl Radicals. J. Phys. Chem. A 2009, 113, 8857–8870. (30) Hammond, G. S. A Correlation of Reaction Rates. J. Am. Chem. Soc. 1955, 77, 334–338. (31) Shepherd, R. A.; Doyle, T. J.; Graham, W. R. M. A Fourier Transform Infrared Study of the D and 13C Substituted C2H3 Vinyl Radical in Solid Argon. J. Chem. Phys. 1988, 89, 2738–2742. (32) Kanamori, H.; Endo, Y.; Hirota, E. The Vinyl Radical Investigated by Infrared Diode Laser. J. Chem. Phys. 1990, 92, 197–205. (33) Zhang, J. S.; Xu, K. S. Photodissociation of the Vinyl Radical (C2H3) via the First Excited State: The C2H2(~x1 ∑g+)+H Channel. J. Chem. Phys. 1999, 111, 3783–3786. (34) Pibel, C. D.; McIllroy, A.; Taatjes, C. A.; Alfred, S.; Patrick, K.; ~2A00 r X ~ 2A0 ) Spectrum between 530 Halpern, J. B. The Vinyl Radical (A and 415 nm Measured by Cavity Ring-Down Spectroscopy. J. Chem. Phys. 1999, 110, 1841–1843. (35) Ahmed, M.; Peterka, D. S.; Suits, A. G. The Photodissociation of the Vinyl Radical (C2H3) at 243 nm Studied by Velocity Map Imaging. J. Chem. Phys. 1999, 110, 4248–4253. (36) Fahr, A.; Hassanzadeh, P.; Atkinson, D. B. Ultraviolet Absorption Spectrum and Cross-Sections of Vinyl (C2H3) Radical in the 225238 nm Region. Chem. Phys. 1998, 236, 43–51. (37) Fahr, A.; Laufer, A. H. Ultraviolet Absorption of the Vinyl Radical and Reaction with Oxygen. J. Phys. Chem. 1988, 92, 7229–7232. (38) Hunziker, H. E.; Kneppe, H.; McLean, A. D.; Siegbahn, P.; Wendt, H. R. Visible Electronic Absorption Spectrum of Vinyl Radical. Can. J. Chem. 1983, 61, 993–995. (39) Sattelmeyer, K. W.; Schaefer, H. F. The v5 Vibrational Frequency of the Vinyl Radical: Conflict Between Theory and Experiment. J. Chem. Phys. 2002, 117, 7914–7916. (40) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A Fifth-Order Pertubation Comparison of Electron Correlation Theories. Chem. Phys. Lett. 1989, 157, 479–483. (41) Stanton, J. F. Why CCSD(T) Works: A Different Perspective. Chem. Phys. Lett. 1997, 281, 130–134. (42) Watts, J. D.; Gauss, J.; Bartlett, R. J. Open-Shell Analytical Energy Gradients for Triple Excitation Many-Body, Coupled-Cluster Methods: MBPT(4), CCSD+T(CCSD), CCSD(T), and QCISD(T). Chem. Phys. Lett. 1992, 200, 1–2. (43) Watts, J.; Gauss, J.; Bartlett, R. Coupled-Cluster Methods with Noniterative Triple Excitations for Restricted Open-Shell Hartree-Fock and Other General Single Determinant Reference Functions. Energies and Analytical Gradients. J. Chem. Phys. 1993, 98, 8718–8733. (44) Hampel, C.; Peterson, K. A.; Werner, H.-J. A Comparison of the Efficiency and Accuracy of the Quadratic Configuration Interaction (QCISD), Coupled Cluster (CCSD), and Bruekner Coupled Cluster (BCCD) Methods. Chem. Phys. Lett. 1992, 190, 1–2. (45) Deegan, M. J. O.; Knowles, P. J. Pertubative Corrections to Account for Triple Excitations in Closed and Open Shell Coupled Cluster Theories. Chem. Phys. Lett. 1992, 227, 321–326. (46) Woon, D. E.; Dunning, T. H. Gaussian-Basis Sets for Use in Correlated Molecular Calculations. 5. Core-Valence Basis-Sets for Boron Through Neon. J. Chem. Phys. 1995, 103, 4572–4585. (47) Bak, K. L.; Gauss, J.; Jørgensen, P.; Olsen, J.; Helgaker, T.; Stanton, J. F. The Accurate Determination of Molecular Equilibrium Structures. J. Chem. Phys. 2001, 114, 6548–6556. (48) Bomble, Y.; Kallay, M.; Gauss, J.; Stanton, J. F. Coupled-Cluster Methods Including Noniterative Corrections for Quadruple Excitations. J. Chem. Phys. 2005, 123, 054101.

LETTER

(49) Kallay, M.; Gauss, J. Approximate Treatment of Higher Excitations in Coupled-Cluster Theory. J. Chem. Phys. 2005, 123, 214105. (50) Kallay, M.; Gauss, J. Approximate Treatment of Higher Excitations in Coupled-Cluster Theory. II. Extension to General SingleDeterminant Reference Functions and Improved Approaches for the Canonical Hartree-Fock Case. J. Chem. Phys. 2008, 129, 144101. (51) Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron Through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. (52) Feller, D. The Use of Systematic Sequences of Wave Functions for Estimating the Complete Basis Set, Full Configuration Interaction Limit in Water. J. Chem. Phys. 1993, 98, 7059–7071. (53) Helgaker, T.; Klopper, W.; Koch, H.; Noga, J. Basis-Set Convergence of Correlated Calculations on Water. J. Chem. Phys. 1997, 106, 9639–9646. (54) Handy, N. C.; Yamaguchi, Y.; Schaefer, H. F. The Diagonal Correction to the Born-Oppenheimer Approximation: Its Effect on the Singlet-Triplet Splitting of CH2 and Other Molecular Effects. J. Chem. Phys. 1986, 84, 4481–4484. (55) Sellers, H.; Pulay, P. The Adiabatic Correction to Molecular Potential Surfaces in the SCF Approximation. Chem. Phys. Lett. 1984, 103, 463–465. (56) Perera, S. A.; Bartlett, R. J. Relativistic Effects at the Correlated Level. An Application to Interhalogens. Chem. Phys. Lett. 1993, 216, 606–612. (57) Cowan, R. D.; Griffin, D. C. Approximate Relativistic Corrections to Atomic Radial Wave-Functions. J. Opt. Soc. Am. 1976, 66, 1010–1014. (58) Werner, H.-J. et al. MOLPRO, a Package of Ab Initio Programs, version 2006.1, 2006. http://www.molpro.net. (59) Kallay, M.; Surjan, P. Higher Excitations in Coupled-Cluster Theory. J. Chem. Phys. 2001, 115, 2945–2954. (60) CFOUR, a quantum chemical program package written by Stanton, J. F.; Gauss, J.; Watts, J. D.; Szalay, P. G.; Bartlett, R. J. with contributions from Auer, A. A.; Bernholdt, D. E.; Christiansen, O., Harding, M. E.; Heckert, M.; Heun, O.; Huber, C.; Jonsson, D.; Juselius, J.; Lauderdale, W. J.; Metzroth, T.; Michauk, C.; O’Neill, D. P.; Price, D. R.; Ruud, K.; Schiffmann, F.; Tajti, A.; Varner, M. E.; Vazquez, J. and the integral packages: MOLECULE (Almof, J.; Taylor, P. R.), PROPS (Taylor, P. R.), ABACUS (Helgaker, T.; Jensen, H. J. Aa.; Jørgensen, P.; Olsen, J.), and ECP routines by Mitin, A. V.; van W€ullen, C. For the current version, see http://www.cfour.de. (61) Glendening, E. D.; Badenhoop, J. K.; Reed, A. E.; Carpenter, J. E.; Bohmann, J. A.; Morales, C. M.; Weinhold, F. NBO 5.0; Theoretical Chemistry Institute, University of Wisconsin, Madison: Madison, WI, 2001.

2592

dx.doi.org/10.1021/jz201124j |J. Phys. Chem. Lett. 2011, 2, 2587–2592