Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Color profile: Generic CMYK printer profile

Composite Default screen

431

Alkoxy radicals in the gaseous phase: ␤-scission


reactions and formation by radical addition to
carbonyl compounds
Arvi Rauk, Russell J. Boyd, Susan L. Boyd, David J. Henry, and Leo Radom

Abstract: The structures and reactivities of the alkoxy radicals methoxy (CH3O·), ethoxy (CH3CH2O·), 1-propoxy
(CH3CH2CH2O·), 2-propoxy ((CH3)2CHO·), 2-butoxy (CH3CH2CH(CH3)O·), tert-butoxy ((CH3)3CO·), prop-2-enoxy
(CH2=CHCH2O·), and but-3-en-2-oxy (CH2=CHCH(CH3)O·) have been investigated at the B3-LYP/6-31G(d) and CBS-
o ) were calculated with CBS-RAD for all the alkoxy radicals, the
RAD levels of theory. Enthalpies of formation (∆ f H298
carbonyl and radical products of β-scission reactions, and the transition structures leading to them. The mean absolute
o values is 5.4 kJ mol–1. Eyring (∆H ‡, ∆H ‡ , ∆G ‡ )
deviation between the predicted and available experimental ∆ f H298 0 298 298
and Arrhenius (log A, Ea) activation parameters for both the forward (β-scission) and reverse (radical addition to car-
bonyl) pathways were calculated. Agreement with available experimental data is very good, generally within 1–5 kJ
mol–1 for Ea, and 0.5 for log A. The transition structures are found to be substantially polarized, with the departing
radical slightly positive, the O atom negative, and the rest of the molecule positive. The barriers for the β-scission reac-
tions decrease with decreasing endothermicity and with decreasing ionization energy of the departing radical.

Key words: alkoxy, alkoxyl, radical, addition, carbonyl, β-scission, calculaton, electronic structure, B3LYP, CBS-RAD,
thermochemistry.

Résumé : Faisant appel à des calculs théoriques aux niveaux B3-LYP/6-31G(d) et CBS-RAD de la théorie, on a étudié
les structures et les réactivités des radicaux alkoxy, méthoxy (CH3O·), éthoxy (CH3CH2O·), 1-propoxy (CH3CH2CH2O·),
2-propoxy [(CH3)2CHO·], 2-butoxy [CH3CH2CH(CH3)O·], tert-butoxy [(CH3)3CO·], prop-2-énoxy (CH2=CHCH2O·) et
but-3-én-2-oxy [CH2=CHCH(CH3)O·]. On a calculé les enthalpies de formation, ∆ f H298 o , au niveau CBS-RAD de la
théorie pour tous les radicaux alkoxy, tous les produits carbonylés et tous les produits radicalaires provenant de réac-
tions de β-scission et toutes les structures de transition qui y conduisent. La déviation absolue moyenne entre les
o expérimentales disponibles est de 5,4 kJ mol–1. On a aussi calculé les paramètres
valeurs prédites et les valeurs de ∆ f H298
d’activation d’Eyring (∆H0 , ∆H298 , ∆G298
‡ ‡ ‡
) et d’Arrhenius (log A, Ea) pour la réaction vers la droite (β-scission) et pour
la réaction inverse (addition d’un radical sur le carbonyle). L’accord entre les valeurs calculées et les valeurs expéri-
mentales disponibles est bon, généralement entre 1 et 5 kJ mol–1 pour les valeurs de Ea et de 0,5 pour le log A. On a
trouvé que les structures de transition sont assez polarisées alors que le radical qui se détache est légèrement positif,
l’atome d’oxygène est négatif et que le reste de la molécule est positif. Les barrières aux réactions de β-scission dimi-
nuent avec une augmentation du caractère endothermique et avec une diminution de l’énergie d’ionisation du radical
qui se détache.

Mots clés : alkoxy, alkoxyle, radical, addition, carbonyle, scission-β, calculs théoriques, structure électronique, B3LYP,
CBS-RAD, thermochimie.

[Traduit par la Rédaction] Rauk et al. 442

Received 24 October 2002. Published on the NRC Research Press Web site at http://canjchem.nrc.ca on 9 April 2003.
Dedicated to Professor Don Arnold for his contributions to chemistry.
A. Rauk.1 Department of Chemistry, University of Calgary, Calgary, AB T2N 1N4, Canada.
R.J. Boyd. Department of Chemistry, Dalhousie University, Halifax, NS B3H 4J3, Canada.
S.L. Boyd. Department of Chemistry, Mount St. Vincent University, Halifax, NS B3M 2J6, Canada.
D.J. Henry and L. Radom. Research School of Chemistry, Australian National University, Canberra, ACT 0200, Australia.
1
Corresponding author (e-mail: rauk@ucalgary.ca).

Can. J. Chem. 81: 431–442 (2003) doi: 10.1139/V02-206 © 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:38 PM
Color profile: Generic CMYK printer profile
Composite Default screen

432 Can. J. Chem. Vol. 81, 2003

Introduction and computationally, by means of ab initio calculations, by


Viskolcz and co-workers (14–18), and others (19, 20). Their
Alkoxy radicals (R1R2R3C–O·) are important intermediate work is discussed in the present context below.
species in the combustion of hydrocarbons (1, 2), and in at-
mospheric chemistry (3–5). They can react by intramolecular
hydrogen abstraction, or by intermolecular hydrogen abstrac- Theoretical methods
tion from a suitable donor (R′–H) to yield the corresponding Quantum mechanical calculations
alcohol: Conventional ab initio (21) and density functional theory
[1] R1R2R3C-O· + R′-H → R1R2R3C-OH + R′· (22) calculations were carried out with the Gaussian 94
(23a), Gaussian 98 (23b), and Molpro 2000 (24) suites of
Alternatively, alkoxy radicals can undergo β-cleavage to electronic structure codes. Geometries and harmonic vibra-
yield an alkyl radical plus a ketone or aldehyde in up to tional frequencies were obtained for all species at the B3-
three ways: LYP/6-31G(d) level of theory. The vibrational frequencies
[2a] R1R2R3C-O· → R1· + R2R3C=O were used, after appropriate scaling (25), in the calculation
of thermodynamic and kinetic parameters including zero-
[2b] → R2· + R1R3C=O point vibrational energies (ZPVEs), and to characterize each
stationary point as a local minimum (no imaginary frequen-
[2c] → R3· + R1R2C=O
cies), or as a transition structure (one imaginary frequency).
In proteins, production of such radicals by oxidative pro- Higher-level energies were determined for all structures by
cesses may lead to protein backbone cleavage (6). In gen- the CBS-RAD procedure, which has been developed to pro-
eral, the pathway that leads to the most stable radical is vide accurate thermochemistry for radical species (26). The
found to dominate the endothermic cleavage (7, 8) but ex- CBS-RAD(B3-LYP) combination employed here has been
ceptions have been noted in polycyclic systems (9) and have found to yield good reaction enthalpies and barriers for the
been investigated computationally (10). addition of C-centered radicals to alkenes (13, 27).
The reverse reaction, alkyl radical addition to the carbonyl
double bond: Thermochemical procedures
Gas-phase thermodynamic functions, including thermal
[3] R1· + R2R3C=O → R1R2R3C-O· enthalpy contributions (H° – H0o ) (where H° refers to
is hindered by an activation energy (Ea(3)) which has empiri- 298.15 K) and entropies (S) may be accurately calculated
cally been found to depend on the ionization energy of the by application of standard statistical procedures (28).
radical (IE(R1·)) and the exothermicity of the reaction Values derived experimentally are available in standard
(∆H(3)) (11): sources (29–32).
Theoretical enthalpies of formation at 298.15 K for spe-
[4] Ea(3) = 8.8[IE(R1·)] – 0.58|∆H(3)| – 26 cies CaHbO may be calculated from the energies of the at-
where the ionization energy is in units of eV and the omization reactions:
enthalpy of the reaction is in kJ mol–1. The relationship with [5] aC + bH + O → CaHbO
the ionization energy was attributed (11) to a tight transition
structure with substantial polarization in the sense: by eqs. [6]–[8]:

[6] o (C H O) = ∆H + a∆ H o (C)
∆ f H298 a b (5) f 298
o (H) + ∆ H o (O)
+ b∆ f H298 f 298

where
[7] ∆H(5)(298.15 K) = ∆H(5)(0 K)
Additional evidence for charge separation in the TS may
be deduced from the observation of a large solvent effect for + (H° – H0o )(CaHbO) – a(H° – H0o )(C)
the β-scission of tert-butoxy radical (12). In a similar man- – b(H° – H0o )(H) – (H° – H0o )(O)
ner, barriers for the addition of alkyl and related radicals to
alkenes have also been found to be dependent on the reac- and
tion enthalpy and the ionization energy of the radical but a
multiplicative-type relationship is found to better describe [8] ∆H(5)(0 K) = ECBS(CaHbO) – aECBS(C)
this dependence (13). – bECBS(H) – ECBS(O)
We investigate here the factors affecting both the decom-
position and radical-addition pathways for a systematic se- o ) and
In this procedure, the enthalpies of formation (∆ f H298
ries of saturated and unsaturated alkoxy radicals, by using o
the (H° – H0 ) values for the gaseous atoms in their lowest
high-level quantum mechanical methods. A number of electronic states are taken from experiment (29) (kJ mol–1): C
alkoxy radicals, including some of the systems of interest (716.7, 6.535), H (218.0, 6.197), O (249.2, 6.724). The
here, have been previously studied both experimentally, by CBS-RAD energies (ECBS(CaHbO)) for all species discussed
means of laser flash photolysis – laser-induced fluorescence, here are listed in Table 1. For the gaseous atoms in their

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:38 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Rauk et al. 433

o – H o (kJ mol–1), S (J K–1 mol–1), 298.15 K), and


Table 1. CBS-RAD energies (0 K, hartrees), thermodynamic functions (H298 0
enthalpies of formation (∆ f H o (kJ mol–1), 298.15 and 0 K).
S(g) o c
∆ f H298 ∆ f H0o
Species CBS-RADa nb o – Ho
H298 Exptl. Calcd. Exptl. Calcd. Calcd.
0
–0.49982 6.2 d
H· 1 114.7 114.6 218.0
CH3· –39.74402 1 10.6 194.2d 195.4 146.4 150.0 153.1
CH3CH2· –78.96993 1 12.9 252.0d 255.7 120.9 127.8 138.2
CH2=CH· –77.74177 1 10.6 233.6 300.0 301.9 306.1
CH3O· –114.87243 1 10.4 234.0e,f 17.2, 21g 21.0 28.7
CH3OH –115.53807 1 11.1 239.7d 237.7 –201.0 –201.8 –190.6
CH2=O –114.34253 1 10.0 218.8d 218.6 –108.6 –112.2 –108.3
H···CH2O·TS –114.83531 1 11.5 243.9 119.6 126.2
CH3CH2O· –154.10201 1 13.5 278.9e –15.5, –12g –9.9 4.2
h
CH3CH=O –153.58023 1 12.7 264.2 262.3 –166.2 –164.8 –154.2
H···CH(CH3)O·TS –154.06979 2 14.0 282.0 75.2 88.8
CH3···CH2O· TS –154.07668 1 14.9 285.9 58.0 70.7
CH3CH2CH2O· –193.32697 3 16.1 316.3e –41.4 –29.3 –8.2
CH3CH2CH=O –192.80566 3 15.6 304.7h 301.5 –185.6 –185.2 –167.8
H···CH(CH2CH3)O·TS –193.29559 3 17.1 316.4 54.1 74.2
CH3CH2···CH2O·TS –193.30643 3 17.8 326.6 26.4 45.8
(CH3)2CHO· –193.33181 1 16.1 305.4e –52.3, –46g –41.9 –20.9
(CH3)2C=O –192.81740 1 15.7i 294.9 h
299.5 –217.1 –215.9 –198.6
H···C(CH3)2O· TS –193.30390 1 17.0 304.9 32.2 52.4
CH3···CH(CH3)O·TS –193.31051 2 17.8 319.7 15.7 35.0
CH3CH2CH(CH3)O· –232.55793 6 19.6 351.4e –69.5 –63.4 –36.4
CH3CH2C(CH3)=O –232.04323 1 19.8 338.1h 337.1 –238.5 –235.9 –213.3
H···C(CH3)(CH2CH3)O·TS –232.53034 6 20.3 350.5 9.7 36.1
CH3···CH(CH2CH3)O·TS –232.53579 2 21.2 351.8 –3.6 21.8
CH3CH2···CH(CH3)O·TS –232.54080 6 21.1 360.7 –16.9 8.6
(CH3)3CO· –232.56511 1 19.6 322.6e,f –90.8, –83g –82.2 –55.2
CH3···C(CH3)2=O·TS –232.54493 1 21.1 338.3 –27.8 –2.2
CH2=CHCH2O· –192.11518 2 14.7 303.8e 102.8 116.7
CH2=CHCH=Oj –191.59751 1 13.9 278.0 –77.0k –62.9 –52.4
H···CH(CH=CH2)O·TS –192.08709 2 15.6 300.2 177.5 190.5
CH2=CH···CH2O·TS –192.08108l 3 17.6 328.7 195.2 206.3
CH2=CHCH(CH3)O· –231.34649 6 18.2 342.8e 67.6 87.7
CH2=CHC(CH3)=O –230.83280 1 17.7 313.4 –138.0k –108.2 –92.0
H···C(CH3)(CH=CH2)O·TS –231.32022 6 18.9 338.3 137.3 156.6
CH3···CH(CH=CH2)O·TS –231.32863l 6 19.5 346.6 115.9 134.6
CH2=CH···CH(CH3)O·TS –231.31577l 6 20.6 362.1 150.7 168.3
a
Energies for transition structures correspond to CBS-RAD–Scanmax values (see Table 2 and text) unless otherwise noted.
b
Number of significantly populated conformations (used to calculate entropies of mixing).
c
From ref. 32 unless otherwise noted.
d
Reference 29.
e
First excited state assumed to be significantly populated (adds R ln 2).
f
Rotational symmetry number = 3 assumed.
g o values from ref. 36.
∆ f H 300
h
Reference 34.
i
One methyl group rotation treated as a free rotation (contributes RT/2).
j
Energies refer to s-trans conformation.
k
Reference 30.
l
Standard CBS-RAD value.

lowest electronic states, the CBS-RAD energies are ECBS for convenience in Table 1 the enthalpies of formation for
(hartrees): C (–37.785143), H (–0.499818), O (–74.987034). the transition structures, even though the ∆fH would be a
Calculated enthalpies of formation are compared with somewhat artificial quantity for such species. In cases where
available experimental values in Table 1. We have included a substance exists as an equilibrium mixture of two or more

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:38 PM
Color profile: Generic CMYK printer profile
Composite Default screen

434 Can. J. Chem. Vol. 81, 2003

conformers, the enthalpy of formation listed in Table 1 cor- properties of the transition structures are presented in Ta-
responds to that of the most stable conformer. ble 2. Table 3 gives the kinetic parameters (∆H‡, ∆G‡, log A,
The entropy of a pure substance in the gas phase may be and Ea) and the reaction enthalpies (∆H) for β-scission reac-
expressed as a weighted sum of the entropies of individual tions (eq. [2]), while corresponding data for the reverse reac-
conformers, plus an entropy of mixing term (eq. [9]): tion, radical addition to carbonyl (eq. [3]), are collected in
Table 4.
[9] Stotal = ΣixiSi – RΣixi ln xi ≅ S1 + R ln n
where xi is the mol fraction of conformer i, S1 is the entropy Enthalpies of formation
of the most stable conformation, and n denotes the number Enthalpies of formation, calculated from atomization ener-
of significantly populated conformations (33). gies as described above, are listed in Table 1. Comparison with
The alkoxy radicals investigated in this study exhibit experimental values may be made for 17 species. The largest
Jahn–Teller distortions. We have accounted for the near de- discrepancy occurs for but-3-en-2-one (CH2=CHC(CH3)=O) for
o = –108.2 kJ mol–1) is
generacy of the A′ and A′ ′ states (Cs symmetry) by the addi- which the calculated value (∆ f H298
–1
29.8 kJ mol less negative than the experimental (∆ f H298 o =
tion of R ln 2 to the electronic entropy. In addition, for the
–1
methoxy and tert-butoxy radicals we have adopted a rota- –138.0 kJ mol ). For the remaining 16 species, the mean
tional symmetry number of 3. Calculated entropies are in- absolute deviation is 3.8 kJ mol–1, consistent with the accu-
cluded in Table 1, together with experimental values where racy expected from CBS-RAD (26). The second-largest dis-
available. crepancy (+14.1 kJ mol–1) also occurs for an α,β-unsaturated
carbonyl compound, acrolein ((E)-CH2=CHCH=O). The cal-
Rates of reaction culated values for the heat of formation of the methoxy,
In the thermodynamic formulation of transition-state the- ethoxy, 2-propoxy, and tert-butoxy radicals are in close
ory (TST) (34), the rate constant (kT) for a reaction at tem- agreement with the recent experimental results of Ramond et
perature T is related to the free energy barrier (∆G‡) by al. (36), and support these values over the older results (32)
eq. [10] (35): (which are generally less positive or more negative). Larger-
than-average differences between theory and experiment oc-
[10] kT = (kBT/h)(c°)1 – m exp(–∆G‡/RT) cur for two alkoxy radicals 1-propoxy (CH3CH2CH2O·,
where kB and h are Boltzmann’s and Planck’s constants, re- +12.1 kJ mol–1) and 2-butoxy (CH3CH2CH(CH3)O·, +6.1 kJ
spectively, R is the ideal gas constant, c° is the standard unit mol–1), and for the ethyl radical (CH3CH2·, +6.9 kJ mol–1).
of concentration (mol L–1), and m is the order of the reac- Experimental reexamination may be warranted in these in-
tion. The factor ((c°)1 – m) provides the correct units for the stances. We are not aware of any experimental enthalpies of
rate constant of unimolecular (m = 1) and bimolecular (m = formation for the prop-2-en-1-oxy (CH2=CHCH2O·) or but-
2) reactions. In the gas-phase standard state (T = 298.15 K) 3-en-2-oxy (CH2=CHCH(CH3)O·) radicals.
the pressure is 1 atm and c° = 0.0408 mol L–1. We apply
eq. [10] for the calculation of both the (forward) Sensitivity of calculated barriers to TS geometry and
unimolecular β-scission reaction (eq. [2]) and the (reverse) theoretical level
bimolecular radical-addition reaction (eq. [3]). The barriers for the β-scission (forward, eq. [2]) and radi-
In Arrhenius rate theory, the rate constant is given by: cal addition (reverse, eq. [3]) reactions are crucial for deter-
mining the kinetic fate of alkoxy radicals in atmospheric
[11] kT = A exp(–Ea/RT) chemistry and for understanding the combustion of hydro-
carbons. Columns 4 and 5 of Table 2 compare the B3-
The Arrhenius activation energy (Ea) is given by: LYP/6-31G(d) and CBS-RAD barriers (∆H0‡ ) for β-scission
[12] Ea = ∆H‡ + mRT of a number of the alkoxy radicals, in both cases using ge-
ometries and ZPVEs obtained at the B3-LYP/6-31G(d) level.
The Arrhenius frequency factor (A) contains the entropy It is evident that the B3-LYP values are significantly higher
of activation (∆S‡): than the CBS-RAD values, by 13–20 kJ mol–1 for C-H
[13] A = (kBT/h)(c°)1 – mem exp(∆S‡/R) scission, and by 3–14 kJ mol–1 for C-C scission. By either
method, methyl-radical loss or ethyl-radical loss has a sig-
Enthalpy barriers (∆H‡ (generally referred to in the course nificantly lower barrier than H-atom loss in species where
of this paper simply as barriers)) are calculated directly from both channels are available.
enthalpies of formation of the appropriate species (or equiva- The CBS-RAD procedure is expected to produce reason-
lently from the calculated total energies and, if appropriate, ably reliable estimates of the barriers provided that the ge-
the enthalpy temperature corrections). In the present study, we ometries of the species derived at the B3-LYP/6-31G(d)
have not considered quantum mechanical tunneling. This level are adequate. The results should be most sensitive to
could in principle be important in the reactions involving hy- the geometries of the transition structures. We have exam-
drogen-atom scission and (or) addition but preliminary calcu- ined the appropriateness of the B3-LYP/6-31G(d) methodol-
lations suggest that the effect is not particularly significant. ogy for determining the transition structures by calculating
the CBS-RAD energy at points along the bond-breaking co-
Results and discussion ordinate for most of the β-scission reactions. The breaking
C···H or C···C bond is stepped in 0.02 Å increments from the
The energies of all species considered here are given in B3-LYP/6-31G(d) transition structure with reoptimization of
Table 1. The variation with the theoretical level of selected all the remaining geometric parameters. The zero-point vi-

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:39 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Rauk et al. 435

Table 2. Dependence of transition structure bond length (r, Å) and reaction barrier (∆H0‡, kJ mol–1) on theoretical level.
B3-LYP/6-31G(d) CBS-RADb CBS-RAD–Scanmaxc Differences
Alkoxy radical Bond a
r ∆H0‡ ∆H0‡ r ∆H0‡ ∆rd ∆H0‡e
R1R2CHO· → H· + R1R2C=O
CH3O· C···H 1.956 114.9 95.1 1.80 97.5 –0.16 2.4
CH3CH2O· C···H 1.852 100.6 82.6 1.72 84.6 –0.13 2.0
CH3CH2CH2O· C···H 1.851 98.4 80.6 1.72 82.4 –0.13 1.8
(CH3)2CHO· C···H 1.795 87.1 71.2 1.67 73.2 –0.13 2.0
CH3CH2CH(CH3)O· C···H 1.797 85.6 70.8 1.70 72.4 –0.10 1.6
CH2=CHCH2O· C···H 1.803 85.4 71.8 1.70 73.8 –0.10 2.0
CH2=CHCH(CH3)O· C···H 1.763 80.3 67.1 1.66 69.0 –0.10 1.9
R1R2(CH3)CO· → CH3· + R1R2C=O
CH3CH2O· C···C 2.250 79.2 65.2 2.15 66.5 –0.10 1.3
(CH3)2CHO· C···C 2.182 65.2 55.0 2.10 55.9 –0.08 0.9
CH3CH2CH(CH3)O· C···C 2.192 65.2 57.4 2.13 58.1 –0.06 0.7
(CH3)3CO· C···C 2.148 59.2 52.2 2.09 53.0 –0.06 0.8
R1R2(CH3CH2)CO· → CH3CH2· + R1R2C=O
CH3CH2CH2O· C···C 2.227 61.8 52.9 2.15 53.9 –0.08 1.0
CH3CH2CH(CH3)O· C···C 2.153 47.2 44.6 2.11 45.0 –0.04 0.4
a
Bond being broken in TS.
b
Calculated using B3-LYP/6-31G(d) geometries and (scaled) zero-point vibrational energies.
c
Calculated using the CBS-RAD–Scanmax procedure (see text); ZPVEs obtained with B3-LYP/6-31G(d) at the B3-LYP/6-31G(d) TS geometry.
d
CBS-RAD–Scanmax – B3-LYP/6-31G(d).
e
CBS-RAD–Scanmax – CBS-RAD.

brational energy is held constant at the TS value. The geom- zero-point energy correction (10). The values so obtained
etry at which the CBS-RAD energy, obtained in this manner, are uniformly 30–50 kJ mol–1 higher than the values pre-
is a maximum, is an approximation to the CBS-RAD transi- sented in Tables 3 and 4. In another study of β-scission reac-
tion structure, while the corresponding ∆H0‡ represents an tions (39), AM1 theory was used to calculate barriers for a
improved CBS-RAD estimate of the barrier. This procedure number of alkoxy radicals, only one of which is common to
has been referred to as Scanmax (37) and is closely related the present study. The AM1 barrier for the decomposition of
to the IRCMax procedure of Petersson and co-workers (38). the tert-butoxy radical decomposition is 13 kJ mol–1 higher
Table 2 lists the length (r) of the bond being broken in the than the CBS-RAD value (Table 3).
TS as obtained with B3-LYP/6-31G(d), the predicted B3-
LYP/6-31G(d), and CBS-RAD values for ∆H0‡ at the B3- Individual alkoxy radicals
LYP/6-31G(d) optimized geometry, and the values of r and We discuss some of the individual alkoxy radicals prior to
∆H0‡ obtained using the CBS-RAD–Scanmax procedure. In examination of the reactivity trends.
each case, the CBS-RAD–Scanmax energy is a maximum at
a value of r shorter than the B3-LYP/6-31G(d) optimized Methoxy radical
value, leading to an increase in the predicted ∆H0‡ . The dif- The CBS-RAD value of the bond dissociation enthalpy
ferences are listed in the last two columns of Table 2. (BDE) of the O—H bond of methanol may be derived from
Compared with CBS-RAD–Scanmax, B3-LYP/6-31G(d) the data in Table 1. The values at 0 and 298 K are 435.4 and
overestimates the length of the breaking C—H bonds in the 440.8 kJ mol–1, respectively, in good agreement with experi-
TS by 0.10–0.16 Å, leading to an underestimation of ∆H0‡ mental values 431 ± 4 and 436 ± 4 kJ mol–1, respectively
by the standard CBS-RAD procedure by about 1.6–2.4 kJ (40). Our calculated enthalpies of formation of the methoxy
mol–1. For C—C bond cleavage the corresponding differ- radical are ∆ f H0o = 28.7 kJ mol–1 and ∆ f H298
o = 21.0 kJ mol–1.
ences in r and ∆H0‡ are somewhat smaller, 0.04–0.10 Å and Berkowitz et al. (40) reported ∆ f H0o = 24.7 ± 3.8 kJ mol–1
0.4–1.3 kJ mol–1, respectively. Where available, we use the o = 17.2 ± 3.8 kJ mol–1 while Ramond et al. (36)
and ∆ f H298
CBS-RAD–Scanmax values of ∆H ‡ rather than standard o = 21 ± 4 kJ mol–1. Osborn et al. (41) re-
reported ∆ f H300
CBS-RAD values in the remainder of this paper. o
ported ∆ f H0 = 28.5 ± 1.7 kJ mol–1, in close agreement with
We note that a detailed analysis of the dependence of the our calculated value and also with the recent high-level theo-
results for some of the alkoxy radicals and their C-C uni- retical value of Petraco et al. (19) (∆ f H0o = 27.2 kJ mol–1).
molecular decomposition reaction characteristics on theoreti- The methoxy radical is predicted to undergo β-scission of
cal level and basis set, at levels comparable to those a C—H bond with an Arrhenius activation energy (Ea =
employed here, may be found in the study by Somnitz and 101.1 kJ mol–1 (Table 3)) in close agreement with the value
Zellner (20). Their work is discussed further below. We note proposed by Dibble (42) (100 ± 4 kJ mol–1) on the basis of
also that for most of the alkoxy radicals studied here, the theoretical and experimental results. The length of the break-
barriers for the β-cleavage reactions have been previously ing C···H bond in the transition structure is 1.80 Å (Table 2),
determined at the CASSCF/6-31G(d) level of theory, without close to that found by Petraco et al. (19) (1.79 Å). The rate

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:39 PM
Table 3. Enthalpy barriers (∆H‡, kJ mol–1), free energy barriers (∆G‡, kJ mol–1), frequency factors (A, s–1), activation energies (Ea, kJ mol–1), rate constants (s–1, 298.15 K),
436
reaction enthalpies (∆H, kJ mol–1) and reaction free energies (∆G) for β-scission reactions calculated with CBS-RAD and B3-LYP/6-31G(d).a
‡ ‡
Alkoxy radical Reaction ∆H298 ∆G298 log A Ea kb ∆H0 ∆H298 ∆G298

June 9, 2003 2:48:40 PM


55.2

I:\cjc\cjc8106\V02-206.vp
Composite Default screen

Methoxy CH3O· → H· + CH2=O 98.6 95.7 13.74 101.1 1.1 × 10–5 79.0 84.8
Ethoxy CH3CH2O· → H· + CH3CH=O 85.1 84.2 13.39 87.6 1.1 × 10–2 57.7 63.1 33.3
1-Propoxy CH3CH2CH2O· → H· + CH3CH2CH=O 83.4 83.3 13.23 85.8 1.6 × 10–2 56.4 62.1 32.3
2-Propoxy (CH3)2CHO· → H· + (CH3)2C=O 74.1 74.3 13.20 76.6 6.0 × 10–1 38.3 44.1 13.0
2-Butoxy CH3CH2CH(CH3)O· → H· + CH3CH2C(CH3)=O 73.1 73.4 13.18 75.6 8.6 × 10–1 39.1 45.5 15.6
Prop-2-en-1-oxy CH2=CHCH2O· → H· + (E)-CH2=CHCH=O 74.7 75.7 13.04 77.2 3.3 × 10–1 46.9 52.3 25.8
Color profile: Generic CMYK printer profile

But-3-en-2-oxy CH2=CHCH(CH3)O· → H· + (E)-CH2=CHC(CH3)=O 69.7 70.9 12.99 72.2 2.3 36.4 42.1 16.7
Ethoxy CH3CH2O· → CH3· + CH2=O 67.9 65.8 13.60 70.4 1.9 × 10 40.6 47.8 7.8
2-Propoxy (CH3)2CHO· → CH3· + CH3CH=O 57.7 53.4 13.96 60.1 2.8 × 103 19.8 27.1 –18.5
2-Butoxy CH3CH2CH(CH3)O· → CH3· + CH3CH2CH=O 59.8 59.6 13.25 62.3 2.2 × 102 21.7 28.3 –14.8
tert-Butoxy (CH3)3CO· → CH3· + (CH3)2C=O 54.5 49.7 14.06 56.9 1.2 × 104 9.7 16.4 –33.3
But-3-en-2-oxy CH2=CHCH(CH3)O· → CH3· + (E)-CH2=CHCH=O 48.2 47.1 13.43 50.7 3.6 × 104 13.0 19.4 –19.2
1-Propoxy CH3CH2CH2O· → CH3CH2· + CH2=O 55.7 52.6 13.77 58.1 3.8 × 103 38.1 44.9 –1.2
2-Butoxy CH3CH2CH(CH3)O· → CH3CH2· + CH3CH=O 46.5 43.7 13.72 48.9 1.4 × 105 20.4 26.4 –22.8
Prop-2-en-1-oxy CH2=CHCH2O· → CH2=CH· + CH2=O 92.4 85.0 14.53 94.8 7.9 × 10–3 81.0 87.0 42.7
But-3-en-2-oxy CH2=CHCH(CH3)O· → CH2=CH· + CH3CH=O 83.1 77.3 14.24 85.6 1.8 × 10–1 64.3 69.5 23.2
a
See text for details.
b
298.15 K.

Table 4. Enthalpy barriers (∆H‡, kJ mol–1), free energy barriers (∆G‡, kJ mol–1), frequency factors (A, L mol–1 s–1), activation energies (Ea, kJ mol–1), rate constants
(L mol–1 s–1, 298.15 K), reaction enthalpies (∆H, kJ mol–1), and reaction free energies (∆G) for radical addition reactions calculated with CBS-RAD and B3-LYP/6-31G(d).a
‡ ‡
Radical IEb Reaction ∆H298 ∆G298 log A Ea kc ∆H0 ∆H298 ∆G298
H· 13.6 H· + CH2=O → CH3O· 13.8 40.5 10.37 18.7 1.2 × 107 –79.0 –84.8 –55.2
H· + CH3CH=O → CH3CH2O· 22.0 50.9 9.99 27.0 1.8 × 105 –57.7 –63.1 –33.3
H· + CH3CH2CH=O → CH3CH2CH2O· 21.3 51.0 9.84 26.2 1.8 × 105 –56.5 –62.1 –32.3
H· + (CH3)2C=O → (CH3)2CHO· 30.1 61.3 9.58 35.0 2.8 × 103 –38.3 –44.1 –13.0
H· + CH3CH2C(CH3)=O → CH3CH2CH(CH3)O· 27.6 57.8 9.76 32.6 1.1 × 104 –39.1 –45.5 –15.6
H· + (E)-CH2=CHCH=O → CH2=CHCH2O· 22.4 50.0 10.22 27.4 2.7 × 105 –46.9 –52.3 –25.9
H· + (E)-CH2=CHC(CH3)=O → CH2=CHCH(CH3)O· 27.5 54.3 10.36 32.5 4.6 × 104 –36.4 –42.1 –16.7
Methyl 9.80 CH3· + CH2=O → CH3CH2O· 20.1 58.0 8.41 25.1 1.1 × 104 –40.6 –47.8 –7.8
CH3· + CH3CH=O → (CH3)2CHO· 30.6 71.9 7.81 35.5 3.8 × 10 –19.8 –27.1 18.5
CH3· + CH3CH2CH=O → CH3CH2CH(CH3)O· 31.5 74.4 7.53 36.5 1.4 × 10 –21.7 –28.3 14.8
CH3· + (CH3)2C=O → (CH3)3CO· 38.1 83.0 7.18 43.0 4.4 × 10–1 –9.7 –16.4 33.3
CH3· + (E)-CH2=CHCH=O → CH2=CHCH(CH3)O· 28.8 66.3 8.49 33.8 3.8 × 102 –13.0 –19.4 19.3
Ethyl 8.12 CH3CH2· + CH2=O → CH3CH2CH2O· 10.7 53.8 7.52 15.7 5.9 × 104 –38.1 –44.9 1.2
CH3CH2· + CH3CH=O → CH3CH2CH(CH3)O· 20.1 66.4 6.93 25.0 3.5 × 102 –20.4 –26.4 22.8
Vinyl 8.25 CH2=CH· + CH2=O → CH2=CHCH2O· 5.4 42.3 8.59 10.4 5.9 × 106 –81.0 –86.9 –42.8
CH2=CH· + CH3CH=O → CH2=CHCH(CH3)O· 11.6 52.0 7.97 16.5 1.2 × 105 –64.3 –69.5 –23.2
a
See text for details.
b

© 2003 NRC Canada


Can. J. Chem. Vol. 81, 2003

Ionization energy of the radical (in eV).


c
298.15 K.
Color profile: Generic CMYK printer profile
Composite Default screen

Rauk et al. 437

constant for unimolecular decomposition at 298 K is pre- nificantly lower than ours (13.60). Their predicted rate con-
dicted to be 1.1 × 10–5 s–1. stant at 300 K (1.46 s–1) is in close agreement with the value
The β-scission reaction (eq. [3]) is highly endothermic for of Batt and Milne (44). β-Scission of the ethoxy radical
the methoxy radical and the enthalpy change corresponds to yielding CH3· + CH2=O is predicted to be endothermic by
the bond dissociation enthalpy (BDE) of the C—H bond 40.6 kJ mol–1 at 0 K and 47.8 kJ mol–1 at 298 K (Table 3).
(∆H0 = 79.0 kJ mol–1 and ∆H298 = 84.8 kJ mol–1 (Table 3)). The experimentally determined activation energy for the
Our calculated 0 K value is slightly lower than that deduced addition of the methyl radical to formaldehyde is 31.0 kJ
experimentally by Osborn et al. (41) (82.8 ± 1.7 kJ mol–1) mol–1 (11), somewhat higher than that found in the present
and that calculated by Petraco et al. (19) (84.1 kJ mol–1). calculations (Ea = 25.1 kJ mol–1) (Table 4).
The reverse reaction, addition of H atom to formaldehyde,
has a predicted barrier height at 0 K (∆H0‡ ) of 18.5 kJ mol–1, 1-Propoxy radical
in good agreement with the recommended theoretical value The predicted values for the enthalpy of formation of
of Petraco et al. (19) (19.7 kJ mol–1). The reverse barrier is the 1-propoxy radical are ∆ f H0o = –8.2 kJ mol–1 and
considerably smaller than the forward barrier because of the o = –29.3 kJ mol–1, considerably less negative than the
∆ f H298
endothermicity of the β-scission reaction. Our Arrhenius pa- experimental value (∆ f H298o = –41.4 kJ mol–1) (Table 1). In
rameters are log A = 10.37 and Ea = 18.7 kJ mol–1. The Ea the light of the good agreement with recent experimental data
value is somewhat lower than values reported by Walch (43) (36) observed for other alkoxy radicals, we believe that exper-
(22 and 26 kJ mol–1), derived experimentally from rates for imental reexamination for 1-propoxy radical is desirable.
deuterium addition at room temperature (log A = 10.30 was β-Scission of the 1-propoxy radical yielding CH3CH2-

assumed) and an RRKM analysis, respectively. Because of CH=O + H· has a predicted barrier (∆H298 ) of 83.4 kJ mol–1
the large negative entropy of activation calculated for the ad- (Table 3). This value is only slightly lower than that predicted
dition reaction (–89.6 J mol–1 K–1) the free energy barrier for β-scission of the H atom from the ethoxy radical (∆H298 ‡
=
(∆G298

= 40.5 kJ mol–1) is substantially higher than the 85.1 kJ mol ). The 0 K value (∆H0 = 82.4 kJ mol , Table 2)
–1 ‡ –1

enthalpy barrier (∆H298 = 13.8 kJ mol–1). Our calculated rate is 13.1 kJ mol–1 lower than that calculated by Viskolcz and
for the addition of H atom to formaldehyde is k = 1.2 × co-workers (14) at the B3-LYP/SVP level of theory. The
107 L mol–1 s–1. higher value of Viskolcz is consistent with our observations
that the B3-LYP procedure tends to yield higher values for β-
Ethoxy radical scission barriers than CBS-RAD (Table 2). The present results
The calculated values for the enthalpy of formation of the predict that the β-scission reaction of the 1-propoxy radical to
ethoxy radical are ∆ f H0o = 4.2 kJ mol–1 or ∆ f H298
o = –9.9 kJ give CH3CH2CH=O + H· is strongly endothermic by 62.1 kJ
–1
mol (Table 1). These are in fair agreement with the values mol–1 (∆H298).
recommended by Berkowitz et al. (40) (∆ f H0o = –1.7 ± The predicted barrier for addition of the H atom to propanal
3.8 kJ mol–1 and ∆ f H298o = –15.5 ± 3.8 kJ mol–1, respec- is 21.3 kJ mol–1 (Table 4) which, not surprisingly, is close to
tively). We find closer agreement with the experimental the value for the addition to acetaldehyde (22.0 kJ mol–1).
value of Ramond et al. (36) (∆ f H300o = –12 ± 7 kJ mol–1). The present calculations predict that β-scission of the 1-
Viskolcz and co-workers (15) used a variety of high theoreti- propoxy radical yielding CH3CH2· + CH2=O, is impeded by
o = 0 ± 3 kJ mol–1, significantly
cal levels to obtain ∆ f H298 a barrier (∆H298‡
) of 55.7 kJ mol–1 (Table 3) and is endother-
higher than our CBS-RAD value and the experimental mic by 44.9 kJ mol–1 (∆H298). This reaction has also been in-
values. vestigated by Somnitz and Zellner (20). Their RRKM-
Our calculated barrier for β-scission of the H atom from derived Ea of 62.4 kJ mol–1 is slightly higher than ours

the ethoxy radical (∆H298 = 85.1 kJ mol–1) is close to that (58.1 kJ mol–1) and, as in the case of ethoxy, their log A
obtained by Viskolcz and co-workers (15) (84.8 kJ mol–1) at value (13.25) is somewhat lower than ours (13.77).

the QCISD(T)/6-311+G(3df,2p) level of theory. It is approx- Our predicted barrier (∆H298 ) for addition of ethyl radical to
imately 14 kJ mol–1 lower than for H-atom scission from the formaldehyde is 10.7 kJ mol–1 (Table 4), approximately 9 kJ

methoxy radical. mol–1 lower than that for methyl-radical addition (∆H298 =
–1 –1
The barrier for the reverse reaction, addition of the H 20.1 kJ mol ) and approximately 3 kJ mol lower than that
‡ ‡
atom to acetaldehyde (∆H298 = 22.0 kJ mol–1), is somewhat for H-atom addition (∆H298 = 13.8 kJ mol–1).

higher than that for addition to formaldehyde (∆H298 =
13.8 kJ mol–1). 2-Propoxy radical
An activation energy of 70.4 kJ mol–1 is predicted for β- The predicted values for the enthalpy of formation of 2-
scission of CH3· from the ethoxy radical, which is in close propoxy radical are ∆ f H0o = –20.9 kJ mol–1 and ∆ f H298
o =
agreement with a recent experimental determination (Ea = −41.9 kJ mol , somewhat less negative than the experimen-
–1

70.3 kJ mol–1) measured by laser flash photolysis – laser-in- o = –52.3 kJ mol–1 and ∆ H o = –46 ± 5 kJ mol–1)
tal (∆ f H298 f 300
duced fluorescence (15). Our predicted rate constant (k = and the value recommended by Sun and Bozzelli (45)
19 s–1) is higher than the experimental values of Viskolcz and o = –49.6 kJ mol–1).
(∆ f H298
co-workers (14) (5.2 s–1) and Batt and Milne (44) (1.5 s–1). Loss of the H atom from the 2-propoxy radical by β-

This reaction has also been investigated by Somnitz and scission is predicted to have a barrier (∆H298 ) of 74.1 kJ
Zellner (20) at a modified G2 level of theory, using RRKM mol , which is close to the G2(MP2) value of 76.5 kJ mol–1
–1

theory to derive Arrhenius activation parameters and reaction obtained by Devolder et al. (16). It is somewhat lower than
rate constants. Their Ea value (73.6 kJ mol–1) is slightly the barriers for H-atom loss from the ethoxy and propoxy
higher than ours and their derived log A value (12.99) is sig- radicals which have only a single α-alkyl substituent (CH3

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:40 PM
Color profile: Generic CMYK printer profile
Composite Default screen

438 Can. J. Chem. Vol. 81, 2003

and CH3CH2, respectively). Conversely, the barrier for addi- barriers for both of these reactions are significantly greater
‡ ‡
tion of the H atom to acetone (∆H298 = 30.1 kJ mol–1) is than for ethyl-radical addition to acetaldehyde (∆H298 =
–1
somewhat higher than that for addition to the aldehydes 20.1 kJ mol ). The reaction enthalpies for all three proces-

CH3CH=O and CH3CH2CH=O (∆H298 = 22.0 and 21.3 kJ ses are generally similar (∆H298 = –26.4 to –28.3 kJ mol–1).
–1
mol , respectively).
β-Scission of the 2-propoxy radical yielding CH3· + tert-Butoxy radical
CH3CH=O is predicted to be endothermic, with ∆H0 = The present calculations predict that β-scission of the tert-
19.8 kJ mol–1 and ∆H298 = 27.1 kJ mol–1 (Table 3). The bar- butoxy radical should occur with a barrier of 54.5 kJ mol–1,
rier (∆H298

) is predicted to be 57.7 kJ mol–1. This value is only 3.2 kJ mol–1 lower than for β-scission of the methyl

similar to that found by Viskolcz and co-workers (16) at the radical from 2-propoxy radical (∆H298 = 57.7 kJ mol–1).

G2(MP2) level (∆H298 = 61.2 kJ mol–1). Our Arrhenius pa- The calculated Arrhenius parameters for the scission re-
rameters log A = 13.96 and Ea = 60.1 kJ mol–1, are in mod- action are log A = 14.06 and Ea = 56.9 kJ mol–1. Ex-
erate agreement with the experimental values, measured by perimentally determined kinetics of tert-butoxy radical de-
laser flash photolysis – laser-induced fluorescence (log A = composition have been reported by Choo and Benson (11)
14.08, Ea = 63.7 kJ mol–1) (16). Our calculated rate at 298 K (log A = 14.1, Ea = 64 kJ mol–1), Batt et al. (47) (log A =
is k = 2.8 × 103 s–1, somewhat higher than the experimental 14.04 ± 0.37, Ea = 62.5 ± 0.6 kJ mol–1), Viskolcz and co-
rate (k = 8.3 × 102 s–1). The corresponding theoretical workers (17) (log A = 14.0, Ea = 60.5 kJ mol–1) and Blitz et
RRKM values of Somnitz and Zellner (20) are Ea = 62.2 kJ al. (48) (log A = 13.15, Ea = 57 ± 2 kJ mol–1). While our Ea
mol–1, log A = 13.38, and k = 3.53 × 102 s–1 at 300 K, in value is close to that of Blitz et al. (48), their log A value is
somewhat better agreement with the experimental values of significantly lower. Our calculations support a value for A
Viskolcz and co-workers (15). close to 1 × 1014, which is also recommended by Viskolcz
Addition of the methyl radical to acetaldehyde is pre- and co-workers (17).

dicted to have a significantly higher barrier (∆H298 = 30.6 kJ As noted earlier for the corresponding H-atom additions
–1
mol ) than the addition of the H atom to acetaldehyde to acetone, methyl-radical addition to acetone is hindered by
(∆H298

= 22.0 kJ mol–1) or the addition of the methyl radical a significantly higher barrier (∆H298‡
= 38.1 kJ mol–1) than

to formaldehyde (∆H298 = 20.1 kJ mol–1). for the corresponding additions to acetaldehyde and

propanal (∆H298 = 30.6 and 31.5 kJ mol–1, respectively).
2-Butoxy radical
β-Scission of the 2-butoxy radical may proceed by three Prop-2-en-1-oxy and but-3-en-2-oxy radicals
pathways: loss of the H atom, methyl radical, or ethyl radi- The two unsaturated oxy radicals, prop-2-en-1-oxy and
cal. Loss of the H atom entails a very similar barrier but-3-en-2-oxy, may undergo β-scission by losing a vinyl

(∆H298 = 73.1 kJ mol–1, Table 3) to that for H-atom loss from radical, or by losing a hydrogen atom or a methyl radical, re-

the 2-propoxy radical (∆H298 = 74.1 kJ mol–1). Likewise the spectively. The structures of the OCCC anticlinal conformers

barrier for H atom addition to butan-2-one (∆H298 = 27.6 kJ of these radicals, which are linked to the most stable (also
–1
mol , Table 4) is similar to that for the addition to acetone OCCC anticlinal) conformations of the transition structures
(∆H298

= 30.1 kJ mol–1). for H loss or CH3 loss, are shown in Fig. 1. In each case, the
Loss of either of the two alkyl groups, methyl or ethyl, is <OCCC dihedral angle is such as to orient one of the allylic
hindered by smaller barriers than that for H-atom loss, with groups (H and CH3, respectively) approximately perpendicu-
∆H298

= 59.8 and 46.5 kJ mol–1, respectively. Somnitz and lar to the plane of the vinyl group and well-positioned for π-
Zellner (20) also investigated the two β-scission pathways in- bond-assisted departure.
volving alkyl group loss. Their theoretical RRKM values ‡
The barriers (∆H298 ) for H-atom loss from prop-2-en-1-
(methyl, ethyl) are Ea = (60.8 kJ mol–1, 50.0 kJ mol–1), log oxy (74.7 kJ mol ) and but-3-en-2-oxy (69.7 kJ mol–1) radi-
–1
A = (12.26, 13.15), and k = (46.0 s–1, 2.67 × 104 s–1) at cals are lower than those for the equivalent processes in the
300 K. Our corresponding values from Table 3 are (methyl, saturated analogues, 1-propoxy (83.4 kJ mol–1) and 2-butoxy
ethyl) Ea = (62.3 kJ mol–1, 48.9 kJ mol–1), log A = (13.25, (73.1 kJ mol–1), consistent with π-bond participation in the
13.72), and k = (2.2 × 102 s–1, 1.4 × 105 s–1) at 298 K. The β-scission process for the unsaturated radicals. The lower
Ea values are in good agreement, but the RRKM log A val- barrier for the departure of the methyl radical from but-3-en-
ues are lower than ours and that recommended by Fittschen 2-oxy (48.2 kJ mol–1) as compared to methyl loss from 2-
et al. (17), who suggest a common value (log A = 14 ± 0.3) butoxy (59.8 kJ mol–1) provides additional support for the π-
for β C-C scission in alkoxy radicals. Experimental estimates bond assistance in the former case.
of the rate constant for the dissociation to the ethyl radical With the exception of H· loss from methoxy, the loss of
plus acetaldehyde lie in the range 7.5 × 102 to 3 × 104 s–1 vinyl radicals are the most endothermic of those listed in Ta-
(see reference (20b) for a review and discussion). An experi- ble 3. The barriers for vinyl group loss from the prop-2-en-
mental investigation in which both channels were observed 1-oxy and but-3-en-2-oxy radicals are ∆H298‡
= 92.4 kJ mol–1
found a difference of 11.2 kJ mol–1 between Ea values, and –1
and 83.1 kJ mol , respectively. The higher values for vinyl
that the log A value for the ethyl channel was higher by 0.6 radical loss partly reflects the greater endothermicity in
(46). Both features are in reasonable agreement with the the- these cases.
oretical results.
The barrier for methyl-radical addition to propanal

Influence of substituents in β-scission reactions
(∆H298 = 31.5 kJ mol–1) is quite similar to that for the addi- The data in Table 3 are grouped according to the depart-

tion to acetaldehyde (∆H298 = 30.6 kJ mol–1). However, the ing radical (H·, CH3·, CH3CH2·, or CH2=CH·). It is apparent

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:40 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Rauk et al. 439

Fig. 1. The unsaturated prop-2-en-1-oxy and but-3-en-2-oxy radicals.

that within any of these groupings, increasing alkyl substitu- Transition structures for the addition of hydrogen, methyl,
tion at the site of oxygen attachment promotes β-scission. ethyl, and vinyl radicals to acetaldehyde are shown in Fig. 2,
Thus the barriers (∆H‡) for β-scission of H· decrease in the along with the calculated Arrhenius activation parameters. A
sequence: CH3O· >> CH3CH2O· ≈ CH3CH2CH2O· > general plot of the calculated activation energies (Ea) against
(CH3)2CHO· ≈ CH3CH2CH(CH3)O·. Likewise, the ∆H‡ val- reaction enthalpies (∆H298) is displayed in Fig. 3. It is clear
ues for β-scission of CH3· decrease in the sequence: that Ea generally decreases with increasing exothermicity, as
CH3CH2O· > (CH3)2CHO· ≈ CH3CH2CH(CH3)O· > expected from the Bell–Evans–Polanyi relationship. In addi-
(CH3)3CO·. The two examples of loss of ethyl and vinyl tion, for a given exothermicity, the barriers tend to be lower
groups also follow this pattern. These results are all consis- for radicals with a lower ionization energy. These observa-
tent with a decrease in ∆H‡ accompanying a decrease in the tions are qualitatively consistent with the empirical relation-
reaction endothermicity ∆H, as expected from the Bell–Ev- ship (eq. [4]) of Choo and Benson (11), and with
ans–Polanyi relationship. The decreased endothermicity may considerations based on a curve-crossing model (13, 51).
in turn be associated with increased hyperconjugative stabi- Lower ionization energies enable greater participation of
lization of the product carbonyl compounds with increased charge-transfer configurations of the type radical+carbonyl–,
alkyl substitution. which in turn leads to a lowering of the barrier. The barriers
In addition to the reaction enthalpy effect, the barriers for follow the pattern (CH3)2C=O ≈ CH3CH2C(CH3)=O >
β-scission (and for the reverse radical addition to carbonyl, CH3CH=O ≈ CH3CH2CH=O > CH2=O, in line with the re-
see below) decrease as the ionization energy of the departing action exothermicity. Particularly low barriers are observed
radical decreases, consistent with the development of posi- for the highly exothermic additions involving the vinyl
tive charge at the incipient radical in the transition structure. radical.
Our calculated Mulliken charges indeed indicate that there is An interesting observation, already noted above for the
a small amount of positive charge of approximately +0.05– addition of H· to CH2=O, is that, in contrast to the situation
0.07 in the incipient radical moiety in the TS. The low net for the β-scission reactions, the ∆G298‡
values for the radical
charge on the departing radical (H·, CH3·, CH3CH2·, or additions are considerably greater than the ∆H298 ‡
values.
CH2=CH·) is consistent with a loose (late) TS with more This arises because of substantial negative entropies of acti-
than 80% of the net spin on the radical. Interestingly, there vation, a result that is not unexpected for a bimolecular reac-
is also a significant polarization of charge in the remainder tion. The differences between ∆G298 ‡
and ∆H298

depend on
of the transition structure, with a negative charge of approxi- the radical that is added, and increase in the sequence H·
mately –0.4 on the incipient carbonyl oxygen and a positive (26–31 kJ mol–1) < CH3· (37–45 kJ mol–1) ≈ CH2=CH· (37–
charge of approximately +0.3 on the remainder of the mole- 40 kJ mol–1) < CH3CH2· (43–46 kJ mol–1).
cule. The extent of this polarization increases with increas- Examination of the log A values shows that the frequency
ing alkyl substitution. factors (A) for H· addition are close to 1 × 1010, while those
for alkyl radical addition are two to three orders of magni-
Radical addition to the carbonyl group tude smaller. Frequency factors for vinyl-radical addition are
The addition of radicals to alkenes has been extensively close to those for methyl-radical addition.
studied by high-level theoretical calculations (for example,
see refs. 13 and 49). However, the addition of radicals to car-
bonyl compounds has received less attention (for example, see Conclusions
refs. 19 and 50). This reaction is the reverse of the β-scission
reaction that we have discussed in detail above and we there- The structures of the alkoxy radicals methoxy (CH3O·),
fore make only a few brief comparative remarks here. ethoxy (CH3CH2O·), 1-propoxy (CH3CH2CH2O·), 2-propoxy
Table 4 contains our computed data for radical additions ((CH3)2CHO·), 2-butoxy (CH3CH2CH(CH3)O·), tert-butoxy
to carbonyl compounds, grouped according to the radical be- ((CH3)3CO·), prop-2-en-1-oxy (CH2=CHCH2O·), and but-3-
ing added. An initial comment is that, because of the en-2-oxy (CH2=CHCH(CH3)O·), and the products of their β-
endothermicity of the β-scission reactions, the barriers for scission reactions, have been determined at the B3-LYP/6-
the reverse radical addition reactions are considerably 31G(d) level of theory. CBS-RAD has been used to calculate
smaller than the β-scission barriers. o ). The predicted ∆ H o
the enthalpies of formation (∆ f H298 f 298

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
July 3, 2003 1:14:52 PM
Color profile: Generic CMYK printer profile
Composite Default screen

440 Can. J. Chem. Vol. 81, 2003

Fig. 2. Calculated transition structures, bond lengths in (Å), and Arrhenius activation parameters (CBS-RAD, kJ mol–1) for the addition
of hydrogen, methyl, ethyl, and vinyl radicals to acetaldehyde. Distances for hydrogen, methyl, and ethyl addition are by CBS-RAD–
Scanmax; for vinyl group addition, the value is from B3-LYP/6-31G(d) optimization.

Fig. 3. Activation energies for the addition of hydrogen, methyl, CBS-RAD level is an underestimation of ∆H0‡ by 1.6–2.4 kJ
ethyl, and vinyl radicals to carbonyl compounds vs. the reaction mol–1 (C—H) or 0.4–1.3 kJ mol–1 (C—C).
enthalpy. Eyring (∆H0‡ , ∆H298

, ∆G298

) and Arrhenius (log A, Ea) ac-
tivation parameters for both the forward (β-scission) and re-
verse (radical addition to carbonyl) pathways were
calculated within the rigid rotor – harmonic oscillator model
and transition-state theory. Most of the available experimen-
tal kinetic data refer to the β-scission pathway. Where direct
comparison was possible, the agreement with the experimen-
tal is very good, generally within 1–5 kJ mol–1 for Ea, and
0.5 for log A.
The values of the barriers for either β-scission or radical
addition to carbonyl are dominated by the reaction enthalpy:
the barrier decreases with decreasing endothermicity (β-
scission) or increasing exothermicity (radical addition). As a
consequence, for example, alkyl or vinyl substituents in the
alkoxy radical lower the barrier for β-scission of H·, CH3·, or
CH3CH2· groups. In addition to the enthalpy effect, the barrier
depends on the ionization energy of the departing radical in
the β-scission reaction (or equivalently, the adding radical in
the radical additions). Smaller barriers are associated with
lower ionization energies. These results are consistent with
expectations based on the curve-crossing model, the smaller
ionization energies allowing greater participation of charge-
transfer configurations of the type radical+carbonyl– in the
values agree well with the available experimental data, with transition structure. We find that in the transition structures,
a mean absolute deviation of 5.4 kJ mol–1. the departing radical carries a small positive charge while
Transition structures (TS) for all the β-scission pathways there is substantial charge polarization in the nascent carbonyl
were located at the B3-LYP/6-31G(d) level and, in most compound, the oxygen atom becoming negatively charged
cases, at the CBS-RAD–Scanmax level. The bond being and the rest of the molecule becoming positively charged.
cleaved in the TS tends to be longer with B3-LYP by 0.10–
0.16 Å (C—H) or 0.04–0.10 Å (C—C). At the B3-LYP ge- Acknowledgements
ometry, the predicted B3-LYP barriers (∆H0‡ ) are 13–20 kJ
mol–1 higher than the CBS-RAD values for C-H scission, Financial support by the Natural Sciences and Engineering
and 3–14 kJ mol–1 higher for C—C bond breaking. The con- Research Council of Canada (NSERC), the Australian Re-
sequence of not reoptimizing the bond being broken at the search Council (ARC), and the award of Visiting Fellowships

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
July 3, 2003 1:14:58 PM
Color profile: Generic CMYK printer profile
Composite Default screen

Rauk et al. 441

at the Australian National University (to AR, RJB, and SLB) Replogle, R. Gomperts, R.L. Martin, D.J. Fox, J.S. Binkley,
are gratefully acknowledged. We also thank Professor Allan D.J. Defrees, J. Baker, J.P. Stewart, M. Head-Gordon, C. Gon-
East for helpful discussions, and the Multimedia Advanced zalez, and J.A. Pople. (1995) Gaussian 94, (SGI-Revision B.3),
Computational Infrastructure (MACI) at the University of Gaussian, Inc., Pittsburgh PA. (b) M.J. Frisch, G.W. Trucks,
Calgary, the Australian Partnership for Advanced Computing, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
and the Australian National University Supercomputer Facil- V.G. Zakrzewski, J.A. Montgomery, Jr., R.E. Stratmann, J.C.
ity for generous allocations of computing resources. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N. Kudin,
M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R.
Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J.
Ochterski, G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma,
References D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman,
1. J.A. Miller, R.J. Knee, and C.I. Westbrook. Annu. Rev. Phys. J. Cioslowski, J.V. Ortiz, A.G. Baboul, B.B. Stefanov, G. Liu,
Chem. 41, 317 (1990). A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R.L.
2. J. Warantz. In Combustion chemistry. Edited by W.C. Gardi- Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A.
ner, Jr. Springer, New York. 1984. Nanayakkara, C. Gonzalez, M. Challacombe, P.M.W. Gill, B.
3. (a) R. Atkinson. J. Phys. Chem. Ref. Data 26, 215 (1997); Johnson, W. Chen, M.W. Wong, J.L. Andres, C. Gonzalez, M.
(b) R. Atkinson. Int. J. Chem. Kinet. 29, 99 (1997); (c) E.S.C. Head-Gordon, E.S. Replogle, and J.A. Pople. Gaussian 98.
Kwok, J. Arey, and R. Atkinson. J. Phys. Chem. 100, 214 Gaussian, Inc., Pittsburgh, Pennsylvania. 1998.
(1996). 24. H.J. Werner, P.J. Knowles, R.D. Amos, A. Bernhardsson, A.
4. D.M. Golden. In Chemical kinetic data needs for modeling the Berning, P. Celani, D.L. Cooper, M.J.O. Deegan, A.J. Dobbyn,
lower troposphere. NBS Special Publication 557, August 1979, F. Eckert, C. Hampel, G. Hetzer, T. Korona, R. Lindh, A.W.
pp. 51–61. Lloyd, S.J. McNicholas, F.R. Manby, W. Meyer, M.E. Mura,
5. (a) L. Batt. Int. Rev. Phys. Chem. 6, 53 (1987); (b) L. Batt. Int. A. Nicklass, P. Palmieri, R. Pitzer, G. Rauhut, M. Schütz, H.
J. Chem. Kinet. 11, 977 (1979). Stoll, A.J. Stone, R. Tarroni, and T. Thorsteinsson. Molpro
6. M.J. Davies. Arch. Biochem. Biophys. 336, 163 (1996). 2000.6. University of Birmingham, Birmingham. 1999.
7. P. Gray and A. Williams. Chem. Rev. 59, 239 (1959). 25. A.P. Scott and L. Radom. J. Phys. Chem. 100, 16 502 (1996).
8. J.K. Kochi. J. Am. Chem. Soc. 84, 1193 (1962). The scale factors used are 0.9806 (ZPVEs), 1.0015 (S), and
9. W. Zhang and P. Dowd. Tetrahedron, 49, 1965 (1993). 0.9989 (H° – H0o).
10. S. Wilsey, P. Dowd, and K.N. Houk. J. Org. Chem. 64, 8801 26. P.M. Mayer, P.J. Parkinson, D.M. Smith, and L. Radom. J.
(1999). Chem. Phys. 108, 604 (1998).
11. K.Y. Choo and S.W. Benson. Int. J. Chem. Kinet. 13, 833 27. (a) M.W. Wong and L. Radom. J. Phys. Chem. 99, 8582
(1981). (1995); (b) M.W. Wong and L. Radom. J. Phys. Chem. A, 102,
12. M. Weber and H. Fischer. J. Am. Chem. Soc. 121, 7381 (1999). 2237 (1998).
13. H. Fischer and L. Radom. Angew. Chem. Int. Ed. 40, 1340 28. D.A. McQuarrie. Statistical thermodynamics. Harper and Row,
(2001). New York. 1973.
14. H. Hippler, F. Striebel, and B. Viskolcz. Phys. Chem. Chem. 29. D.D. Wagman, W.H. Evans, V.B. Parker, R.H. Schumm, I.
Phys. 3, 2450 (2001). Halow, S.M. Bailey, K.L. Chyrney, and R.L. Nuttall. J. Phys.
15. F. Caralp, P. Devolder, C. Fittschen, N. Gomez, H. Hippler, R. Chem. Ref. Data, 11 (Suppl. 2) (1982).
Mereau, M.T. Rayez, F. Striebel, and B. Viskolcz. Phys. Chem. 30. S.G. Lias, G.A. Bartmess, J.F. Liebman, J.L. Holmes, R.D.
Chem. Phys. 1, 2935 (1999). Levin, and W.G. Mallard. J. Phys. Chem. Ref. Data 17 (Suppl.
16. P. Devolder, C. Fittschen, A. Frenzel, H. Hippler, G. 1) (1988)
Poskrebyshev, F. Striebel, and B. Viskolcz. Phys. Chem. 31. J. Cioslowski, M. Schimeczek, G. Liu, and V. Stoyanov. J.
Chem. Phys. 1, 675 (1999). Chem. Phys. 113, 9377 (2000).
17. C. Fittschen, H. Hippler, and B. Viskolcz. Phys. Chem. Chem. 32. D.R. Lide (Editor). CRC Handbook of chemistry and physics.
Phys. 2, 1677 (2000). 82nd ed. CRC Press, West Palm Beach, Florida, U.S.A. 2001.
18. G. Lendvay and B. Viskolcz. J. Phys. Chem. A, 102, 10 777 33. See for example: (a) J.P. Guthrie. J. Phys. Chem. A, 105, 8495
(1998). (2001); (b) P.W. Atkins. Physical Chemistry. Oxford Univer-
19. N.D.K. Petraco, W.D. Allen, and H.F. Schaeffer, III. J. Chem. sity Press, Oxford. 1998.
Phys. 116, 10 229 (2002). 34. D.R. Stull, E.F. Westrum, Jr., and G.C. Sinke. The thermody-
20. (a) H. Somnitz and R. Zellner. Phys. Chem. Chem. Phys. 2, namics of organic compounds. John Wiley and Sons, New
1899 (2000); (b) H. Somnitz and R. Zellner. Phys. Chem. York. 1969.
Chem. Phys. 2, 1907 (2000). 35. P.J. Robinson. J. Chem. Ed. 55, 509 (1978).
21. (a) W.J. Hehre, L. Radom, P.v.R. Schleyer, and J.A. Pople. Ab 36. T.M. Ramond, G.E. Davico, R.L. Schwartz, and W.C.
initio molecular orbital theory. Wiley, New York. 1986; (b) F. Lineberger. J. Chem. Phys. 112, 1158 (2000).
Jensen. Introduction to computational chemistry. Wiley, New 37. M.L. Coote, G.P.F. Wood, and L. Radom. J. Phys. Chem. A,
York. 1999. 106, 12 124 (2002).
22. W. Koch and M.C. Holthausen. A chemist’s guide to density 38. D.K. Malick, G.A. Petersson, and J.A. Montgomery. J. Chem.
functional theory. Wiley-VCH, Weinheim. 2000. Phys. 108, 5704 (1998).
23. (a) M.J. Frisch, G.W. Trucks, H.B. Schlegel, P.M.W. Gill, B.G. 39. O.M. Zarechnaya, I.A. Opeida, and A.F. Dmitruk. Russ. J.
Johnson, M.A. Robb, J.R. Cheeseman, T.A. Keith, G.A. Org. Chem. 37, 1405 (2001).
Petersson, J.A. Montgomery, K. Raghavachari, M.A. Al- 40. J. Berkowitz, G.B. Ellison, and D. Gutman. J. Phys. Chem. 98,
Laham, V.G. Zakrzewski, J.V. Ortiz, J.B. Foresman, J. 2744 (1994).
Cioslowski, B.B. Stefanov, A. Nanayakkara, M. Challacombe, 41. D.L. Osborn, D.J. Leahy, E.R. Ross, and D.M. Neumark.
C.Y. Peng, P.Y. Ayala, W. Chen, M.W. Wong, J.L. Andres, E.S. Chem. Phys. Lett. 235, 484 (1995).

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:41 PM
Color profile: Generic CMYK printer profile
Composite Default screen

442 Can. J. Chem. Vol. 81, 2003

42. T.S. Dibble. J. Mol. Struct. 485–486, 67 (1999), and 48. M. Blitz, M.J. Pilling, S.H. Robertson, and P.W. Seakins. Phys.
refs. therein. Chem. Chem. Phys. 1, 73 (1999).
43. S.P. Walch. J. Chem. Phys. 98, 3076 (1993). (Analysis based 49. C. Selcuki and V. Aviyente. J. Mol. Model. 7, 398 (2001).
on cited private communication by F. Temps). 50. C. Gonzalez, C. Sosa, and H.B. Schlegel. J. Phys. Chem. 93,
44. L. Batt and R.T. Milne. Int. J. Chem. Kinet. 1, 549 (1977). 2435 (1989).
45. H. Sun and J.W. Bozzelli. J. Phys. Chem. A, 106, 3947 (2002). 51. L. Radom, M.W. Wong, and A. Pross. In Controlled radical
46. R.M. Drew, J.A. Kerr, and J. Olive. Int. J. Chem. Kinet. 17, polymerization. Edited by K. Matyjaszewski. ACS Symposium
167 (1985). Series 685. American Chemical Society, Washington, D.C.
47. L. Batt, M.W.M. Hisham, and M. Mackay. Int. J. Chem. Kinet. 1998. pp. 31–49.
21, 535 (1989).

© 2003 NRC Canada

I:\cjc\cjc8106\V02-206.vp
June 9, 2003 2:48:41 PM

You might also like