You are on page 1of 7

Inuence of the chloro substituent position on

the triplet reactivity of benzophenone


derivatives: a time-resolved resonance Raman
and density functional theory study
Wen Li,
a
Jiadan Xue,
a
Shun Cheung Cheng,
a
Yong Du
b
* and
David Lee Phillips
a
*
A nanosecond time-resolved resonance Raman (ns-TR
3
) spectroscopic investigation of the photoreduction reactions and ability
of several chloro-substituted benzophenone (Cl-BP) triplets is described. The TR
3
results show that the 3-chlorobenzophenone
(3-Cl-BP), 4-chlorobenzophenone (4-Cl-BP) and 4,4-dichlorobenzophenone (4,4-dichloro-BP) triplets exhibit similar hydrogen
abstraction ability with the parent BP triplet. In 2-propanol, the 3-Cl-, 4-Cl- and 4,4-dichloro-diphenylketyl (DPK) radicals were
observed and they appear to react with dimethylketyl radicals at the para-position to form a light absorption transient species.
These transient species were characterized with TR
3
spectra, and identied with the help of results from density functional
theory calculations. In an acetontitrile/water (MeCN:H
2
O) 1:1 mixed solvent, these DPK radicals were also observed but with
slower formation rates. However, the 2-Cl-DPK radical was observed to form with a lower yield and a signicantly slower
formation rate than the other chloro-substituted benzophenones examined here in 2-propanol under the same experimental
conditions. These results reveal that the 2-chloro substituent reduces the hydrogen abstraction ability of the substituted BP
triplet, which was not as expected based on the assumption that the electron-withdrawing group could increase its photoreduc-
tion ability. This unusual ortho effect of the chlorine substitution is briey discussed. Copyright 2011 John Wiley & Sons, Ltd.
Supporting information may be found in the online version of this article.
Keywords: chloro-substituted benzophenone; time-resolved resonance Raman spectroscopy; triplet state; ketyl radical; hydrogen
abstraction
Introduction
The photoreduction of aromatic ketones especially benzophenone
(BP) and its derivatives has received much attention over the years
and a number of studies have been reported.
[110,13,27,31,33]
It is
well-known that the hydrogen abstraction reaction of BP and its
derivatives typically arises from their np* excited triplet state that
can react with a hydrogen donor molecule to produce a
corresponding ketyl radical and free radical of the hydrogen donor
molecule. This kind of reaction may proceed by either a direct
hydrogen atom transfer or a coupled electron/proton transfer.
The reactivity of the BP derivatives towards a hydrogen donor in
solution are determined by the nature of their lowest excited
triplet state with three types (namely np*, pp* and charge transfer
state) commonly found. Therefore, the efciency of the photore-
duction of BP derivatives can be affected signicantly by the
solvent polarity and the nature and position of the substituent
on the phenyl ring. For example, an electron-donating substituent
group on the phenyl ring such as NH
2
, OH, OCH
3
usually reduces
the reactivity of the substituted BP towards a hydrogen donor,
because of the relative position of the np* and pp* states being
largely altered in polar and nonpolar solvents.
[1118]
In addition,
the electron-donating group could decrease the electrophilicity
at the carbonyl oxygen and hence make the corresponding BP
derivative less reactive towards a hydrogen atom donor compared
with the unsubstituted molecule. On the other hand, electron-
withdrawing groups generally increase the photoreduction ability
of substituted BP triplet states. The decauoro-BP derivative was
found to be 1040 times more reactive than the parent BP
molecule,
[19,20]
and the hydrogen abstraction rate of the triplet
4-triuoromethylacetophenone molecule is six times faster than
the parent acetophenone molecule.
[21]
Some studies reported for the photochemistry and photophysics
of BP determined that the energy gap of the lowest excited singlet
state (S
1,
np*) and the second triplet state (T
2,
pp*) is fairly small and
the spinorbit coupling between these two states is allowed.
[2224]
Therefore, the intersystemcrossing (ISC) process may be as efcient
as other S
1
processes such as internal conversion (IC) and uores-
cence emission so that the T
1
(np*) state is produced via a rapid
ISC with a high quantum yield, which was determined to be close
* Correspondence to: David Lee Phillips, Department of Chemistry, The University
of Hong Kong, Pokfulam Road, Hong Kong, China. E-mail: phillips@hku.hk;
Yong Du, Centre for THz Research, China Jiliang University, Hangzhou,
310018, China. E-mail: yongdu@cjlu.edu.cn
a Department of Chemistry, The University of Hong Kong, Pokfulam Road, Hong
Kong, China
b Centre for THz Research, China Jiliang University, Hangzhou 310018, China
J. Raman Spectrosc. (2011) Copyright 2011 John Wiley & Sons, Ltd.
Research Article
Received: 12 April 2011 Revised: 9 August 2011 Accepted: 19 August 2011 Published online in Wiley Online Library
(wileyonlinelibrary.com) DOI 10.1002/jrs.3078
to unity.
[25]
Thus, the T
1
state is the predominant precursor for
subsequent processes and reactions after ultraviolet excitation of
BP and a number of its derivatives.
The reaction details and properties of the excited states of BP
and its derivatives have been revealed extensively through
the broad application of various time-resolved spectroscopies,
especially transient absorption (TA). For example, the properties
and reaction mechanisms of the ketyl radical and the radical
from the solvent and their reaction to produce a light absorbing
transient (LAT) have been reported.
[2630]
The intersystem cross-
ing from S
1
to T
1
for BP and its derivatives has been observed
directly by ultrafast TA with measured rates of ~10
11
s
1
,
[31,32]
and a rapid water addition process to the protonated BP triplet
was identied to result in the fast deactivation of the triplet BP
in acidic aqueous solution.
[33]
Moreover, nanosecond time-
resolved resonance Raman (ns-TR
3
) spectroscopy is an effective
technique for further clearly characterizing the structural and
electronic properties of transient species, and has been used to
help identify the intermediates and photoproducts of BP and its
derivatives.
[10,34,35]
In our previous work,
[37,45,46]
the BP triplet
state, ketyl radicals, LAT and some other intermediates formed
from the BP triplet and its derivatives in organic and aqueous
solvents have been successfully identied and characterized
using ns-TR
3
spectra.
In this paper, ns-TA and ns-TR
3
experiments were performed on
chloro-substituted BP derivatives, 2-Cl-BP, 3-Cl-BP, 4-Cl-BP and 4,4-
dichloro-BP in organic and aqueous solvents, in an attempt to
understand the properties and photochemistry of their triplet
states, and the effects of the chloro-substituent on the BP triplet
reactivity. Density functional theory (DFT) calculations were used
to determine the structures and vibrational wavenumbers of
the intermediates and these results were compared with the
experimental TR
3
spectra to help in the interpretation of the
experimental results. We compare the structure, properties and
chemical reactivity of the chloro-substituted BP derivatives studied
here and briey examine how the substituent position of the
chlorine atom affects the experimental results.
Experimental and Computational Methods
The chloro-substituted BP samples were commercially obtained
from Alfa Aesar, China (with 98% purity) and used as received.
Spectroscopic grade acetonitrile (MeCN), 2-propanol (IPA) and
distilled water (H
2
O) were used as solvents for the experiments.
Sample solutions of ~1.5 mM concentration were employed in
the TR
3
experiments.
The ns-TA measurements were performed on a LP-920 Laser
ash photolysis setup (Edinburgh Instruments, UK). The 266 nm
pump laser pulse was obtained from the fourth harmonic output
of an Nd:YAG Q-switched laser, and the probe light was provided
by a 450 W Xe arc lamp. These two light beams were focused
onto a 1 cm quartz cell. The signals analyzed by a symmetrical
CzernyTurner monochromator were detected by a Hamamatsu
R928 photomultiplier and the signal processed via an interfaced
computer and analytical software.
A homemade ns-TR
3
spectroscopy described previously
[4749]
was used in this work to acquire the ns-TR
3
spectra. The 266 nm
pump laser pulse was obtained from the fourth harmonic of a
Nd:YAG Q-switched laser. The 319.9 nm pump laser pulse came
from the third anti-Stokes of a hydrogen Raman-shifted line from
the second harmonic of a Nd:YAG Q-switched laser. These two
lasers were electronically synchronized by a pulse delay genera-
tor and the relative timing of the pump and probe laser pulses
were display on a 500 MHz oscilloscope. The time resolution of
the experiments was about 5 ns. The pump and probe laser
beams were focused onto a owing liquid stream of sample
using a near-collinear geometry. The Raman scattered light was
collected in a backscattering geometry and detected by a liquid
nitrogen-cooled charge-coupled device detector. The Raman
signal was read out by an interfaced personal computer (PC) and
the Raman spectra were obtained from subtraction of an
appropriately scaled probe-before-pump spectrum from the
corresponding pump-probe spectrum. The TR
3
spectra were
calibrated by utilizing the known wavenumbers of the MeCN
Raman bands. The areas of the Raman bands were determined
by ttings using Lorentzian functions.
All of the geometry optimizations, vibrational modes and
vibrational frequencies were obtained from calculations
employing Becke, three-parameter, LeeYangParr (B3LYP)
and unrestricted Becke, three-parameter, LeeYangParr
(UB3LYP) methods and utilizing a 6-311 G** basis set. A
Lorentzian function with a 15 cm
1
bandwidth was used with
the Raman vibrational frequencies and relative intensities to
determine the calculated Raman spectra presented in this work.
Solvent effects were calculated by self-consistent reaction eld
(SCRF) method based on a polarizable continuum model. All of
the DFT calculations were performed using the GAUSSIAN 03
programs
[36]
operated on the High Performance Computing
Cluster installed at the University of Hong Kong.
Results and Discussions
Ns-TR
3
results in MeCN and IPA
The 266 nm laser photolysis of 2-Cl-BP, 3-Cl-BP, 4-Cl-BP and 4,
4-dichloro-BP in MeCN generates transient resonance Raman
spectra shown in Fig. 1, which were obtained using a 319.9 nm
probe wavelength at 10 ns after the photolysis pulse. These
transient spectra are assigned to the corresponding Cl-BP triplets
because they exhibit high similarity with that of the parent BP
triplet, as shown in the comparison presented in Fig. 1. Most of
the Raman bands in Fig. 1 for the BP analogue triplets are mostly
associated with the CC stretching and CH bending motions,
except that the ~1500 cm
1
Raman bands are associated with
aromatic CC stretching, and the ~1200 cm
1
Raman bands for
the 3-Cl-BP, 4-Cl-BP and 4,4-dichloro-BP triplets are due to
carbonyl CO stretching motions. All of these triplets in MeCN
have a lifetime of ~80 ns under ambient conditions and ~25 ns
under O
2
saturated conditions. This result is in agreement with
those that we obtained previously for the parent BP triplet.
[37]
Close examination of Fig. 1 shows that all of the transient species
except the 2-Cl-BP triplet have a similar pattern at ~1200 cm
1
,
which is consistent with the typical np* character triplet.
[3840]
Figure 2 shows the ns-TR
3
spectra of 3-Cl-BP in IPA at different
time delays from 5 ns to 300 ms acquired with a 319.9 nm probe
wavelength following 266 nm photolysis. Three transient species
are observed in Fig. 2 over the examined time range. At early
delay times such as 5 ns, the Raman spectrum is due to a mixture
of the rst and second species. The rst species has Raman bands
at 701, 967, 1210 and 1535 cm
1
, identical to that obtained in
neat MeCN. Therefore, this species is believed to be the 3-Cl-BP
triplet. The second species has the intense bands at 986, 1121,
1180, 1478, 1557 and 1576 cm
1
, and grows in to reach their
W. Li et al.
wileyonlinelibrary.com/journal/jrs Copyright 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2011)
maximum at ~50 ns. The 3-Cl-BP triplet has a np* conguration so
that the intermolecular hydrogen abstraction reaction can easily
take place with hydrogen donor solvents such as IPA to produce
a ketyl radical and we tentatively assign the second species to the
3-Cl-diphenylketyl (3-Cl-DPK) radical. With the decay of the
3-Cl-DPK radical, a third species with Raman bands at 1597 and
1648 cm
1
was generated and its growth is correlated with the
decay of the 3-Cl-DPK radical. Previous studies
[5052]
suggested
that the DPK radical would react with an associated dimethylketyl
(DMK) to produce the para-LAT as the main photoproduct in IPA
(Scheme 1). The reaction of the 3-Cl-DPK radical examined here
and the DMK radical may form two products because the DMK
radical could attack the para-position of either the chlorophenyl
ring or the phenyl ring. According to the DFT calculation, the free
energies difference of these two kinds of LAT species is only
0.6 kcal/mol, and the DFT calculation predicted Raman spectra
of these two products are both in reasonable agreement with
the experimental TR
3
spectra as shown in Fig. 3. Therefore, the
third species observed at post-delay times in Fig. 2 may be
carried by two types of para-LAT.
The DPK radicals and LAT species are also observed for 4-Cl-BP
and 4,4-dichloro-BP in IPA and are displayed as the ns-TR
3
spectra
in Figs S1 and S2. Similar with BP, the photoreduction of 3-Cl-BP,
4-Cl-BP, 4,4-dichloro-BP triplets proceeds very fast with the time
constant of ~20ns for the ketyl radicals formation in IPA. However,
it is a little slower for the hydrogen abstraction reaction of the
2-Cl-BP triplet than the others examined here. Figure 2 also shows
the ns-TR
3
spectra of 2-Cl-BP in IPA obtained with a 319.9nm
probe wavelength following 266 nm excitation. Global analysis of
the most intense Raman band at 1580 cm
1
for the 2-Cl-DPK
radical in Fig. 2 demonstrates that the 2-Cl-DPK radical grows in
with a time constant of ~80 ns (Fig. S3). This result indicates that
the chloro-substitution at the ortho-position reduces the hydrogen
abstraction reactivity of the substituted BP triplet compared with
the meta, para and unsubstituted BP derivatives.
Ns-TR
3
results in aqueous solution
The 266-nm laser photolysis of 3-Cl-BP in a MeCN: H
2
O/1:1
solvent generates the ns-TR
3
spectra at different time delays from
5 ns to 10 ms obtained with a 319.9 nm probe wavelength ,which
are shown in Fig. 4. Three species are observed: the rst transient
species has its strongest Raman bands at 669, 967, 1170, 1213
and 1535 cm
1
and is assigned to the triplet state because of its
resemblance with the 3-Cl-BP triplet spectrum obtained in neat
MeCN. The decay of the triplet gives rise to a second species with
the strongest Raman bands at 985, 1178 and 1576 cm
1
. Integra-
tion of the areas of the 1535 and 1576 cm
1
Raman band
obtained in Fig. 4 shows that the triplet decays with the time
constant of ~130 ns, the second species grows and decays with
time constants of ~100 ns and ~500 ns, respectively. The second
species observed in aqueous solution at post-delay times is
identied as the 3-Cl-DPK radical by comparison of the Raman
spectrum obtained in aqueous solution at 400 ns to that obtained
in IPA at 200 ns, and the comparison of these spectra are
presented in Fig. 5. A close examination of Fig. 5 shows that the
two experimental spectra are almost identical, which implies they
are due to the same species. After the decay of the ketyl radical, a
new species is observed, and tentatively assigned to a LAT
species produced by coupling of the DPK radical with a hydroxide
radical, similar to the mechanism of the ketyl radical with the
associated DMK radical in IPA. The DFT calculation predicted
Figure 2. Ns-TR
3
spectra of 2-Cl-BP (upper) and 3-Cl-BP (lower) in IPA
obtained with 266 nm excitation and a 319.9-nm probe wavelength at
various delay times that are indicated next to the spectra. The asterisk
(*) marks solvent subtraction artifacts and the # band is due to stray
laser line.
Figure 1. Resonance Raman spectra of 2-Cl-BP (A), 3-Cl-BP (B), 4-Cl-BP
(C) and 4,4-dichloro-BP (D) triplets and the parent BP triplet (E) obtained
in MeCN using a 319.9-nm probe wavelength at a 10-ns delay time after
266 nm excitation. The asterisk (*) marks solvent subtraction artifacts
and the # band is due to stray laser line.
Inuence of the chloro substituent position on triplet reactivity of benzophenone derivatives
J. Raman Spectrosc. (2011) Copyright 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs
Raman spectra for such LAT photoproducts in water are provided
and seem to agree well with the experimental TR
3
spectrum
obtained after 266 nm photolysis of 3-Cl-BP in MeCN: H
2
O/1:1 at
6 ms as shown in Fig. S4.
The ns-TR
3
experiments in aqueous solution were also
performed for the other chloro-substituted BP derivatives. The
TR
3
results in Figs S5 and S6 suggest that the fate of 4-Cl-BP and
4,4-dichloro-BP triplets strongly resembles the parent BP and
3-Cl-BP triplets. In aqueous solution, the 4-Cl-DPK and 4,
4-dichloro-DPK radicals were produced and their formation corre-
lates with the triplets decay. However, no ketyl radical was
observed after the 2-Cl-BP triplet decay in aqueous solution.
Figure 4 provides the ns-TR
3
spectra of 2-Cl-BP in the solvent of
MeCN:H
2
O/1:1 at different time delays from 5 ns to 10ms. Kinetic
analysis of the 1561 cm
1
Raman band integration area shows that
the triplet decays with the time constant of ~370 ns under ambient
condition and ~100 ns in the O
2
saturated condition.
Reactions of chloro-substituted BP derivative triplets in water
In our TR
3
experiments, the satisfactory agreement of the
transient species generated after the 3-Cl-BP triplet decay in
the aqueous solution with the authentic Raman spectrum of
the 3-Cl-DPK radical obtained in IPA under essentially the same
experimental conditions clearly demonstrates that the transient
species in aqueous solution is the ketyl radical. The generation
of the 3-Cl-DPK radical may simultaneously produce OH radicals
and because these OH radicals could possibly then attack the
DPK radical to form a cyclohexadienyl complex, which has
similar structure to the LAT species.
The DPK radical formation mechanism in aqueous solutions has
been a question for many years. A thermochemical calculation
[41]
indicated the direct hydrogen abstraction of the BP triplet from a
H
2
O molecule could be energetically feasible, but the reaction rate
has an upper limit of only ~10
3
M
1
s
1
.
[42,43]
Assuming the DPK
Scheme 1. Mechanism for the formation of para-LAT in IPA.
Figure 3. Comparison of the experimental resonance Raman spectra of 3-
Cl-BP obtained in neat MeCN at 6ms delay time with 266-nm excitation and
a 319.9-nm probe wavelength (B), to the DFT calculation predicted spectra
for the para-LAT products with the DMK radical attack on the phenyl ring
(A) and the chlorophenyl ring (C). The asterisk (*) marks solvent subtraction
artifacts and the # band is due to stray laser line. Dotted lines display the
correlation between the experimental and calculated Raman bands.
Figure 4. Ns-TR
3
spectra of 2-Cl-BP (upper) and 3-Cl-BP (lower) in the
solvent of MeCN:H
2
O/1:1 obtained with a 266-nm excitation and a
319.9-nm probe wavelength at various delay times that are indicated next
to the spectra. The asterisk (*) marks solvent subtraction artifacts and the
# band is due to stray laser line.
Figure 5. Ns-TR
3
spectra of 3-Cl-BP obtained in neat IPA at a 200-ns delay
time (A) and in MeCN:H
2
O/1:1 at a 400-ns delay time (B). The asterisk (*)
marks solvent subtraction artifacts and the # band is due to stray laser line.
W. Li et al.
wileyonlinelibrary.com/journal/jrs Copyright 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2011)
radical has the same Raman intensity under the same concentra-
tion in IPA and in aqueous solution, comparison of the 3-Cl-DPK
radical Raman spectra in IPA and in a MeCN:H
2
O/1:1 mixed solvent
obtained under the same experimental conditions suggests that
the yield of the 3-Cl-DPK radical is only about 0.09 in the MeCN:
H
2
O/1:1 solvent (see supplemental information for details). This
result suggests that the 3-Cl-DPK radical formation rate in a
MeCN:H
2
O/1:1 solvent is ~3.610
4
M
1
s
1
, which is faster than
the predicted limit but much slower than the hydrogen abstrac-
tion rate of the BP triplet from IPA (2.3 10
6
M
1
s
1
).
[21]
The
generation of the DPK radical in the aqueous solvent examined
here appears to not be due to the reaction of the triplet with
the ground state BP derivatives or the triplet itself, because the
formation rate of the DPK radical in our TR
3
experiments is not
dependent on the concentration of the BP derivative precursors
or the pump power used during the TR
3
experiments. Another
possibility for the DPK radical formation in aqueous solutions could
be an electron transfer followed by a proton transfer. Hydrated
electrons could possibly be generated when the 266 nm pump
was used during TR
3
experiments, but it is hard to detect because
of its small Raman scattering cross-section.
[44]
No DPK radical was observed after the 2-Cl-BP triplet decay in
the aqueous solutions in the TR
3
experiments, and the 2-Cl-BP
triplet has a lifetime of ~370 ns in a MeCN:H
2
O/1:1 mixed solvent,
much longer than that in neat MeCN. This prolonged lifetime for
the 2-Cl-BP triplet in aqueous solvent is probably mainly due to
the O
2
solubility being much smaller in water than in MeCN.
The nanosecond transient absorption result showed that the 2-
Cl-BP triplet has a lifetime of ~2 ms in MeCN:H
2
O/1:99 mixed
solvent under ambient condition (Fig. S7).
Assignment of the 2-Cl-BP triplet and its conguration
The transient Raman spectra obtained in MeCN after 266 nm
photolysis of 2-Cl-BP in Fig. 1 was assigned to its triplet because
the lifetime of the transient shortens when with O
2
purging
compared with that observed under an ambient condition. We
note that the Raman band pattern of the 2-Cl-BP triplet looks
somewhat different from the other Cl-BP triplets or the parent
BP triplet in the 11001300 cm
1
region as shown in Fig. 1. The
2-Cl-BP triplet has two congurations, with the chlorine atom
close to or far away from the carbonyl oxygen (denoted as
U-2-Cl-BP and D-2-Cl-BP triplet, respectively). DFT calculations
predict that the U-2-Cl-BP triplet form has a 1.5 kcal/mol lower
energy than the D-2-Cl-BP triplet form in the gas phase and a
2.5 kcal/mol lower energy in MeCN and H
2
O solutions, therefore,
the U-2-Cl-BP triplet form should be more favorable.
Figure 6 presents a comparison of the experimental resonance
Raman spectrum for the 2-Cl-BP triplet obtained in MeCN to the
DFT calculation predicted normal Raman spectrum for the U-2-
Cl-BP and D-2-Cl-BP triplet species. This comparison suggests the
calculated vibrational frequencies for the U-2-Cl-BP triplet agrees
better with the experimental values. Tables S1S4 list the tentative
vibrational assignments with nominal descriptions of the normal
modes obtained from the DFT calculation for the U-2-Cl-BP triplet
and D-2-Cl-BP, 3-Cl-BP and 4-Cl-BP triplets. A close exanimation of
Tables S1S4 shows that all of the D-2-Cl-BP, 3-Cl-BP and 4-Cl-BP
triplets have a strong vibration at ~1230 cm
1
corresponding to
the carbonyl CO stretch, which agrees with the TR
3
spectra for
the 3-Cl-BP and 4-Cl-BP triplets shown in Fig. 1 and is also consis-
tent with their triplet np* character exhibited in IPA solvent.
However, for the U-2-Cl-BP triplet species, vibration analyses show
that all of the vibrational frequencies in the 1100 1300 cm
1
region correspond to the CH bend in the phenyl or the chloro-
phenyl planes except the bands 1123, 1195 and 1342 cm
1
, which
are the combination of CH bend and pretty weak CO stretch
vibration. This difference is due to the geometry change of the
U-2-Cl-BP triplet species compared with the other Cl-BP triplets
and accounts for the unique Raman band pattern observed in
the 11001300 cm
1
region for the 2-Cl-BP triplet. If the D-2-Cl-BP
triplet conguration was predominant, the intense Raman band
at ~1200cm
1
should be observed in the TR
3
experiment for the
2-Cl-BP triplet, similar to the BP and 3-Cl-BP and 4-Cl-BP triplets.
However, this is not the case in Fig. 1. The resonance enhancement
effect in the TR
3
experiments is similar for all of the triplets
examined here as time-dependent (TD) DFT calculation results indi-
cate that all of these triplets have a strong oscillator strength at
~320nm (Table S5), the probe wavelength used in TR
3
experiments.
The chloro-substituent at the meta and para positions does not
noticeably change the CO bond geometry relative to the two
phenyl planes. In contrast, the 2-chloro substituent inuences
the Cl-BP triplet geometry signicantly as shown in Table S6, in
which parameters were obtained from the DFT calculation. For
U-2-Cl-BP triplet, the CO bond is close to a co-planar structure
with the chlorophenyl ring (7

), but it is about 25

away from
the chlorophenyl ring for the 3-Cl-BP and 4-Cl-BP triplets. The
2-chloro substituent also makes the charge distribution on the
triplets different from the others as shown in Table S7, especially
the charge on the chlorine atom. In the U-2-Cl-BP triplet congu-
ration, the chlorine atom bears a positive charge while chlorine
has a negative charge in the 3-Cl-BP and 4-Cl-BP triplets as
expected. All of these could result in the reactivity reduction of
the hydrogen abstraction reaction for the 2-Cl-BP triplet.
2-Chloro substituent reduction of the hydrogen abstraction
ability of the substituted BP triplet
The TR
3
results obtained after 266 nm photolysis of 2-Cl-BP in
IPA with a 319.9 nm probe wavelength clearly showed that the
2-Cl-DPK radical forms slower than the parent and other
chloro-substituted DPK radicals, which demonstrates that the
Figure 6. Comparison the experimental resonance Raman spectrum of
2-Cl-BP obtained in neat MeCN (B) to the DFT calculation predicted
normal Raman spectrum for the D-2-Cl-BP triplet (A) and U-2-Cl-BP triplet
(C) species. The asterisk (*) marks solvent subtraction artifacts and the #
band is due to stray laser line. Dotted lines display the correlation
between the experimental and calculated Raman bands.
Inuence of the chloro substituent position on triplet reactivity of benzophenone derivatives
J. Raman Spectrosc. (2011) Copyright 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs
2-chloro substituent reduces the hydrogen abstraction ability of
the 2-Cl-BP triplet. In general, carbonyl compounds having a pp*
triplet as the lowest excited state exhibit no or low photoreduction
efciency, such as some electron-donating group substituted BP
derivatives.
[1417]
In addition, the solvent polarity can affect the
hydrogen abstraction ability of the carbonyl compound triplet.
For example, if the lowest excited np* triplet is too close to the
pp* triplet state, the polar solvent can invert these two states or
promote the contribution to the unreactive pp* state. However,
the 2-Cl-BP triplet observed here still has an np* character in both
nonpolar and polar solvents, as presented in Fig. S8, which provides
the TA spectra of 2-Cl-BP triplet in cyclohexane, MeCN and MeCN:
H
2
O/1:1 mixed solvents. The TA spectra for the 2-Cl-BP triplet
obtained in these three solvents are almost identical and indicate
that the solvent polarity does not substantially change the lowest
triplet property and the contributions on the np* and pp* states.
The geometry comparison has revealed that the chloro-
substitution at the ortho- position has a large inuence on the
substituted BP triplet geometries. In addition, the 2-chloro
substituent also strongly changes the electron density of the
lowest unoccupied molecular orbital (LUMO) for the 2-Cl-BP
triplets as illustrated in Fig. 7, which can provide some insight
into how the 2-chloro substituent reduces the hydrogen
abstraction ability of the 2-Cl-BP triplet. Figure 7 shows that
for 2-Cl-BP the LUMO orbit delocalizes the chlorine atom and
carbonyl. The chlorine atom at the 2-position leads to signicant
interaction with the nearby carbonyl so that it appears like a
pipi interaction that is not present in the 3- and 4- chlorine
substituted BP derivatives. This clearly suggests the chemical re-
activity of the 2-position substituted chloro BP species will be
signicantly different from the 3-Cl-BP and 4-Cl-BP triplets,
which is consistent with the observed experimental results de-
scribed earlier.
Conclusion
A ns-TR
3
spectroscopic investigation of the reactivity of the Cl-BP
triplets has been described. The 3-Cl-BP, 4-Cl-BP and 4,
4-dichloro-BP triplets exhibit similar hydrogen abstraction ability
as has been found previously for the parent BP triplet. In IPA, the
3-Cl-DPK, 4-Cl-DPK and 4,4-dichloro-DPK radicals were observed
and react with DMK radical most likely at the para-position to
form a para-LAT. In a MeCN:H
2
O/1:1 aqueous solvent, these
DPK radicals were also observed but with a slower formation rate.
However, a smaller yield and a slower rate of formation of the
2-Cl-DPK radical were observed compared with the other
chloro-substituted BP derivatives that were investigated here in
IPA. All of these results indicate that the 2-chloro substituent
reduces the hydrogen abstraction ability of the substituted BP
triplet, not as may be expected that an electron-withdrawing
group could increase its photoreduction activity. DFT calculation
results suggest that the chloro-substituent at the ortho position
changes the BP triplet geometry and makes the carbonyl oxygen
less electrophilic from its interaction with the carbonyl moiety.
Acknowledgements
This research was supported by grants from the Research Grants
Council of Hong Kong (HKU 7035/08P) and the University Grants
Committee Special Equipment Grant (SEG-HKU-07) to DLP.
Supporting Information
Supporting information may be found in the online version of
this article.
References
[1] W. Moore, G. Hammond, R. Foss, J. Am. Chem. Soc. 1961, 83, 2789.
[2] R. Hochstrasser, J. Wessel, Chem. Phys. Lett. 1973, 19, 156.
[3] R. Hochstrasser, H. Lutz, G. Scott, Chem. Phys. Lett. 1974, 24, 162.
[4] M. Topp, Chem. Phys. Lett. 1975, 32, 144.
[5] R. V. Bensasson, J. C. Gramain, J. Chem. Soc., Faraday Trans. 1 1980,
76, 1801.
[6] H. Miyasaka, N. Mataga, Bull. Chem. Soc. Jpn. 1990, 63, 131.
[7] X. Cai, M. Sakamoto, M. Fujitsuka, T. Majima, Chem. Eur. J. 2005, 11,
6471.
[8] J. Chilton, L. Giering, C. Steel, J. Am. Chem. Soc. 1976, 98, 1865.
[9] A. Demeter, T. Berces, J. Photochem. Photobiol. A: Chem 1989, 46, 27.
[10] T. Tahara, H. Hamaguchi, M. Tasumi, J. Phys. Chem. 1987, 91, 5875.
[11] G. Porter, P. Suppan, Trans. Faraday Soc. 1966, 62, 3375.
[12] G. Porter, P. Suppan, Trans. Faraday Soc. 1965, 61, 1664.
[13] A. Beckett, G. Porter, Trans. Faraday Soc. 1963, 59, 2038.
[14] A. C. Bhasikuttan, A. K. Singh, D. K. Palit, J. P. Mittal, J. Phys. Chem. A
1998, 102, 3470.
[15] A. K. Singh, D. K. Palit, J. P. Mittal, J. Phys. Chem. A 2000, 104, 7002.
[16] P. Aspari, N. Ghoneim, E. Haselbach, M. Von Raumer, P. Suppan,
E. Vauthey, J. Chem. Soc., Faraday Trans. 1996, 92, 1689.
[17] N. Ghoeneim, A. Monbelli, D. Pilloud, P. Suppan, J. Photochem.
Photobiol. A: Chem. 1996, 94, 145.
[18] D. I. Scuster, M. D. Goldstein, P. Bane, J. Am. Chem. Soc. 1997, 99, 187.
[19] L. C. T. Shoute, J. P. Mittal, J. Phys. Chem. 1993, 97, 8630.
[20] R. Anandhi, S. Umapathy, J. Raman Spectrosc. 2000, 31, 331.
[21] A. Gilbert, J. Baggott, P. J. Wagner, Essentials of Molecular Photochem-
sitry, CRC Press, Inc., Boca Raton, FL, 2000, pp. 307
[22] M. El-Sayed, J. Chem. Phys. 1963, 38, 2834.
[23] B. Shah, D. Neckers, J. Am. Chem. Soc. 2004, 126, 1830.
[24] J. Scaiano, J. Photochem. 1973/74, 2, 81.
[25] N. Turro, Modern molecular photochemistry, Univ. Science Books,
Sausalito, Ca 1991.
[26] L. C. T. Shoute, R. E. Huie, J. Phys. Chem. A 1997, 101, 3467.
[27] M. Sakamoto, X. Cai, M. Fujitsuka, T. Majima, J. Phys. Chem. A 2006,
110, 11800.
[28] A. C. Bhasikuttan, A. K. Singh, D. K. Palit, A. V. Sapre, J. P. Mittal,
J. Phys. Chem. A 1999, 103, 4703.
Figure 7. Simple schematic diagrams depicting the electron density for the LUMO of the Cl-BP triplets predicted from the DFT calculations. The green
atom in the geometry represents chlorine.
W. Li et al.
wileyonlinelibrary.com/journal/jrs Copyright 2011 John Wiley & Sons, Ltd. J. Raman Spectrosc. (2011)
[29] M. Sakamoto, X. Cai, M. Hara, S. Tojo, M. Fujitsuka, T. Majima, J. Phys.
Chem. A 2004, 108, 8147.
[30] A. Demeter, T. Berces, J. Photochem. Photobiol. A: Chem. 1989, 46, 27.
[31] B. K. Shah, M. A. J. Rodgers, D. C. Neckers, J. Phys. Chem. A 2004, 108, 6087.
[32] Y. Du, J. Xue, M. D. Li, X. Guan, D. W. McCamant, D. L. Phillips, Chem.
Eur. J. 2010, 16, 6961.
[33] M. Ramseier, P. Senn, J. Wirz, J. Phys. Chem. A 2003, 107, 3305.
[34] T. Tahara, H. Hamaguchi, M. Tasumi, Chem. Phys. Lett. 1988, 152, 135.
[35] T. Tahara, H. Hamaguchi, M. Tasumi, J. Phys. Chem. 1990, 94, 170.
[36] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.
Cheeseman, J. A. Montgomery Jr., T. Vreven, K. N. Kudin, J. C. Burant,
J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi,
G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara,
K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross,
V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O.
Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala,
K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G.
Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K.
Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q.
Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A.
Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith,
M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M.
W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, J. A. Pople,
Gaussian, Inc., Wallingford CT, 2004.
[37] Y. Du, C. Ma, W. M. Kwok, J. Xue, D. L. Phillips, J. Org. Chem. 2007, 72,
7148.
[38] S. P. Webb, S. W. Yeh, L. A. Phillips, M. A. Tolbert, J. H. Clark, J. Am.
Chem. Soc. 1984, 106, 7286.
[39] S. P. Webb, S. W. Yeh, L. A. Phillips, M. A. Tolbert, J. H. Clark, J. Phys.
Chem. 1986, 90, 5154.
[40] B. J. Schwartz, L. A. Peteanu, C. B. Harris, J. Phys. Chem. 1992, 96, 3591.
[41] M. B. Ledger, G. Porter, J. Chem. Soc., Faraday Trans. 1 1972, 68, 539.
[42] D. J. Lougnot, P. Jacques, J. P. Fouassier, J. Photochem. 1982, 19, 59.
[43] D. J. Lougnot, P. Jacques, J. P. Fouassier, H. L. Casal, N. Kim-Thuan,
J. C. Scaiano, Can. J. Chem. 1985, 63, 3001.
[44] M. Mizuno, T. Tahara, J. Phys. Chem. A 2001, 105, 8823.
[45] Y. Du, J. Xue, C. Ma, W. M. Kwok, D. L. Phillips, J. Raman. Spectrosc.
2008, 39, 1518.
[46] Y. Du, J. Xue, C. Ma, W. M. Kwok, D. L. Phillips, J. Raman. Spectrosc.
2008, 39, 503.
[47] P. Y. Chan, W. M. Kwok, S. K. Lam, P. Chiu, D. L. Phillips, J. Am. Chem.
Soc. 2005, 127, 8246.
[48] P. Zhu, S. Y. Ong, P. Y. Chan, K. H. Leung, D. L. Phillips, J. Am. Chem.
Soc. 2001, 123, 2645.
[49] Y. L. Li, K. H. Leung, D. L. Phillips, J. Phys. Chem. A 2001, 105, 10621.
[50] J. N. Pitts Jr., R. L. Letsinger, R. P. Taylor, J. M. Patterson, G. Recktenwald,
R. B. Martin, J. Am. Chem. Soc. 1959, 81, 1068.
[51] M. B. Rubin, Tetrahedron Lett. 1982, 23, 4615.
[52] J. C. Fister, J. M. Harris, Anal. Chem. 1995, 67, 701.
Inuence of the chloro substituent position on triplet reactivity of benzophenone derivatives
J. Raman Spectrosc. (2011) Copyright 2011 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs

You might also like