Agapito et al.
Photoacoustic Calorimetry. The basis of photoacoustic calo-
rimetry,12,18 our photoacoustic calorimeter setup,19,20 and the
experimental technique are described in detail elsewhere.21,22
Briefly, argon-purged solutions in benzene of ca. 0.4 M di-tert-
butyl peroxide and an adequate concentration (see Analysis of
Thermochemical Data) of each organic molecule studied (cyclo-
hexene, 1,3-cyclohexadiene, and 1,4-cyclohexadiene) were flowed
through a quartz flow cell and photolyzed with pulses from a
nitrogen laser (337.1 nm, pulse width 800 ps). To check for
multiphoton effects, the incident laser energy was varied by using
neutral density filters (ca. 5-30 µJ/pulse at the cell, flux <40 J
m-2). Each pulse produced photolysis of di-tert-butyl peroxide (t-
BuOOBu-t), generating tert-butoxyl radicals (reaction 3), which in
turn abstracted an allylic hydrogen from the organic molecule RH,
reaction 4.
R-C-H bond, were determined by using density functional theory
(DFT).25 In this approach the energy of a system, E[F], is given by
eq 6, where VNN is the nucleus-nucleus repulsion energy, Hcore is
the one-electron kinetic and electron-nuclei potential energy
contribution to the total energy, and Vee is the Coulombic electron-
electron repulsion energy.
E[F] ) VNN + Hcore + Vee + Ex[F] + Ec[F]
(6)
The terms Ex[F] and Ec[F] are respectively the exchange and
correlation functionals of the electronic density, F. The optimized
geometry for a molecule is found by determining the set of nuclear
coordinates that minimizes the energy given by eq 6. In this work
the geometry optimizations were carried out with Becke’s three-
parameter hybrid method26 with the correlation functional of Lee,
Yang, and Parr (B3LYP).27 The accuracy of the energy also depends
on the completeness of the basis set in which the molecular orbitals
are expanded. For these geometry optimizations Dunning’s triple-ú
correlation consistent basis set (cc-pVTZ) was used.28 Vibrational
analysis was performed for all optimized geometries to ensure that
they represented minima of the energy surfaces. The choice of
B3LYP/cc-pVTZ geometries for the structural analysis was dictated
by its cost-effectiveness and the fact that several works indicate
that the molecular geometries thus obtained are in good agreement
with experimental data.29-31 Nevertheless, it is well-known that DFT
methods systematically underestimate bond dissociation enthalp-
ies.32,33 Therefore, in addition to B3LYP, BDEs were also computed
by using two composite theoretical procedures, namely CBS-Q and
CBS-QB3.34-36 These were specifically devised to allow an accurate
determination of thermochemical properties for large systems, by
resorting to extrapolation to the complete basis set limit. We note,
however, that the geometry optimizations of CBS-Q and CBS-QB3
are performed respectively with MP2(FC)/6-31G† (frozen-core
Møller-Plesset second-order perturbation theory,37 in which the
electrons from inner shells are excluded from the calculation of
the correlation energy) and B3LYP/6-31G†, and therefore are
slightly less accurate than B3LYP/cc-pVTZ geometries.29,31
Complete basis set extrapolated coupled cluster calculations with
single and double excitations and perturbative inclusion of triple
excitations (CCSD(T)),38 using B3LYP/cc-pVTZ geometries, are
also reported. Extrapolation of CCSD(T) energies to complete basis
set was carried out through a dual (2)cc-pVDZ,3)cc-pVTZ)
scheme proposed by Truhlar for both the Hartree-Fock and
correlation energies.39 This procedure has proven to be very reliable
for the determination of BDEs,40,41 although computationally more
demanding than any of the aforementioned methods.
t-BuOOBu-t (sln)
9
hν8 2 t-BuO• (sln)
(3)
2RH (sln) + 2 t-BuO• (sln) f 2R• (sln) + 2 t-BuOH (sln) (4)
Each laser pulse induced a sudden volume change in solution,
which generated an acoustic wave, detected by a piezoelectric
transducer (0.5 MHz) in contact with the bottom of the cell. The
signals were amplified and measured by a digital oscilloscope. The
signal-to-noise ratio was improved by averaging 32 acquisitions
for each data point obtained at a given laser energy. The apparatus
was calibrated by carrying out a photoacoustic run with an optically
matched solution of o-hydroxybenzophenone (in the same mixtures
but without the peroxide), which dissipates all of the absorbed
energy as heat.18 For each run (experiment or calibration), four data
points were collected corresponding to four different laser intensities
obtained with the neutral density filters. The resulting waveforms
from each data point were recorded for subsequent mathematical
analysis, affording two waveforms for each point: sample and
calibration. The analysis involved, for each laser energy, first the
normalization of both waveforms and then their deconvolution,
using the software Sound Analysis.23 This analysis first allowed
the confirmation of the reaction scheme indicated above (reactions
3 and 4) and then afforded the observed fraction of photon energy
released as heat, φobs,i, for each process, and the lifetime of the
second, τ2. An estimate of the rate constant can be obtained from
this lifetime.24 The enthalpy of the hydrogen abstraction reaction
was derived from eq 5,
-∆obsH2
∆rH2 )
(5)
Φr
(25) Zhou, Z.; Parr, R. G. J. Am. Chem. Soc. 1989, 111, 7371-7379.
(26) Becke, A. D. J. Chem. Phys. 1993, 98, 5648.
(27) Lucas, C. R.; Gabe, E. J.; Lee, F. L. Can. J. Chem. 1988, 66, 429-
434.
where ∆obsH2 corresponds to the observed enthalpy change and is
calculated by multiplying Em ) NAhν (the molar photon energy)
by φobs,2 (the observed heat fraction associated with reaction 2).
Φr is the reaction quantum yield for the photolysis of di-tert-butyl
peroxide. All experiments were performed at 293 ( 0.5 K.
Theoretical Calculations. The structures of propene, isobutene,
1- and (E)-2-butene, 3-methylbut-1-ene, (E)-2-pentene, (E)-1,3- and
1,4-pentadiene, cyclohexene, and 1,3- and 1,4-cyclohexadiene, as
well as the respective radicals resulting from homolysis of an
(28) Dunning, T. H., Jr. J. Chem. Phys. 1989, 90, 1007.
(29) Byrd, E. F. C.; Sherrill, C. D.; Head-Gordon, M. J. Phys. Chem. A
2001, 105, 9736-9747.
(30) Agapito, F.; Cabral, B. J. C.; Martinho Simo˜es, J. A. THEOCHEM
2005, 719, 109-114.
(31) Wang, J. T.; Feng, Y.; Liu, L.; Li, X.-S.; Guo, Q.-X. Chin. J. Chem.
2004, 22, 642-648.
(32) Cabral do Couto, P.; Guedes, R. C.; Cabral, B. J. C.; Martinho
Simo˜es, J. A. Int. J. Quantum Chem. 2002, 86, 297-304.
(33) Kern, R. D.; Zhang, Q.; Yao, J.; Jursic, R. S.; Tranter, R. S.; Greybill,
M. A.; Kiefer, J. H. Proc. Combust. Inst. 1998, 102, 143-150.
(34) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A. J. Chem.
Phys. 1996, 104, 2598-2619.
(18) Braslavsky, S. E.; Heibel, G. E. Chem. ReV. 1992, 92, 1381-1410.
(19) Borges dos Santos, R. M.; Lagoa, A. L. C.; Martinho Simo˜es, J. A.
J. Chem. Thermodyn. 1999, 31, 1483-1510.
(20) Nunes, P. M.; Correia, C. F.; Dos Santos, R. M. B.; Simoes, J. A.
M. Int. J. Chem. Kinet. 2006, 38, 357-363.
(21) Correia, C. F.; Nunes, P. M.; Borges, dos Santos, R. M.; Martinho
Simo˜es, J. A. Thermochim. Acta 2004, 420, 3-11.
(22) Nunes, P. M.; Agapito, F.; Costa Cabral, B. J.; Borges dos Santos,
R. M.; Martinho Simo˜es, J. A. J. Phys. Chem. A 2006, 110, 5130-5134.
(23) Sound Analysis, version 1.50D; Quantum Northwest: Spokane, WA,
1999.
(35) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G.
A. J. Chem. Phys. 1999, 110, 2822-2827.
(36) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G.
A. J. Chem. Phys. 2000, 112, 6532-6542.
(37) Møller, C.; Plesset, M. S. Phys. ReV. 1934, 46, 618.
(38) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Headgordon, M.
Chem. Phys. Lett. 1989, 157, 479-483.
(24) Nunes, P. M.; Correia, C. F.; Borges dos Santos, R. M.; Martinho
Simo˜es, J. A. Int. J. Chem. Kinet. 2006, 38, 357-363.
(39) Truhlar, D. G. Chem. Phys. Lett. 1998, 294, 45-48.
8772 J. Org. Chem., Vol. 72, No. 23, 2007