Pfaendtner Et Al 2006 Quantum Chemical Investigation of Low Temperature Intramolecular Hydrogen Transfer Reactions of
Pfaendtner Et Al 2006 Quantum Chemical Investigation of Low Temperature Intramolecular Hydrogen Transfer Reactions of
The B3LYP functional was evaluated as a method to calculate reaction barriers and structure-reactivity
relationships for intramolecular hydrogen transfer reactions involving peroxy radicals. Nine different basis
sets as well as five other MO/DFT and hybrid methods were used in comparing three reactions to available
experimental data. It was shown that B3LYP/6-311+G(d,p) offers a good compromise between speed and
accuracy for studies in which thermodynamic and kinetic data of many reactions are required. Sixteen reactions
were studied to develop structure-reactivity relationships to correlate the activation energy with the heat of
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
reaction. As long as no structural heterogeneities were present in the transition state ring, a simple Evans-
Polanyı́ relationship was shown to capture the activation energy as a function of heat of reaction for reactions
Downloaded via NORTHWESTERN UNIV on October 27, 2023 at 19:58:13 (UTC).
in the 1,5-hydrogen shift family. For peroxy radicals undergoing self-abstraction of a hydrogen atom in the
1,5-position, the activation energy was calculated as Ea (kcal mol-1) ) 6.3 + ∆Hrxn (kcal mol-1). For reactions
with a carbonyl group embedded in the ring of the transition state, the activation energy of peroxy radicals
undergoing self-abstraction was correlated as Ea (kcal mol-1) ) 18.1 + 0.74*∆Hrxn (kcal mol-1). The impact
of the size of the transition state ring on the activation energy and pre-exponential factor was also probed,
and it was shown that these effects can be described using simple nonlinear and linear fits, respectively.
tions were also performed using G3(MP2)27 in order to probe bonds, angles, and dihedrals were analyzed using the internal
the effect of a high-level correction using MP2 optimized mode analysis provided by Gaussian 03. In large molecules,
geometries. Accepted B3LYP scaling factors of 0.9806 for zero- there is often substantial mixing of vibrational modes resulting
point energies and 0.9614 for other calculations, e.g., partition in a number of low frequencies corresponding to multiple
functions, were used.28 Scott and Radom optimized these scaling different internal rotations. Potential problems arising from this
factors specifically for the 6-31G(d) basis set, but given the mixing of vibrational modes have recently been outlined.39 Our
absence of scale factors for other basis sets, it has become present approach is to use the internal mode analysis to identify
commonplace to scale results using these values for B3LYP, a single low-frequency vibration for each internal rotation in a
regardless of the basis set.17,29 Finally, solvation effects were given molecule. Thus, the HO partition function for a molecule
studied using a polarizable continuum model (PCM).30 The with n dihedral angles was corrected by (i) removing the
solvent cavity was established using radii obtained from the contribution of n low-frequency vibrational modes and (ii)
universal force field31 with hydrogen atoms explicitly defined. including the contribution of n internal rotations.
Full geometry optimization and energy calculations were Once the partition functions for each reactant and transition
performed with n-heptane as the solvent using B3LYP/6-31G- state were corrected for internal rotation, the rate constant as a
(d). function of temperature was calculated using the standard
Using the optimized geometry and frequencies obtained from expression,
a Hessian calculation, the total microcanonical partition function
can be calculated for each molecule within the rigid-rotor kbT Qq(T)
harmonic oscillator (HO) assumption. The formulas for elec- kTST ) rpd*κ(T) exp(-Eo/RT) (2)
tronic, translational, rotational and vibrational partition functions
h Qr(T)
are well established.32 However, the HO approximation incor-
rectly treats low-frequency rotations as harmonic oscillators. If where rpd is the reaction path degeneracy, determined by the
left uncorrected, this can induce substantial error in calculated number of identical hydrogen atoms a particular radical can self-
kinetic and thermodynamic data. Internal rotations whose barrier abstract, Qq and Qr are the total partition functions for the TS
to rotation is much less than kT are characterized as free rotors and reactant, respectively, and Eo is the reaction barrier, i.e.,
and corrected easily.33 In the temperature range of interest to the zero-point corrected difference between the ground-state
the current study (300-1000 K), the barrier to rotation for all electronic energies of the TS and the reactant. The quantity κ(T)
dihedrals is on the order of kT, therefore requiring that internal is the correction for quantum tunneling. To quantify the effect
rotations be treated as hindered rotors. Our approach in treating of quantum tunneling on the reactions of interest here, we
internal rotation was similar in principle to previously reported calculated various approximations to the transmission coefficient
methods.29,34 First, the hindrance potential energy surface (PES) including the Wigner40 correction, the Eckart approximation,41
for each rotation about polyvalent single bonds was obtained the zero-curvature approximation (ZCT),41 and the small
by scanning the dihedral angle in 30° increments. The dihedral curvature tunneling (SCT)41 approximation. Calculation of the
angle was fixed and the rest of the structure was allowed to Eckart, SCT, and ZCT corrections was performed using the
relax with the geometry optimized using the B3LYP/6-31G(d) software package “Virtual Kinetic Laboratory” of Zhang and
level of theory. Although rigid scans are more computationally Truong.42 Once the temperature-dependent partition functions
efficient, significant errors were encountered due to the branch- were known, rate constants as a function of temperature were
ing of the molecules under consideration. In particular, the calculated, and the Arrhenius parameters A and Ea were obtained
peroxy radical and hydroperoxide moieties gave nonphysical from a straight-line fit of ln k vs 1/T. Unless otherwise stated,
PES results when the superstructure was kept rigid. Examples we used a temperature range of 300-1000 K for regression of
of this are provided in the Supporting Information. To preserve A and Ea. To calculate the enthalpy of reaction, all thermody-
the structure of the reactive center during relaxed PES scans of namic parameters were obtained from ensemble energy averages
transition states, all coordinates that constituted the ring of each 〈E〉 using standard formulas.32
transition state were frozen, while all other coordinates except Thermochemistry of Peroxide Species. Although the B3LYP
the dihedral angle being scanned were optimized. Once the PES functional has been repeatedly shown to give accurate results
for each rotation was obtained, the one-dimensional Schrödinger for the heat of reaction and heat of formation, especially when
equation was solved in order to obtain the energy levels for compared to other methods of similar computational cost, it has
each rotation. This step was performed using the Fourier grid been documented that B3LYP systematically gives errors in the
Hamiltonian (FGH) method developed by Marston and co- bond dissociation energy of the O-O and O-H bonds in
workers.35,36 To obtain the correct energy levels, the numerical hydroperoxide moieties.3,43 The standard approach is to use
algorithm for the FGH method was implemented using published isodesmic reactions to obtain a corrected heat of formation.44
source code.37 The energy levels from this calculation were used Although this correction for reactants and products is straight-
to calculate the partition function for each internal rotation as forward, an analogous treatment for the TS is not obvious.
a function of temperature: Therefore, the heats of reaction reported are the difference in
the quantum chemical enthalpies between the reactant and
Qir )
1
σ i
r
( )
∑gi exp -kT
i
(1)
product species, which offered a method that was consistent
with the manner in which the activation energy was calculated
for the structure/reactivity relationships that were developed.
Treatment of Rotational Isomers. It was observed that in
where σr is the internal symmetry number of the rotating top some cases the IRC-derived reactants were not the minimum
and gi is the degeneracy of the ith energy level, i.29 energy structures. Subsequent rotation of the C-O bond in the
In making the correction to the HO partition function, the peroxy radical or rotation about the backbone of the molecule
vibrational modes corresponding to internal rotation must be could yield a reactant lower in energy typically by no more
identified and their contribution to the HO partition function than 1 kcal mol-1. While these were relatively subtle differences,
removed.34,38 For each vibration, the contributions from various it was still important to account for them in the evaluation of
10866 J. Phys. Chem. A, Vol. 110, No. 37, 2006 Pfaendtner et al.
methods and development of the structure/reactivity relation- using B3LYP/6-31G(d). As shown in Table 1, the PCM model
ships. For reactants in which a lower energy structure was found, does not result in any appreciable change in the calculated
the following modified rate constant was used: kinetic properties or enthalpy of reaction. This agreement
between gas-phase TST calculations and condensed-phase
kapparent ) Krot
eq (T)*kTST(T) (3) kinetic data has also been observed for other reaction classes.46
However, it is important to note the rate of hydrogen abstraction
where Krot
eq is the equilibrium constant between two rotational
in the condensed phase has also been shown to be a strong
isomers: function of the polarity of the solvent.47
These results add to the growing database of activation
-∆G rot(T)/RT
eq ) e
Krot (4) energies predicted by B3LYP, which collectively suggests that
B3LYP does not exclusively under- or overestimate reaction
Thus, the activation energy and pre-exponential factor were barriers. Furthermore, it is not generally possible a priori to state
regressed from a fit of ln(kapparent) versus 1/T. A full derivation in which direction B3LYP will err. For example, Wijaya and
of this relationship is provided in the Supporting Information. co-workers3 report an underestimation by B3LYP of the barrier
height for different cyclization reactions while Henry and
3. Results and Discussion Radom46 report that B3LYP always overpredicts the barrier for
the cyclization of but-3-enyl-radical. The heats of reaction are
Effect of Basis Set on Results Obtained from the B3LYP
not available from experiment, but two observations can be
Functional. Results for the comparison of 16 different methods
drawn based on the results provided. First, results from very
and the 3 different reactions studied in detail (reactions 1-3 in
small basis sets deviate strongly from the CBS and G3 methods,
Figure 1) are listed in Table 1. The first 11 entries in each section
with errors as large as 4 kcal mol-1. Moving to very large basis
of the table list B3LYP calculations performed with increasingly
sets improves the agreement with the CBS and G3 results, with
larger basis sets. The one non-Pople basis set used was listed
some B3LYP heats of reaction higher than the CBS-QB3 values
as the last of these 11 for clarity. The bottom five lines in each
by only 1 kcal mol-1. Finally, it is worthwhile to note that the
section of Table 1 contain results for these reactions with
barriers calculated using B3LYP/6-311+G(3df,2p) differ from
different MO, DFT, and hybrid methods. The results show that
those calculated using single point calculations from the same
the choice of the basis set has a strong impact on both the
method but with geometries optimized using B3LYP/6-31G-
reaction barrier and the heat of reaction. As expected, increasing
(d). Using the smallest basis set (6-31G(d)) for geometry
the size of the basis set generally improves the agreement
optimization and the largest basis set (6-311+G(3df,2p)) for
between the B3LYP results and experimental data or results
single point calculations gives activation energies that may
from the high-level compound methods. With one exception in
deviate significantly from the experimental, CBS, and G3 results.
which the calculated activation energy was equal to the
experimental value, all of the B3LYP methods overestimate the It is next interesting to compare the five higher level methods
experimental activation energies. This corresponds to an un- in the bottom of each section of Table 1 with each other. Barriers
derprediction of the reaction rate coefficient. One possible calculated at the MP2 and G3(MP2) levels of theory are in very
explanation for this could be the type of calculation used to poor agreement with the experimental data for all three reactions.
estimate the contribution of quantum tunneling, but our results To test if the cause of this discrepancy was from the optimized
show that the type of tunneling correction used did not change geometries or from the single point/hybrid energies, we
the calculated activation energies significantly. We calculated calculated the G3//B3LYP energy of the reactant and TS of the
the Eckart, ZCT, and SCT transmission coefficients for reactions MP2/6-311+G(d,p) optimized structures for reaction 2. The
1 and 2 using the B3LYP/6-31G(d) and B3LYP/6-311+G(d,p) calculated activation energy was nearly identical to the full G3//
basis sets. The Wigner correction was a good approximation B3LYP activation energy, thus indicating that the energies
of the SCT transmission coefficient (within a factor of 2 or less calculated using MP2 are poor. The BHandHLYP/6-311G(d,p)
at the temperature range of interest (373 K)). Accordingly, we method also substantially overpredicts the activation barrier
calculated all kinetic data in Table 1 using the Wigner correction. compared to the CBS-QB3 method. Despite their poor predic-
Furthermore, the transmission coefficient had a maximum value tions of the barrier height, BHandHLYP/6-311G(d,p), MP2/6-
of 3.7 at 300 K for the reactions we studied, which is much 311+G(d,p), and G3(MP2) still give pre-exponential values that
lower than that observed for intermolecular hydrogen transfer are in good agreement with the value obtained from CBS-QB3,
reactions of small molecules studied by Truong and co-workers45 and the heats of reaction are in reasonable agreement with the
and Gonzalez-Lafont and co-workers.41 The small contribution B3LYP values using large basis sets. In comparing the two
of tunneling is consistent with the adiabatic ground state hybrid methods, the results reveal that in the case of the first
potential clearly showing a Very late transition state. The and second reactions, CBS-QB3 gives results that are much
potential is nearly flat as the reaction goes from the transition closer to the experimentally reported values, whereas the
state to the product hydroperoxide radical (forward direction experimental values for the third reaction more closely match
as depicted in Figure 1). Because tunneling is not observed to the G3B3 results. However, each activation energy was obtained
have a strong impact in these reactions, it is not surprising that from a single source without any repeatability reported.
for each reaction in Table 1, a plot of ln k versus 1/T gave a Therefore, the comparison set is too limited to draw any
perfect straight-line fit, thereby validating our method of definitive conclusion about the superiority of either of these
obtaining the Arrhenius parameters A and Ea. methods. Additionally, in all cases, the calculated rate coef-
Because we are seeking kinetic correlations for use in ficients (kTST(373 K)) from the CBS-QB3 method give the best
condensed-phase hydrocarbon oxidation chemistry, we tested agreement to the available experimental data. For calculating
the sensitivity of the calculated data to the presence of a nonpolar rate coefficients, the G3//B3LYP method performs noticeably
solvent, i.e., a typical oxidation substrate of interest. We used better than pure B3LYP methods for reaction three, but at a
the PCM model to simulate a solvent cavity of n-heptane and comparable (or lower) level of accuracy for reactions 1 and 2.
reoptimized the reactants, TS, and products for reactions 1-3 For example, in the data set for reaction 1, B3LYP/cc-pVTZ+d
Low-Temperature Oxidation of Hydrocarbons J. Phys. Chem. A, Vol. 110, No. 37, 2006 10867
TABLE 1: Comparison of Basis Set and Method for Three Intramolecular Hydrogen Abstraction (Reactions 1-3 in Figure 1)
reactants transition state
methoda Eob ∆Hrxn c Ead log Ae kTST(373 K)f νirg 〈S2〉h νir g νii 〈S2〉h
reaction 1j C(CH3)2(O2 •)CH
2CH(CH3)2 f C(CH3)2(O2H)CH2C‚(CH3)2
6-31G(d) 19.8 17.8 21.1 10.8 2.73 × 10-2 47,75,110,181,207,229,242 0.753 193,201,205,225 -1626 0.756
6-31G(d)k 19.4 17.2 20.6 10.7 4.26 × 10-2
RO 6-31G(d)l 20.0 17.7 21.2 10.8 2.38 × 10-2 47,75,111,181,207,229,242 194,201,205,225 -1695
6-31G(d,p) 18.0 15.5 19.2 10.8 3.54 × 10-1 48,76,111,180,209,219,243 0.753 195,200,206,226 -1610 0.756
6-311G(d,p) 17.7 14.3 19.0 10.7 3.69 × 10-1 47,76,111,178,202,215,243 0.754 194,197,206,227 -1653 0.756
6-311+G(d,p) 17.7 13.7 18.7 10.6 4.39 × 10-1 48,75,103,178,208,228,242 0.754 193,200,205,227 -1654 0.756
RO 6-311+G(d,p)l 17.8 13.6 18.8 10.6 3.83 × 10-1 48,75,104,178,208,228,242 194,200,205,227 -1713
6-311+G(df) 20.2 16.7 21.3 10.7 1.65 × 10-2 55,77,99,178,213,220,245 0.754 194,203,206,228 -1694 0.757
6-311+G(3df,2p) 17.4 13.4 18.5 10.7 7.24 × 10-1 47,72,101,177,208,228,242 0.755 192,199,204,226 -1675 0.757
6-311+G(3df,2p)m 16.6 13.4 19.1 10.8 4.05 × 10-1 47,75,110,181,207,229,242 0.753 193,201,205,225 -1626 0.753
cc-pVTZ+d 17.4 13.6 18.5 10.7 7.24 × 10-1 44,74,106,178,204,227,240 0.754 194,196,204,226 -1666 0.756
MP2/6-311+G(d,p)e 34.1 17.7 35.7 10.6 4.79 × 10-11 60,66,93,180,221,232,259 0.763 208,215,228,252 -2033 0.794
BHandHLYP/6-311G(d,p) 23.4 14.7 24.7 10.7 1.68 × 10-4 51,76,119,187,217,234,256 0.757 206,210,219,242 -2133 0.765
G3(MP2) 43.7 13.7 44.8 10.9 4.45 × 10-16 48,71,124,193,241,252,267 0.761 262,302,315,336 -2777 0.797
G3\\B3LYP 17.9 12.9 19.2 10.8 3.54 × 10-1 47,75,110,181,207,229,242 0.753 193,201,205,225 -1626 0.756
CBS-QB3 16.3 12.3 17.5 10.6 2.22 44,79,111,178,200,219,244 0.754 185,202,215,238 -1653 0.756
experimental 17.5 11.5 18
reaction 2 CH3C(O2•)H(CH2)2CH3 f CH3C(O2H)HCH2C‚HCH3
6-31G(d) 24.0 19.6 23.1 10.9 2.31 × 10-3 71,79,108,169,206 0.753 193,210 -1628 0.757
6-31G(d)k 23.5 19.1 22.5 10.8 4.13 × 10-3
RO 6-31G(d)l 24.9 20.2 23.9 10.8 6.24 × 10-4 78,86,107,176,241 187,208 -1711
6-31G(d,p) 22.7 17.9 21.7 10.8 1.21 × 10-2 73,84,106,176,234 0.753 189,206 -1637 0.756
6-311G(d,p) 22.5 16.8 21.5 10.8 1.59 × 10-2 77,82,104,176,232 0.753 189,197 -1682 0.756
6-311+G(d,p) 22.3 16.3 21.3 10.8 2.08 × 10-2 71,79,108,176,230 0.754 190,203 -1690 0.757
RO 6-311+G(d,p)l 22.5 16.0 21.5 10.8 1.59 × 10-2 71,81,108,176,230 191,203 -1753
6-311+G(df) 25.1 19.2 24.1 10.8 4.76 × 10-4 68,78,104,176,231 0.754 191,206 -1726 0.757
6-311+G(3df,2p) 22.1 15.8 21.1 10.8 2.73 × 10-2 72,80,109,176,227 0.755 187,202 -1716 0.757
6-311+G(3df,2p)m 20.6 19.7 10.9 2.27 × 10-1 71,79,108,169,206 0.753 193,210 -1628 0.757
cc-pVTZ+d 21.3 15.3 20.4 10.8 7.02 × 10-2 39,76,106,167,200 0.754 188,201 -1703 0.757
MP2/6-311+G(d,p)n 40.6 19.3 39.4 10.7 4.10 × 10-13 67,73,109,175,242 0.763 213,215 -2098 0.796
BHandHLYP/6-311G(d,p) 27.6 16.0 26.3 10.8 2.45 × 10-5 54,80,112,171,214 0.756 202,211 -2190 0.765
G3(MP2) 40.9 15.3 38.9 9.9 1.28 × 10-13 73,87,117,179,237 0.761 124,163 -2979 0.800
G3\\B3LYP 21.7 14.7 20.8 10.9 5.15 × 10-2 71,79,108,169,206 0.753 193,210 -1628 0.757
CBS-QB3 20.2 14.4 19.3 10.9 3.90 × 10-1 47,77,107,167,202 0.754 189,197 -1682 0.756
experimental 19.7 11.5 8.7 × 10-1
reaction 3j C(CH3)2(O2•)(CH2)2CH(CH3)2 f C(CH3)2(O2H)(CH2)2C‚(CH3)2
6-31G(d) 20.0 16.6 19.9 9.4 5.48 × 10-3 49,79,99,139,198,207,230,230 0.753 197,204,219,230 -1628 0.757
6-31G(d)k 19.7 16.2 19.6 9.4 8.22 × 10-3
RO 6-31G(d)l 20.3 16.4 20.1 9.4 4.19 × 10-3 49,79,99,140,198,207,230,231 198,204,220,231 -1758
6-31G(d,p) 18.3 14.3 18.0 9.4 7.12 × 10-2 50,80,100,141,197,209,230,231 0.753 199,205,219,231 -1654 0.756
6-311G(d,p) 17.9 13.1 17.7 9.3 8.48 × 10-2 50,81,100,143,193,207,225,229 0.754 197,206,216,232 -1695 0.757
6-311+G(d,p) 17.8 13.1 18.0 9.3 5.66 × 10-2 49,76,95,138,198,209,226,228 0.754 196,207,218,232 -1702 0.757
RO 6-311+G(d,p)l 17.9 12.8 18.1 9.3 4.94 × 10-2 49,76,95,138,199,208,226,229 196,207,218,232 -1773
6-311+G(df) 20.2 16.0 20.4 9.3 2.22 × 10-3 51,72,96,140,201,211,227,232 0.754 196,207,220,232 -1740 0.757
6-311+G(3df,2p) 17.7 12.8 18.0 9.3 5.66 × 10-2 46,72,92,133,198,206,226,228 0.755 194,205,216,231 -1730 0.757
6-311+G(3df,2p)m 16.7 19.3 9.4 1.23 × 10-2 49,79,99,139,198,207,230,230 0.753 197,204,219,230 -1675 0.757
cc-pVTZ+d 17.8 12.9 18.0 9.3 5.66 × 10-2 46,75,95,137,196,206,226,227 0.754 195,204,215,231 -1721 0.757
MP2/6-311+G(d,p)n 34.2 18.4 34.9 9.3 7.07 × 10-12 60,86,108,143,203,223,228,242 0.763 213,223,227,238 -2164 0.794
BHandHLYP 23.4 13.7 23.0 9.2 5.28 × 10-5 50,85,106,149,206,218,233,241 0.757 210,220,223,239 -2214 0.765
G3(MP2) 47.0 12.0 44.7 7.5 2.03 × 10-19 43,79,103,153,217,239,246,257 0.761 83,95,198,254 -2875 0.797
G3\\B3LYP 17.9 10.6 16.7 9.4 4.11 × 10-1 49,79,99,139,198,207,230,230 0.753 197,204,219,230 -1675 0.757
CBS-QB3 16.5 10.6 15.9 9.3 9.62 × 10-1 50,81,100,144,194,208,226,229 0.754 197,206,216,232 -1695 0.757
experimental 17.2 11 8
a Unless alternate MO or hybrid method is given, the B3LYP functional is used and only the basis set is provided. b Reaction barrier (in kcal/
mol). c Heat of reaction (in kcal/mol) at 298 K. d Arrhenius activation energy (in kcal/mol). e Pre-exponential factor (in 1/s). f Rate coefficient (in
1/s) at 373 K. g Frequencies (in cm-1, unscaled) which were identified as hindered rotors. h Expectation value of the total spin for the reactants and
the transition states. i Imaginary vibrational mode (in cm-1) of the transition state. j Lower energy rotational isomer of IRC-derived reactant or
product was found, and eq 3 was used to calculate the pre-exponential factor and activation energy for the forward reaction. k Geometry optimization
and energy calculation performed using UB3LYP/6-31G(d) and a PCM model using the properties of n-heptane to describe the solvent cavity.
l Calculation performed with restricted-open shell wavefunctions. m Geometry optimization performed at UB3LYP/6-31G(d) with single point and
frequency calculation at given basis set. n MP2 scaling factors of 0.9670 (ZPE) and 0.9434 (fundamental frequencies) were used according to the
recommendation of Scott and Radom.28
gives a rate coefficient of 7.24 × 10-1 (1/s), whereas the G3//B3LYP for the reactants of reactions 2 and 3 in Figure 1.
prediction of G3//B3LYP is 3.54 × 10-1 (1/s). Both calculations The largest basis sets that most closely approximate the hybrid
are far from the single experimental value of 18 (1/s), but the methods cost slightly more than the CBS-QB3 calculation,
cc-pVTZ+d value is a factor of 2 closer. whereas G3//B3LYP is noticeably more expensive than the other
Given the variability of the results from all of the different methods. The high computational cost of G3//B3LYP is due
methods and basis sets, it is instructive to compare their almost entirely to the single point calculation using the MP4/
computational cost. Figure 2 gives the relative computational 6-31G(2df,p) level of theory. On the basis of the relatively good
cost for B3LYP using the 6-31G(d), 6-311+G(d,p), 6-311+G- performance of the B3LYP/6-311+G(d,p) calculations on all
(3df,2p), and cc-pVTZ+d basis sets, as well as CBS-QB3 and three reactions and its modest computational cost, we selected
10868 J. Phys. Chem. A, Vol. 110, No. 37, 2006 Pfaendtner et al.
TABLE 2: Comparison of Activation Energy and Pre-Exponential Factor as the Ring Size of the Transition State Is Varied
from 4 to 7a
reaction TS ring size Eab log A (1/s)
CH3CH(O2•)CH3 f CH3COCH3 + HO• c 4 41.6 12.6
(CH3)2C(O2•)CH(CH3)2 f (CH3)2C(O2H)C‚(CH3)2 5 28.0 11.5
C(CH3)2(O2•)CH2CH(CH3)2 f C(CH3)2(O2H)CH2C‚(CH3)2 6 18.7 10.6
C(CH3)2(O2•)(CH2)2CH(CH3)2 f C(CH3)2(O2H)(CH2)2C‚(CH3)2 7 18.0 9.3
a
Geometry optimization and energy calculations were performed using B3LYP/6-311+G(d,p). b Quantity in kcal mol-1. c The 1,3-hydrogen
shift creates the unstable •COOH containing radical, which directly yields the β-scission products shown in the table.
{
between 1.6 and 2.8 kcal mol ; however, the same form of
for ∆Hrxn/4Eo < -1 structure-reactivity relationship was observed, with Eo and R
0
values of 15.6 kcal mol -1 and 0.83/0.17.
Ea ) Eo(1 + ∆Hrxn/4Eo) for -1 e ∆Hrxn/4Eo e 1 (5)
2
TABLE 3: Data Used in Development of Structure-Reactivity Relationship for 1,5-Hydrogen Shift Familya
forward reaction reverse reaction
number reaction ∆Hrxn b Eab log A (1/s) Eab log A (1/s)
1 C(CH3)2(O2•)CH2CH(CH3)2 f C(CH3)2(O2H)CH2C‚(CH3)2 c 13.7 18.7 10.6 5.1 9.7
2 CH3C(O2•)H(CH2)2CH3 f CH3C(O2H)HCH2C‚HCH3 16.3 21.3 10.8 5.8 10.7
4 (CH3)2CHCH2C(O2•)(CH3)CH2CH(CH3)2 f 8.6 16.8 11.3 6.1 11.1
(CH3)2C‚CH2C(O2H)(CH3)CH2CH(CH3)2
5 (CH3)CH(O2•)CH2CH(CH3)2 f (CH3)CH(O2H)CH2C‚(CH3)2 c 14.9 21.6 10.6 8.0 10.9
6 (CH3)CH(O2•)CH2CH3 f (CH3)CH(O2H)CH2C‚H2c 19.6 26.2 11.2 7.1 11.6
7 (CH3)CH(O2•)COCH3 f (CH3)CH(O2H)COC‚H2 11.6 26.3 11.1 15.3 12.5
8 (CH3CH(O2•)HCOCH2CH3 f (CH3CH(O2H)HCOCH‚CH3 4.0 21.3 11.6 17.6 12.0
8 (CH3CH(O2•)HCOCH2CH3 f (CH3CH(O2H)HCOCH‚CH3 d,e 2.7 20.8 11.7 17.8 11.0
9 (CH3)CH(O2•)COCH(CH3)2 f (CH3)CH(O2H)COC‚(CH3)2 1.5 19.1 10.4 17.4 10.1
a
When applicable, lower energy rotational isomers were used, according to eq 3. Unless noted, all calculations were performed using B3LYP/
6-311+G(d,p) b Quantity in kcal mol-1. c Lower energy rotational isomer of IRC-derived reactant was found and eq 3 was used to calculate the
pre-exponential factor and activation energy for the forward reaction. d Value not used in regression of structure/reactivity relationship. e Reaction
thermodynamics and kinetics calculated using CBS-QB3.
(11) Chan, W.; Hamilton, I.; Pritchard, H. J. Chem. Soc., Faraday Trans. (28) Scott, A.; Radom, L. J. Phys. Chem. 1996, 100 (41), 16502-16513.
1998, 94 (16), 2303-2306. (29) Van Speybroeck, V.; Van Neck, D.; Waroquier, M.; Wauters, S.;
(12) Denisova, T. G.; Denisov, E. T. Kinet. Catal. 2001, 42 (5), 620- Saeys, M.; Marin, G. B. J. Phys. Chem. A 2000, 104(46), 10939-10950.
630. (30) Cancés, E.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 1997, 107
(13) Mill, T.; Hendry, D. In ComprehensiVe Chemical Kinetics; Else- (8), 3032-3041.
vier: New York, 1980; Vol. 16, pp 2-87. (31) Rappé, A.; Casewit, C.; Colwell, K.; Goddard, W.; Skiff, W. J.
(14) Sickle, D. V.; Mill, T.; Mayo, F.; Richardson, H.; Gould, C. J. Am. Chem. Soc. 1992, 114 (25), 10024-10035.
Org. Chem. 1973, 38 (26), 4435-4440. (32) McQuarrie, D. A.; Simon, J. D. Molecular Thermodynamics;
(15) Mill, T.; Montosori, G. Int. J. Chem. Kinet. 1973, 5, 119-136. University Science Books: Sausalito, CA, 1999.
(16) Rust, F. J. Am. Chem. Soc. 1957, 79 (15), 4000-4003. (33) Frenkel, M.; Marsh, K. N.; Wilhoi, R. C.; Kabo, G. J.; Roganov,
(17) Merle, J. K.; Hayes, C.; Zalyubovsky, S.; Glover, B.; Miller, T.; G. N. Thermodynamics of Organic Compounds in the Gas State; Thermo-
Hadad, C. J. Phys. Chem. A 2005, 109 (16), 3637-3646. dynamics Research Center: College Station, TX, 1994; Vol. 1.
(18) Gomez-Balderas, R.; Coote, M. L.; Hendry, D.; Radom, L. J. Phys. (34) Sumathi, R.; Carstensen, H.; Green, W. H., Jr. J. Phys. Chem. A
Chem. A 2004, 108 (15), 2874-2883. 2001, 105 (28), 6910-6925.
(19) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, (35) Marston, C. C.; Balint-Kurti, G. G. J. Chem. Phys. 1989, 91 (6),
M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. 3571-3576.
N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; (36) Balint-Kurti, G. G.; Dixon, R.; Marston, C. C. Int. ReV. Phys. Chem.
Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; 1992, 11, 317-344.
Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; (37) Balint-Kurti, G. G.; Ward, C. L.; Marston, C. C. Comput. Phys.
Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, Commun. 1991, 67, 285-292.
X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; (38) Sebbarand, N.; Bockhorn, H.; Bozzelli, J. W. J. Phys. Chem. A
Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; 2005, 109 (10), 2233-2253.
Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; (39) Cauter, K. V.; Speybroeck, V. V.; Vansteenkiste, P.; Reyniers, M.
Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, F.; Waroquier, M. ChemPhysChem 2006, 7, 131-140.
S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. (40) Hirschfelder, J. O.; Wigner, E. J. J. Chem. Phys. 1939, 7 (8), 616-
D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. 628.
G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; (41) Gonzalez-Lafont, A.; Truong, T. N.; Truhlar, D. G. J. Chem. Phys.
Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, 1991, 95 (12), 8875-8894.
M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; (42) Zhang, S.; Truong, T. N. VKLab version 1.0. University of Utah,
Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 2001.
03, revision C.02; Gaussian, Inc.: Wallingford, CT, 2004. (43) Brinck, T.; Lee, H.; Jonsson, M. J. Phys. Chem. A 1999, 103 (35),
(20) Peng, C.; Ayala, P. Y.; Schlegel, H. B. J. Comput. Chem. 1996, 7094-7104.
17 (1), 49-56. (44) Redfern, P. C.; Zapol, P.; Curtiss, L. A.; Raghavachari, K. J. Phys.
(21) Peng, C.; Schlegel, H. B. Israel. J. Chem. 1993, 33, 449-454. Chem. A 2000, 104 (24), 5850-5854.
(22) Gonzalez, C.; Schlegel, H. B. J. Chem. Phys. 1989, 90 (4), 2154- (45) Truong, T. N.; Duncan, W.; Tirtowidjojor, M. Phys. Chem. Chem.
2161. Phys. 1999, 1 (6), 1061-1065.
(23) Becke, A. J. Chem. Phys. 1993, 98 (7), 5648-5652. (46) Henry, D. J.; Radom, L. In Quantum-Mechanical Prediction of
(24) Coote, M. L.; Wood, G. P. F.; Radom, L. J. Phys. Chem. A 2002, Thermochemical Data; Understanding Chemical Reactivity; Cioslowski, J.,
106 (50), 12124-12138. Ed.; Kluwer Academic Publishers: Norwell, MA, 2001; Vol. 22, pp 161-
(25) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. 197.
A. J. Chem. Phys. 1999, 110 (6), 2822-2827. (47) Avila, D. V.; Ingold, K. U.; Lusztyk, J.; Green, W. H., Jr.; Procopio,
(26) Baboul, A. G.; Curtiss, L. A.; Redfern, P. C. J. Chem. Phys. 1999, D. R. J. Am. Chem. Soc. 1995, 117 (10), 2929-2930.
110 (16), 7650-7657. (48) Broadbelt, L. J.; Pfaendtner, J. AIChE J. 2005, 51 (8), 2112-2121.
(27) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K.; Rassolov, V.; (49) Van Geem, K. M.; Reyniers, M. F.; Marin, G. B.; Song, J.; Green,
Pople, J. A. J. Chem. Phys. 1999, 110 (10), 4703-4709. W. H., Jr.; Matheu, D. M. AIChE J. 2006, 52 (2), 718-730.