Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/46217709

The Photophysics of a Polar Molecule in a Nonpolar Cryogenic Glass - The


Effects of Dimerization on (1-Butyl-4-(1H-inden-1-ylidene)-1,4-
dihydropyridine (BIDP)

Article  in  The Journal of Physical Chemistry A · October 2010


DOI: 10.1021/jp1047375 · Source: PubMed

CITATIONS READS
0 98

3 authors:

Anat Kahan Boris Bazanov


California Institute of Technology Hebrew University of Jerusalem
54 PUBLICATIONS   588 CITATIONS    5 PUBLICATIONS   120 CITATIONS   

SEE PROFILE SEE PROFILE

Yehuda Haas
Hebrew University of Jerusalem
71 PUBLICATIONS   1,788 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Synthesis of the pentazolate anion View project

Reproduction View project

All content following this page was uploaded by Anat Kahan on 19 January 2018.

The user has requested enhancement of the downloaded file.


J. Phys. Chem. A 2010, 114, 10563–10574 10563

The Photophysics of a Polar Molecule in a Nonpolar Cryogenic Glass - The Effects of


Dimerization on (1-Butyl-4-(1H-inden-1-ylidene)-1,4-dihydropyridine (BIDP)

Anat Kahan, Boris Bazanov, and Yehuda Haas*


Institute of Chemistry and the Farkas Center for Light Induced Processes, the Hebrew UniVersity of Jerusalem,
Jerusalem, Israel
ReceiVed: May 24, 2010; ReVised Manuscript ReceiVed: August 9, 2010

The first excited state of BIDP was shown in a previous communication to exhibit an ultrafast decay in fluid
nonpolar, nonprotic solutions due to the presence of a S1/S0 conical intersection (CI). In frozen polar and
nonpolar glasses a strong fluorescence was observed, rationalized by the hindering of the internal torsion
required to reach the geometry of the CI. Complete analysis of the data was hampered by some unusual
observations in nonpolar glasses. In this paper we show that they can be explained by assuming dimer formation,
with a formation constant of Keq ) (4 ( 3) × 105 M-1 at 83 K and ∆Hdim ) 7 ( 2 kcal/mol. A complete
analysis of the spectra is presented, and fluorescence quantum yields of the monomer and dimer are reported.
Efficient self-quenching is found, with a Stern Volmer constant, KSV ) (1.5 ( 0.1) × 106 M-1, assigned to
static quenching. The dimer absorption spectrum was extracted from the data and is compared to Kasha’s
exciton model and to quantum chemical (QC) calculations. The basic features of exciton splitting are reproduced
by quantum mechanical calculations, but complete quantitative agreement of the QC computations with the
experimental results is not attained. The previous analysis of the monomer spectra using the displaced harmonic
oscillator model is extended to the more demanding conditions prevailing at cryogenic temperatures. The
derived ∆Hdim is in good agreement with other dimers formation enthalpies and with the quantum mechanical
calculation presented. The new analysis corrects τf in MCHIP to 2.9 × 10-13 s, somewhat smaller than the
value reported in polar solvent in a previous communication, thereby strengthening the assumption that polarity
can reduce the efficiency of CI.

I. Introduction SCHEME 1: Structure of BIDP


The photophysics of 1-butyl-4-(1H-inden-1-ylidene)-1,4-
dihydropyridine (BIDP, Scheme 1) were recently discussed
experimentally1,2 and theoretically.3 The molecule is a model
molecule that was used to explore solvent and electric field
effects on conical intersection (CI).1,4 This molecule belongs
to a family of molecules such as merocyanines5 and 1-fluoro-
2,4-hexadiene,6 whose CIs were considered to be controlled by
the solvent. In fluid solution’s internal conversion was found
to be very efficient in all solvents as confirmed by the absence
of fluorescence, whereas in glassy solutions at cryogenic
ization in cryogenic temperatures is a well-known phenomenon,
temperatures, the fluorescence yield was found to be high (0.59
reported12 already over 50 years ago. In addition to new spectral
in a polar solvent). In nonpolar solvents quantitative measure-
features, dimer formation often also decreases the emission
ments were hampered by dramatic changes of the fluorescence
intensity of the monomer (self-quenching). The classical
spectra. It was suggested that in the nonpolar glasses aggregation
Stern-Volmer mechanism (eq 1) is frequently a satisfactory
takes place, leading to more complicated photophysical behavior.
approximation.
In this paper we report novel measurements performed to check
this hypothesis.
BIDP, as other polar molecules having an extended conju- φf(0)
gated π-electron system, has a well-known tendency to form ) 1 + KSV[Q] (1)
φf(Q)
dimers and higher aggregates, especially in nonpolar solvents.
The phenomenon has been extensively studied due to its
theoretical and practical importance.7-10 A well-known, physi- In fluid environments, dynamic quenching dominates and the
cally transparent model has been developed by Kasha and co- Stern-Volmer constant KSV is interpreted as KSV ) (kq)/(kf),
workers.11 This exciton model predicts the formation of dimers where kq is the bimolecular quenching rate constant, often
and higher aggregates, and analyzes the resulting absorption equaling the diffusion coefficient, and kf is the inverse lifetime
and emission spectra of the aggregates in terms of the spectra of the fluorescent molecule in the absence of the quencher (in
of the monomer constituents and the dimer structure. Dimer- the case of self-quenching, extrapolated to zero concentration).
In glasses where translational motion is restricted, static
* To whom correspondence should be addressed. Fax: +97225618033. quenching is more likely.13 When static quenching results from
10.1021/jp1047375  2010 American Chemical Society
Published on Web 09/10/2010
10564 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

the formation of a complex (or dimer) in the ground state KSV QCHEM program suite.18 The calculations were made for the
often equals the dimerization equilibrium constant Keq for the IDP molecule, without the n-butyl tail connected to the pyridine
reaction pair ring, which is a change that does not effect the electronic
distribution. The monomer ground state structure was fully
k1 optimized using B3LYP/6-31G(d) calculation, and the optimized
2M 98 D structure was used for all other calculations. TDDFT/B3LYP/
6-31G(d) calculations were used for excited states of the
k-1 (2) monomer. In addition, TDDFT/BNL/6-31G(d) calculations were
D 98 2M preformed. This functional, developed by Baer and Livshits,19
helps to correct long-range behavior by a splitting functional
and a complementary exchange-correlation functional, both
the equilibrium constant, Keq, is given by depend on tuning factor γ. The tuning process20 gave γ ) 222.
This functional was chosen as it appears to be more suitable
k1 [D]eq for charge transfer (CT) states, especially for systems too large
Keq ) ) (3) for CAS calculations, such as BIDP dimers, and should yield
k-1 [M]2eq better agreement with experiment than the B3LYP functional.21
For comparison, ground state was also calculated using the MP2
energy correction, to check whether DFT can be used to
Using the exciton model as a guide, it is shown that the dimer
compute intermolecular interactions.22 Different dimer structures
spectrum can be extracted from the low temperature absorption
were calculated using the three methods, based on the optimized
and emission spectra. The data are also compared to DFT
structure of the monomer. Two of the structures are based on
calculations of the ground state dimer structure, and TDDFT
Kasha’s “in line” and “parallel” dimer geometries, with the two
calculations of the dimer spectrum.
dipole moments in line, pointing in the same direction (dimer1,
Experimental and Computational Details see Table 1 in the Results section) or parallel, having opposite
directions (dimer2). Dimer3 was inspired by the T structure of
Experimental Details. BIDP was prepared and purified using the benzene dimer. Most of the structures were previously
literature procedures.14,15 It was characterized by visible and calculated for pyridine dimers.23 For dimer1, 2, and 3, different
NMR spectra and by its fluorescence. The solvents used were distances between the monomers were calculated, and in the
methylcyclohexane (MCH, Sigma-Aldrich, spectroscopic grade), range between 3 and 5 Å only minor differences were found.
2-methyl-butane (isopentane, IP, Sigma-Aldrich, HPLC grade), As a result, a 4 Å distance was chosen. Other structures were
and 2-methyltetrahydrofurane (2MTHF, Sigma-Aldrich, >99%). tested starting with various geometries and performing full
Low temperature spectra were taken in a methylcyclohexane/ optimization; the most stable in energy is labeled dimer4. It
isopentane (MCHIP, 2:3) composition. This solvent was thor- has an anti parallel structure, with a 3.78 Å distance between
oughly dried in preliminary experiments (sodium metal) to the monomers, with the center of mass of one monomer shifted
ensure the exclusion of hydrogen bonding, water aggregates, from the other by 5.16 Å (See Table 1).
etc. In further experiments it was found that the solvent was The 6-31G(d) basis set was the largest basis for which the
water-free, and it was used as is. optimization process converged, and was consequently chosen
For spectroscopic experiments at cryogenic temperatures, a for all further calculations. For a comparison, single-point
homemade apparatus was used, built around an Oxford Instru- calculations using the 6-311G(2d) basis set were also used
ments double-wall cryostat equipped with a liquid nitrogen in TDDFT/B3LYP calculations. Other geometries were
container and heating circuit operated by a temperature controller calculated but were found to be less stable than those four,
(Eurotherm Inc., model 2216 L). The white light source (Ocean which were therefore used in order to discuss the experi-
Optics Inc., halogen-deuterium, UV-vis-NIR DT-MINI-2-GS) mental results.
and miniature spectrometer (Ocean Optics Inc., USB-4000) The PCM model24,25 was used to simulate the solvent effect
collinearly coupled through the cryogenic system by quartz of cyclohexane on dimer4 using DFT/B3LYP/6-31G(d). A few
fibers were used in absorption experiments. For fluorescence trimer structures (Scheme S1 in the Supporting Information)
and excitation profiles a white light source (Hamamatsu CM were also calculated, using B3LYP functional, but as their
150 W xenon-Arc-Lamp E7536) at perpendicular configuration contribution to the absorption spectra was found to be marginal,
was used. For the fluorescence experiments the excitation without any new contribution that can explain the absorption
wavelength was selected by passing the light beam through a spectrum, no TDDFT/BNL calculations were performed on
monochromator, equipped with a 2400 L/mm grating. To them.
calculate quantum yield (QY), coumarine 540 in acetonitrile As in ref 1, an analysis separating the observed experimental
was used as a reference (QY ) 0.63).16 For excitation profiles spectrum to the first two allowed transitions theory predicts for
measurements the monochromator was used in order to change the monomer, S0-S1 and S0-S31 was performed. The displaced
the excitation wavelength and a R928 photomultiplier, with harmonic oscillator (DHO) model, developed by Heller, Myers,
interference filters at 560 or 660 nm (40 nm fwhm), were used and Mathies,26,27 was used, (eq 4).
to collect the emitted light. Prior to the measurements the

[ ]
solution was saturated with dry nitrogen gas to avoid photo-
oxidation during spectroscopic experiments.2 All measurements 4π2e2M2EL -(E0 - Ej 0)2
were carried out in a cylindrical quartz cuvette, 13 mm inner σA(EL) ) exp ×
3pcn 2θ

[∫
diameter. ∞
exp[i(EL - V)t/p]exp(-Γt/p) × (4)
Computational Details. Quantum mechanical (QM) calcula- -∞

]
tions were used to analyze different dimer geometries and their
effect on the spectroscopic characterizations. Calculations were ∏ [exp{-sk[1 - exp(-iωkt)]}]dt
performed using the PCGAMESS programs suite17 and the k)1
Photophysics of a Polar Molecule in a Nonpolar Glass J. Phys. Chem. A, Vol. 114, No. 39, 2010 10565

Figure 1. (a) The UV-vis absorption spectrum of BIDP in MCHIP as a function of the temperature ([M]corr 83K
) 3.3 × 10-5). Inset: normalized
spectra at three temperaturess298, 198, and 143 Ksshowing more clearly three main temperature-dependent features of BIDP absorption. (b) The
OD at 395 nm (black), 450 nm (red), and 505 nm (green) as a function of the temperature. Three regions appear to characterize the temperature
dependence of the absorption; the transition temperatures are T1 ) 147 K and T2 ) 200 K. See text for more details.

σA(EL) is the absorption cross section, EL is the energy of the model. The amount of shifting was determined by the shift
incident photon, e is the charge of the electron, M is the needed for comparison to the monomer excitation spectrum at
electronic transition length between the ground and excited 83 K at low concentration.
electronic states, p is Planck’s constant, c is the velocity of light,
and n is the refraction index. θ is the inhomogeneous width, Experimental and Computational Results
and each band is broadened inhomogeneous. E0 is the energy
Experimental Results. As briefly discussed in ref 1 and
separation between the lowest vibrational level of the ground
shown in more detail in Figure 1, the absorption spectrum of
state and the electronic excited state, exp(-Γt/p) is a damping
BIDP in MCHIP depends strongly on temperature. Whereas at
factor with Γ as the homogeneous line-width, ωk is the frequency
room temperature the spectrum is similar to that in more polar
of the k mode, sk ) ∆k2/2, and ∆k is a displacement between
solvents (2-methyltetrahydrofuran (MTHF), acetonitrile), in the
ground and excited state harmonic surfaces. The model was used
glassy solvent at 83 K the spectrum is much broader (Figures
to fit the excitation spectrum at 83 K of monomer absorption
3 and 5 in ref 1). A second important observation is that at 83
(Figure S1) and the absorption spectrum at 298 K (detailed
K the molecule fluoresces, and its emission spectrum bears an
information in Table S1). Several iterations were performed until
approximate mirror image to the absorption spectrum in polar
the only major changes between room temperature and low
solvents, whereas in MCHIP glass no such relation is observed.
temperature were the homogeneous and the inhomogeneous line
A third noteworthy difference was that in MTHF the fluores-
width factors, and the refractive index, n, while the other
cence quantum yield (0.57) was much higher than in MCHIP
parameters should not change much with temperature. The
glass (∼0.1). It was suggested that, although the BIDP
separation made it possible to calculate the oscillator strength
concentration was relatively low (5 × 10-5 M) dimerization
for each transition, and the radiative rate constant kf0 was
might be involved in the low temperature samples. Indeed,
calculated using the Strickler-Berg equation28 (5).
preliminary experiments showed that the shape of the absorption
spectrum changed upon reducing the solute’s concentration.
∫ I(ν̃)dν̃ ε(ν̃) It transpired that the concentration range in which further
dilution did not change the spectrum’s shape was around 3 ×
kf0 )
1
) 2.88 × 10-9n2 ∫ dν̃ ≈ 10-7 M. Unfortunately absorption spectra in this range are too
τ0 ∫ I(ν̃)ν̃-3dν̃ ν̃ noisy for quantitative analysis, so that fluorescence excitation
0.667n2〈ν̃〉2avf (5) spectra were used to obtain the absorption line-shape at these
low concentrations. In this concentration range (∼10-7 M) the
emission spectrum was found, in contrast with higher concentra-
τ0 is the radiative lifetime, n is the refractive index of the solvent, tions, to be independent of the excitation wavelength. These
I is the fluorescence emission, ν˜ is wavenumber, and ε(ν˜) is data indicate that a single species is present in the dilute
the extinction coefficient, f is oscillator strength of the transition, solutions, whereas at least two are present at higher concentrations.
〈ν〉av is the average wavenumber corresponding to the 0-0 Figure 2a shows the absorption/excitation spectra recorded
transition. for BIDP in MCHIP glass at two different concentrations, 3.6
Attempts to simulate dimer absorption using the DHO model × 10-7 M and 5.4 × 10-5 M. The main differences are the
were also made. The QM calculated transitions were used, and much broader spectrum at the higher concentration and the more
were shifted all by the same amount. The shift is a result of the prominent vibronic structure in the lower concentration. These
difference between the experiment and the QM calculations, features bring to mind the predictions of the exciton model of
and also a result of the difference between the Franck-Condon Kasha for parallel transitions: two electronically excited dimer
QM calculation results and the zero-zero energy in the DHO levels are expected, an allowed one shifted to the blue with
10566 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

TABLE 1: The Energy of IDP Monomer and Dimer and the Difference in Energy between Two Monomers and Dimer,
Calculated by TDDFT/B3LYP/6-31G(d), by TDDFT/BNL/6-31G(d) and by MP2/6-31G(d), as well as Two Trimer Structures
Calculated by TDDFT/B3LYP/6-31G(d)

respect to the monomer, and a forbidden one shifted to the red. procedure similar to that applied to the absorption spectrum.
Several experiments were run at different initial concentrations, The subtraction was made under the approximation that the
and the dimer spectrum was extracted by subtracting the emission at 525 nm and shorter wavelengths is due to the
monomer spectrum from the total one. monomer species only.
In order to do so, we assume that at low concentration of The emission spectra are displayed, along with the absorption
BIDP (∼10-7 M) the excitation spectrum is due to the spectrum spectra, in Figure 3. It is seen that an approximate mirror image
of monomer only. Furthermore, the oscillator strength of the relation is found for the monomer, and also an approximate
monomer at 83 K was assumed to differ only slightly from that one between the dimer emission and the weak, red-shifted,
at room temperature. Thus, by taking into account also the absorption band centered at 505 nm (for comparison, see
changes in the refraction index and in the density of the solvent, Figure 5 in ref 1).
it is possible to estimate the extinction coefficients at 83 K. A The separation of the spectra allowed the measurement of
further assumption is that at 470 nm the absorption is due to the fluorescence quantum yields of the monomer and dimer.
the monomer only, whereas for the red and blue sides of it we The case of the monomer is straightforward, provided the molar
used the equation ODtot ) εM[M]l + ODD. That way we could absorption coefficient in the MCHIP glass is known. An
obtain the monomer and dimer concentrations, and their relative approximate value can be obtained from the absorption spectra,
contributions. This procedure was repeated for several initial but as mentioned above they were too noisy to be reliable. Under
total concentrations of BIDP. the assumption that the oscillator strength of the transitions is
Figure 2b displays the spectrum, assigned to the dimer, independent of the temperature, the molar absorption coefficient
resulting from the subtraction of the monomer contribution from values at room temperature were used to deduce the low
the complete experimental spectrum. The dimer absorption temperature ones. Using this procedure, and the solvent’s density
exhibits two main bands, a broad strong one, about 5000 cm-1 temperature dependence, the molar absorption coefficient of the
fwhm, blue-shifted with respect to the monomer, and a weaker monomer at the maximum (457 nm) is 30000 ( 3000 M-1 cm-1
one, narrower, about 1100 cm-1 fwhm, that is red-shifted. The (room temperature value is (λmax) ) 21 6001), and the molar
splitting between their maxima is about 6000 cm-1, with 4000 absorption coefficient of the dimer obtained upon separation of
cm-1 (2000 cm-1) from the monomer maximum absorption to the spectra is 21 000 ( 2000 M-1 cm-1 at λmax ) 390 nm. These
the blue (red). values lead to successful reconstruction of the low temperature
The monomer emission spectrum was measured directly using spectra (Figure S4). The low temperature values are consistent
a dilute solution, and the dimer spectrum was obtained using a with the narrowing of the monomer band upon cooling.
Photophysics of a Polar Molecule in a Nonpolar Glass J. Phys. Chem. A, Vol. 114, No. 39, 2010 10567

Figure 2. The proposed partition of the BIDP absorption spectrum to the contributions of the monomer and the dimer in MCHIP at 83 K. (a) Full
line: The total absorption spectrum, 5.4 × 10-5 M; Dashed line: The monomer excitation spectrum detected at 660 nm, 3.6 × 10-7 M. (b) The
dimer spectra extracted from spectra shown in (a). See text for details.

Figure 3. Absorption/excitation and emission spectra of BIDP monomer (a) and dimer (b), normalized, at 83 K in MCHIP. (a) Full line: excitation
spectrum, λem ) 560 nm, dashed: emission spectrum, λexc ) 450 nm. (b) Derived absorption (full line) and emission spectra (dashed) of BIDP
dimer.

With the monomer absorption coefficient at hand, the reported in ref 1, in which the dimer contribution was not
fluorescence quantum yield at low concentration at 83 K was considered. It was also observed that the monomer quantum
found to be 0.9 ( 0.1. This value is much higher than that yield strongly decreased as the BIDP concentration was
10568 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

TABLE 2: Fluorescence Quantum Yields of the Monomer The temperature dependence of the absorption of BIDP in
As a Function of the Nominal BIDP Concentrationa MCHIP was shown in Figure 1. Three main temperature regions
tot
OD457 nm (RT)
83K
[M]corr φ(Q) (φ(0))/(φ(Q)) can be distinguished in the spectrum. The first one is between
-6 80 and 180 K, where the broad spectra display three maxima,
0.16 7.4 × 10 0.08 + 0.01 12.5
0.13 5.9 × 10-6 0.11 + 0.01 9.1 at 505, 450, and 395 nm, which were assigned to a combination
0.10 4.4 × 10-6 0.15 + 0.01 6.7 of the monomer and the dimer, as mentioned above. In the
0.08 3.7 × 10-6 0.16 + 0.02 6.3 second temperature range, (180-220 K), the spectrum is still
0.06 3.0 × 10-6 0.17 + 0.02 5.9 quite broad, but the peaks at 505 nm and especially at 395 nm
0.05 2.2 × 10-6 0.24 + 0.02 4.2 are less prominent. The third region is from 220 to 298 K, a
0.03 1.5 × 10-6 0.40 + 0.04 2.5 region where the monomer spectrum is dominant and no dimer
0.02 7.4 × 10-7 0.91 + 0.09 1.1
contribution can be discerned. A large decrease of the absorption
a 83K
[M]corr is the corrected concentration at 83 K for volume intensity is noted as the system is cooled.
decrease; φ (Q) is the quantum efficiency of BIDP monomer Next, Van’t-Hoff’s equation was used to derive the dimer-
83K
fluorescence when the concentration equals [M]corr (self quenching).
ization enthalpy ∆Hdim: (d(ln Keq))/(d(1/T))) ) -(∆Hdim/R). At
each temperature, Keq can be expressed by Keq )[D]/[M]2
)(([M]0 - [M]) × 0.5)/[M]2. Using the beer-Lambert law, [M]
can be expressed by(ODM)/(εM,ν × l), where l is the cell length,
εM,ν is the extinction coefficient and is temperature dependent,
and ODM is the absorbance of the monomer, which is also
temperature dependent. [M]0 has to be corrected for the volume
change incurred upon temperature change. The absorbance of
BIDP in MCHIP is due to both dimer and monomer at low
temperatures; in order to separate them we used the facts that
the excitation spectrum of BIDP in MCHIP at low concentra-
tions is very similar to BIDP in 2MTHF and that in 2MTHF
no dimer is observed even at 83 K, at all concentrations (Figure
S2). Therefore, we calculated the monomer spectra at different
temperatures for BIDP in 2MTHF, by performing a DHO fit
and using the resulting homogeneous and inhomogeneous width
parameters at each temperature for BIDP in MCHIP, to get εM,ν
as a function of temperature. By scaling it to the experimental
spectra we got ODM as a function of temperature (Figure 5). It
Figure 4. Stern-Volmer plot for the self-quenching of BIDP at 83 was also possible to subtract these monomer spectra from the
K. Squares: (φf(0))/(φf([M])), the ratio of the quantum yield at the lowest experimental absorbance bands to get the dimer line shapes (See
BIDP concentration used, to the QY at higher monomer concentrations, Figure 5b). The procedures used to find Keq at each temperature
vs nominal concentration. Straight line: a linear fit that gives KSV )
(1.5 ( 0.1) × 106 M-1. and the results of detailed calculations are listed in Table S3
(Supporting Information), and a plot of ln Keq versus 1/T is
shown in Figure 6. A straight line from 298 to 198 K yields
increased. In order to obtain quantitative data, the concentrations ∆Hdim ) 7 ( 2 kcal/mol. In the range from 198 to 83 K, Keq
of the dimer and monomer were determined in the total appears to be essentially constant.
concentration range (corrected for 83 K) 7.4 × 10-6 to 5.0 × Computational Results. The main results of the quantum
10-5 M using the separated spectra. From these data, an average chemical calculations are summarized in Table 1. In all cases,
value of Keq ) (4 ( 3) × 105 M-1 was derived using eq 3 (at the calculated dimer structures are more stable than two
83 K). In spite of the relatively large error, the basic assumption separated monomers by 0.3-14 kcal/mol. dimer1, dimer2, and
that dimer formation is the main cause for the spectral changes dimer3 are stable structures and have unique symmetry and/or
appears to be supported by the fact that the equilibrium constant distance constructions. Dimer1 is an in-line structure with Cs
calculated at 83 K for these various concentrations has a realistic symmetry and 4 Å distance between the nearest atoms. Both
value. dipoles of the monomers are aligned along the same direction,
Table S2 of the Supporting Information summarizes the leading to a slight stability, of 0.27 (0.36) kcal/mol using the
results for different nominal concentrations of BIDP. Due to B3LYP (BNL) functionals compared to the two separated
S/N ratio considerations, the monomer and dimer concentrations monomers. Dimer2, which is more stable than dimer1 by 1.6
could be reliably extracted by the subtraction method only for (7.7) kcal/mol at 4A distance, has antiparallel structure, which
solutions whose room temperature OD exceeded 0.2 (total probably increases the dipole interaction that stabilizes it. The
concentration > 4 × 10-6 M). The small variability of the corresponding parallel structure (not shown) is less stable than
equilibrium constant supports the hypothesis that dimerization two separate monomers at all distances checked, from 3 to 7
is the major aggregation process in the concentration range used Å. Dimer3 is a T structure with the two dipoles in antiparallel
in this work, and that formation of bigger adducts for the configuration, which is stabilized by 1.5 (2.5) kcal/mol. Dimer4
detected wavelengths may be neglected in the concentra- was the most stable structure found from several optimization
tion range used. Table 2 displays the results of separation of processes employing DFT/B3LYP starting at different dimer
the fluorescence to monomer and dimer contributions, using the geometries. Dimer4 is stable by 8 kcal/mol (14 kcal/mol) at
derived Keq. The quenching process was analyzed using the the B3LYP (BNL) level, and in solution (B3LYP/PCM) dimer4
Stern-Volmer equation (eq 1). Figure 4 shows that a linear is more stable by 5 kcal/mol. The MP2 calculations for the
relationship is obtained for the self-quenching process with KSV ground state strengthen those results and show the same range
) (1.5 ( 0.1) × 106 M-1. of stabilization energies for all dimers. For another comparison
Photophysics of a Polar Molecule in a Nonpolar Glass J. Phys. Chem. A, Vol. 114, No. 39, 2010 10569

Figure 5. (a) The dimer spectra as extrapolated from the difference between the total absorbance and the monomer DHO fit, at different temperatures
83K
([M]corr ) 3.3 × 10-5). (b) An example for deriving the dimer spectra in (a). Black: experimental BIDP absorbance at 163 K (other temperatures
at figure 1). Red: the monomer line shape based on the DHO model analysis (see text). Blue: the dimer spectrum as obtained from the difference
from the total absorbance and the monomer DHO fit, at 163 K.

of the monomer transitions was obtained in all dimers structures.


An allowed monomer electronic transitions was split to an
allowed and a forbidden transition, in line with Kasha’s model11
for dimer spectra.
The character of the transitions can be analyzed considering
the main orbitals involved in the transition. The orbitals of two
monomers coming close to each other interact to create dimer
orbitals. There are two types of dimers orbitals, (i) those formed
by combination or anticombination of monomer orbitals and
(ii) those in which the electronic distribution resides on one of
the monomers only (This can be viewed as a monomer orbital
“solvated” by the other monomer). Dimer4 was chosen for a
qualitative analysis as it is the most stable dimer and also has
both types of dimer orbitals, (i) and (ii), and as a result it also
has CT transitions. Scheme 2 shows the orbitals involved in
the main optical transitions of the monomer and in dimer4, at
Figure 6. A van’t Hoff plot for BIDP in MCHIP. The region from 298 the TDDFT/B3LYP/6-31G(d) level. On the left panel the
to 198 K yields ∆Hdim ) 7 ( 2 kcal/mol. See Table S3 for details.
dominant orbitals of the monomer in the S0-S1 and S0-S3
transitions are shown. Both involve transitions from orbital 50
6-311G(2d) basis set was also used in TDDFT/B3LYP calcula- to 52 and from 51 to 52, but with different weights. Both have
tions, and the stability energies are 10% larger at the most (Table a mixed character, partly ππ* and partly intramolecular CT.
S4). Since the differences are small for the large basis sets used, The shapes of the orbitals of dimer4 are displayed in the middle
the basis set superposition error (BSSE) correction is not panel. Orbitals 99 and 100 are dimer orbitals that are mainly
considered important (See ref 29). type (ii), originating from orbitals 50 of the two monomers,
Two trimer structures (based on dimer1 and dimer2, respec- while retaining the electronic distribution of the monomers.
tively) are presented in Table 1, using the B3LYP functional, Orbitals 101 and 102 are mainly type (ii), originating from
were found to show no extra stability. orbitals 51 of the two monomers and orbitals 103-104 are
Two allowed transitions of the monomer that have oscillator combination and anticombination of orbitals 52 in the monomers
strength larger than 0.1 were found by both TDDFT functionals, (type (i)). The dimer transitions S0-S1 (a), S0-S1 (b), S0-S3
B3LYP and BNL, as in previous CASSCF calculations (ref 1). (a), and S0-S3 (b) transitions have an intermolecular CT
The B3LYP functional predict that transitions S0-S1 and S0-S3 character, where the electronic density distribution is mostly
are allowed (in agreement with the CAS calculations), with on one molecule of the dimer in the ground state and is
oscillator strengths of 0.21 and 0.49 respectively. Using BNL transferred to an orbital in which the electronic distribution is
functional the allowed transitions are S0-S1 and S0-S2 with spread over the whole dimer. The type (ii) transitions split to
oscillator strengths of 0.50 and 0.18 respectively. A splitting allowed and less allowed transitions in the dimer, which are
10570 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

SCHEME 2: Schematic Orbital Energy Level Diagram

Left and right: the calculated orbital energies (Hartrees) of the dominant orbitals involved in the monomer and the S0-S1 and S0-S3 transitions.
Orbital shapes shown on left. Middle: the new orbitals of dimer4, B3LYP/6-31G(d). The nearly degenerate orbitals 99 and 100 are type (ii) orbitals
related to orbital 50 of the monomer, 101-102 are type (ii) orbitals related to orbital 51 of the monomer and 103-104 are combination and
anti-combination (type (i)) from orbital 52 in the monomer. The transitions originating in type (ii) orbitals split to allowed and less allowed transitions
in the dimer. The allowed and less allowed transitions are 600-800 cm-1apart. Transition S0-S1 (a), S0-S1 (b), S0-S3 (a), and S0-S3 (b) have
partial CT character.

separated by 600-800 cm-1 in dimer4. In other dimer structures


the shift between the allowed and less allowed transitions was
found to vary between 300 and 3300 cm-1, depending on the
geometry. CT and ππ* transitions (though not necessarily both
of them) were also found in these structures. The transitions’
character was found to be functional dependent. For further
details see Table S5 and S6.
Most of the calculated dimer transitions yield an allowed
transition that is blue-shifted with respect to the relevant
monomer transition according to both TDDFT methods. In
Figure 7, the monomer transitions are shown in black squares,
and other relative transitions of all calculated dimer structures
are shown by other black-gray symbols. For the B3LYP
calculations two trimers structures were also calculated, in
addition to dimer1 and dimer2, and transitions of trimer2 are
also shown in Figure 7. The calculated transitions must be
shifted by -7300 cm-1 (7200 cm-1) for B3LYP (BNL) in order Figure 7. The calculated oscillator strengths of electronic transitions
to coincide with the experimental ones (this shift was determined of the IDP monomer, dimer2 and trimer3 structures (See Table 2).
by the 490 nm band of the monomer experimental excitation Black: TDDFT/B3LYP/6-31G(d) level calculations, shifted by -7300
spectra). Figure 7 compares the experimental electronic spectra cm-1. Squares: monomer, circles: dimer2, triangles: trimer2. Gray:
of the dimer and the monomer with the calculated ones and TDDFT/BNL/6-31G(d) level calculations shifted by -7200 cm-1,
their oscillator strengths using the two functionals for dimer2. squares: monomer, circles: dimer2; compared to the experimental
monomer (black line) and dimer spectra (gray dashed).
The calculated transitions of the other structures can be seen in
detail in Tables S4 and S5. The calculated S0-S1 and S0-S3 505 nm, in addition dimer2 upon using the B3LYP functional,
(S2) transitions of the monomer were found to reproduce the was found in the calculation of dimer1 using BNL and in dimer4
monomer experimental absorption spectrum. The agreement structure using the B3LYP functional. Trimer1 and trimer2 (See
between the experimental dimer bands and the calculated ones Scheme S1 in the Supporting Information) also contribute to
is much poorer. The narrow red-shifted band at 505 nm appears the 505 nm band. A contribution to the blue side comes from
weakly by B3LYP calculations, and the broad blue-shifted band all other calculated dimers using both functionals, but again none
is also reproduced by B3LYP calculations, but with a smaller of the calculated transitions of the dimer reproduces quantita-
shift. In addition, no broadening mechanism that can predict tively the total dimer bands.
the large width of the band of only a single transition could be Using the DHO model for the monomer absorption spectra
simulated. A certain contribution to the experimental band at at 83 and 298 K we separated the two main contributions to
Photophysics of a Polar Molecule in a Nonpolar Glass J. Phys. Chem. A, Vol. 114, No. 39, 2010 10571

the absorption band, S0-S1 and S0-S3. The oscillator strength showing that internal conversion is the dominant process in
for S0-S1 was calculated to be 0.28 and for S0-S3 to be 0.25 MCHIP at room temperature.
(See Figure S2 and Table S1 for fitting details). These results It is evident that the fluorescence yield increases dramatically
are somewhat different from those reported in ref 1, but since as the temperature is decreased to less than 143 K (Figure S3).
they were obtained upon iterating between two temperatures, This temperature is close to the reported melting point of the
RT and 83 K, we consider the present ones as more reliable. mixture, 147 K,30 and strengthens the assumption that the
increased viscosity at the cryogenic temperatures inhibits BIDP
Discussion from reaching the geometry of the CI, a 90° torsion angle
In view of the presence of the dimer in the cryogenic glass between the two rings, and therefore increases the efficiency of
(at least in the case of MCHIP), the previous kinetic scheme the fluorescence. As shown in Figure S3, Fluorescence is
(ref 1) is now modified as follows: completely quenched at 143 K, a temperature at which dimer-
ization is quite extensive (essentially complete). At lower
temperatures, fluorescence increases, as expected for increased
M(S0) + hν f M(S*) (6)
rigidity of the solvent. Thus, we surmise that the main effect of
the cooling is due to the hindrance of the torsion in the
2M(S0) a D(S0) Keq Dimer formation (7) monomer. At these lower temperatures the dimer concentration
is large yet fluorescence is quite strong. Thus, the main effect
of cooling is the prevention of access to the conical intersection.
D(S0) + hν f D(S*) (8) From the analysis we derived Keq )(4 ( 3) × 105 M-1, a
value similar to that obtained for other systems such as dyes in
M(S*) f M(S0)) + hν k0f M fluorescence (9) nonpolar solvents.31 This is in line with a strong dipole-dipole
interaction that in nonpolar solvents and leads to more extensive
dimerization than in polar solvents. The blue shift of the dimer
M(S*) f M(S0) kIC M internal conversion (10) is 4000 cm-1, the upper limit of the shifts between the monomers
and dimers for the above-mentioned dyes.31 The derived Keq is
also of the same order as that reported for merocyanine dimers
D(S*) f D(S0)) + hν′ k′0
f D fluorescence (11) that are mainly H-aggregates (blue-shifted) and strongly
quenched.32
D(S*) f D(S0) ′ D internal conversion
kIC (12) The quenching process is analyzed by Stern-Volmer equa-
tion, yielding KSV ) (1.5 ( 0.1) × 106 M-1. Since KSV is much
larger than expected for dynamic quenching,13 and since
Formation of the triplet state is ignored since no evidence for dynamic quenching is not likely to be important in cryogenic
phosphorescence was observed experimentally. glasses, static quenching is proposed as the main process. We
It was shown earlier1 that in fluid nonpolar solvents internal adopt the “dark complex model” offered by Boaz and Rollef-
conversion is very efficient due to the presence of a CI that son,33 which suggests that the quenching is caused by a new
causes a very rapid decay to the ground state, on time scale of ground state complex, whose fluorescence quantum yield is
hundreds of femtoseconds. The decay is faster in nonpolar lower. The model also suggests that the complex constituents
solvents than in polar ones. Since now we have the full interact by van der Waals forces, in agreement with the value
characteristic of the fluorescence and absorptions in the nonpolar of 4 Å calculated for the distance between the components of
solvent, we can try to re-evaluate kf, the measured radiative rate the stable complex. (See also ref 34 and the QM calculation).
constant. Since the electronic absorption spectrum of the BIDP In most static quenching models, KSV should be very similar to
monomer actually consists of two overlapping bands, assigned Keq, but according to our analysis it is four times larger. There
to the S0-S1 and S0-S3 transitions, the DHO model was used is a likelihood of efficient energy transfer from the monomer
to evaluate (by eq 5) that at 83 K the pure radiative lifetime of to the dimer that can apparently increase KSV. Another possible
S1 is kf0 ) 1.4 × 108 s-1. kf, which is defined by eq 13, effect is the creation of an inhomogeneous concentration
distribution in the high-viscosity solvent, due to the tendency
kf ) k0f + kIC (13) of BIDP molecules to be closer to one another and create sites
of dimerization. In such situations the effective concentra-
tion of BIDP is higher than the average one, leading to
is related to kf0 by the monomer quantum yield (φ(0)) by: apparently more efficient quenching.
The temperature dependence (Figure 1) of BIDP absorbance
can be used to analyze at what temperature and how the dimers
φ(0) ) k0f /kf (14)
are formed. Two critical temperatures, T1 ) 147 K and T2 )
200 K appear to arise from the data. Similar values were noted
where (φ(0) signifies the quantum yield in the absence of self- by other experimental papers, such as Mandanici et al.,35 Angell
quenching). Using low concentration solutions, in which the et al.,36 and Abramczyk et al.,37,38 who assigned the lower one
dimer concentration can be neglected, the fluorescence quantum melting point and the higher one to a change from the liquid
yield is 0.9, and thus kf ) 1.6 × 108 s-1, replacing the results phase to a premelting structure, which affects the vibrational
presented in ref 1. At room temperature fluorescence in MCHIP dynamics of MCH. These temperatures match our experimental
is below the noise level, leading to an upper limit for the data. The overall decrease of the absorbance as the system
quantum yield of about 4 × 10-5, 1 order of magnitude less cooled to about 200 K can be explained by the dimer formation
than in acetonitrile, (4 × 10-4). Thus, at 298 K a lower limit and the change in the molar absorption coefficient of the
for kf is1.4 × 108/4 × 10-5 ) 3.5 × 1012 s-1. This large value monomer. (The absorption of all components appears to
is due primarily to kIC, which is estimated to be kIC = 1012 s-1, decrease, note that the spectra were rescaled to the volume
10572 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

TABLE 3: Summary of the Important Constants Evaluated in This Work


units MCHIP
, dielectric constant, 83 K 1.53b
melting point/glass point °K 146b
monomer dimer
, extinction coefficient cm-1 M-1 30 000 ( 3000 21 000 ( 2000
λmax, 83 K @457 nm @390 nm
fS0-S1, 83 K 0.28c
fS0-S3, 83 K 0.25c
fS0-S1, TDDFT/B3LYP 0.49 0.3-0.5d
fS0-S3, TDDFT/B3LYP 0.21 0.6-0.9d
kf0, (A01)c 83 K s-1 1.4 × 108
τ0)1/ kf0, c 83 K ns 7
a
Φfl, 83 K 83K
0.9 at [M]corr ) 7.2 × 10-6 M
KISC + kIC, 83 K using Φfl ) s-1 1.6 × 108
kfl/(kfl + kISC + kIC)
Φfl, RT e10-5
kf ()kIC) s-1 3.5 × 1012
τf fs 290
Κeq, 83 K M-1 (4 ( 3) × 105
Κsv, 83 K M-1 (1.5 ( 0.1) × 106
∆Η kcal/mol 7(2
a
Fluorescence quantum yield compared to Coumarine 540. b from ref 30. c Evaluated using DHO model and Strickler-Berg equation. d The
range for different dimer structures.

change). For concentrations from 7.4 × 10-6 to 5.0 × 10-5 M Every two monomer orbitals create two dimer orbitals, that
the spectra were reconstructed by using the line shape, the cell contribute to electronically excited states that are red or blue-
length, concentrations, and molar absorption coefficients of the shifted with respect to the monomer excited state. The optical
monomer and dimer, respectively transition to one of these is more allowed than to the other. In
Using the DHO model and the resulting separation between most of the structuressKasha’s in-line, parallel, and otherssthe
the monomer and the dimer at different temperature led to two allowed transitions are blue-shifted compare to the same
results. First, it is clear that the two bands of the dimer appear transition of the monomer. Only in two cases, dimer transitions
together (Figure 5a), a fact that was not clear from raw data. were found to be allowed red-shifted transitions: the S0-S3
The second result is the Van’t Hoff analysis, which apparently transition in dimer1, Kasha’s in line dimer, at the B3LYP level,
results in two slopes. In the range 298-198 K, ∆Hdim ) 7 ( 2 and the S0-S1 transition in dimer1 at the BNL level. Therefore,
kcal/mol is extracted from the data; this value is in very good the prediction made by Kasha that the structure is responsible
agreement to the calculated results for ∆E (Table 1, dimer4); for the red/blue shift is partly confirmed in the QM methods.
The agreement is reasonable in this case as the volume change The splittings calculated between the monomer and the dimer
is small. The resulting magnitude is similar to other molecules are in the range of 300-3300 cm-1, similar to the Kasha model
such as cyanines (10-17 kcal/mol).39 At lower temperatures estimates.
the slope is very small. It appears that the dimerization processes Two types of dimer transitions, CT transitions and ππ*, can
proceeds to equilibrium as long as the solvent is fluid, and once be discerned in the dimers; they may be rationalized by the
the solvent is frozen, further changes in spectrum are mainly nature of the dimer’s MOs. In some dimers the electronic
due to changes in the line shapes of the monomer and dimer distribution is spread over the entire structure, and in others it
spectra and not to changes in their relative concentrations. It is is found only on one monomer. Whereas the CT-type transitions
realistic to expect two regions since in the higher temperatures are blue-shifted only, ππ* transitions can be red or blue-shifted
range no solidification of the solvent occurs and dimerization (See Tables S5 and S6 for detailed information). DFT/BNL
can be take place, whereas in the lower temperatures region calculations are expected to yield better agreement with experi-
the relative motion of monomer molecules is inhibited. ment for CT states,21 but in the present case a shift of -7300
It is instructive to compare our results with Kasha’s exciton cm-1 (-7200 cm-1) was needed in the case of both functionals
model in which a structure wherein the two transition dipoles for agreeing with experiment (monomer). Similar differences
of the monomers are parallel produces two new excited states between experiments and TDDFT calculated transition energies
of which the blue-shifted state is allowed. An allowed excited where reported,40,41 and since we are interested in exploring the
state that is red-shifted with respect to the monomer appears transition behavior upon dimerization and not the absolute
for a dimer structure in which the two transition dipoles are numbers, no other action was needed except a linear correction.
in-line. A splitting yielding two allowed transitions occurs at Even upon applying the correction, the agreement between the
oblique transition moments. Kasha’s model predicts for van der calculated spectra to the experimental is rather limited. The large
Waals dimers 1000-2500 cm-1 blue shift compared to the blue shifted band is not reproduced in the calculations (Figure
monomer, followed by fluorescence quenching. In our case a 7), and attempts to include homogeneous and in-homogeneous
splitting to the red and to the blue was found, wherein the shift line width factors, using the DHO model, were not successful
to the blue is larger than the Kasha model prediction. Since in predicting the experimental line shape. It might be that there
according to the Kasha exciton model the dimer’s structure is is a special structure that produces this spectrum or some local
reflected in the resulting dimer absorption spectrum one can effects that shift the spectrum. Since the red and the blue bands
try to find if there is a preferred dimer structure of the BIDP. change similarly upon cooling they must arise from a single
The two new excited states predicted by the model are clearly process of formation. The trimer calculations offered by B3LYP
revealed in the QM calculations, as Scheme 2 shows for dimer4. functional, which show less stable structures, contribute to the
Photophysics of a Polar Molecule in a Nonpolar Glass J. Phys. Chem. A, Vol. 114, No. 39, 2010 10573

line shape only marginally on the red side, but contributions Farkas Center for Light Induced Processes is supported by the
from larger aggregates cannot be ruled out completely. An effort Minerva Gesellschaft mbH. We are grateful to Prof. Roie Bear’s
to perform DHO simulations for the dimer line shape or for the group and especially to Mrs. Tamar Stein, for discussion of the
total band line shape at 83 K with respect to the calculated BNL functional and for the BNL tuning.
transitions did not succeed; it was especially hard to reproduce
the large broadening of the blue-shifted band. Similarly large Supporting Information Available: Experimental spectra
shifts were recorded for dimers bonded by a hydrogen bond,40 of BIDP, tables of experimental parameters and results, and
but in our case they can be ruled out. schemes of the monomers, dimers, and trimers. This material
It is concluded that the computational methods used reproduce is available free of charge via the Internet at http://pubs.acs.org.
qualitatively the red and blue-shifted bands of the exciton model,
but quantitative agreement was not obtained. References and Notes
(1) Cogan, S.; Kahan, A.; Zilberg, S.; Haas, Y. J. Phys. Chem. A 2008,
112, 5604.
Conclusions (2) Cogan, S.; Haas, Y. J. Photochem. Photobiol., A 2008, 193, 25.
In this work it is shown that the low temperature absorption (3) Alfalah, S.; Deeb, O.; Zilberg, S.; Haas, Y. Chem. Phys. Lett. 2008,
459, 100.
of BIDP in nonpolar glasses can be quantitatively understood (4) Haas, Y.; Cogan, S.; Zilberg, S. Inter. J. Quan. Chem. 2005, 102,
if dimer formation is considered. Table 3 summarizes the 961.
important physical constants obtained in this work. Analysis of (5) Xu, X. F.; Kahan, A.; Zilberg, S.; Haas, Y. J. Phys. Chem. A 2009,
113, 9779.
the spectra at different concentrations and a careful separation (6) Squillacote, M.; Wang, J.; Chen, J. J. Am. Chem. Soc. 2004, 126,
between the contributions of the dimer and monomer lead us 1940.
to the determination of Keq at 83 K. Dimer formation is also (7) Speiser, S. Chem. ReV. 1996, 96, 1953.
accompanied by monomer fluorescence quenching, with KSV (8) Kim, K. S.; Tarakeshwar, P.; Lee, J. Y. Chem. ReV. 2000, 100,
4145.
)(1.5 ( 0.1) × 106 M-1, assigned to static quenching. (9) Chen, Z.; Lohr, A.; Saha-Moller, C. R.; Wurthner, F. Chem. Soc.
The low temperature monomer absorption line-shape allows ReV. 2009, 38, 564.
us to make a new DHO fit and to improve our previous estimate (10) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning,
A. P. H. J. Chem. ReV. 2005, 105, 1491.
of the fluorescence efficiently of the monomer. Working under (11) Kasha, M.; Rawls, H. R.; El-Bayoumi, M. A. Pure Appl. Chem.
conditions of negligible dimerization, internal conversion is 1965, 11, 371.
dominant in fluid solutions (kIC = 1012 s-1, compared to kf ≈ (12) Levinson, G. S.; Simpson, W. T.; Curtis, W. J. Am. Chem. Soc.
1957, 79, 4314.
108 s-1), whereas in the glassy solvent at 83 K the nonradiative (13) Eftink, M. R.; Ghiron, C. A. Anal. Biochem. 1981, 114, 199.
process is 10 times slower than the radiative one. (14) Boyd, G. V.; Ellis, A. W.; Harms, M. D. J. Chem. Soc. C 1970, 6,
The dimer line shape, which consists of a narrow red-shifted 800.
band and broad blue-shifted one, is in qualitative agreement (15) Berson, J. A.; Evleth, E. M.; Hamlet, Z. J. Am. Chem. Soc. 1965,
87, 2887.
with Kasha’s exciton model. QM calculations at the DFT level (16) Jones, G.; Jackson, W. R.; Choi, C. Y.; Bergmark, W. R. J. Phys.
were found to predict the two new optically allowed transitions Chem. 1985, 89, 294.
to excited states of the dimer. The two monomer MOs that (17) (a) Granovsky, A. A. PC GAMESS, 7.0 ed. http://classic.chem.msu.su/
gran/gamess/index.html (accessed Oct. 28, 2008). (b) Nemukhin, A. V.;
chiefly contribute to these dimer states interacting with each Grigorenko, B. L.; Granovsky, A. A. Moscow UniV. Chem. Bull. 2004, 45,
other were identified. The ability to predict the dimer spectrum 75.
from the QM calculation is very limited. (18) (a) Gill, P. M. W. Q-Chem 3.2; 2010. (b) Shao, Y.; Fusti-Molnar,
The temperature dependent measurements and the DHO L.; Jung, Y.; Kussmann, J.; Ochsenfeld, C.; Brown, S. T.; Gilbert, A. T. B.;
Slipchenko, L. V.; Levchenko, S. V.; O’Neill, D. P.; Distasio, R. A. Jr.;
analysis indicate the existence of two temperature regimes, in Lochan, R. C.; Wang, T.; Beran, G. J. O.; Besley, N. A.; Herbert, J. M.;
agreement with the solidification and crystallization process of Lin, C. Y.; Van Voorhis, T.; Chien, S. H.; Sodt, A.; Steele, R. P.; Rassolov,
MCHIP.37 In the higher temperature range (195-298 K) V. A.; Maslen, P. E.; Korambath, P. P.; Adamson, R. D.; Austin, B.; Baker,
J.; Byrd, E. F. C.; Dachsel, H.; Doerksen, R. J.; Dreuw, A.; Dunietz, B. D.;
dimerization takes place with an enthalpy of formation, ∆Hdim Dutoi, A. D.; Furlani, T. R.; Gwaltney, S. R.; Heyden, A.; Hirata, S.; Hsu,
) 7 ( 2 kcal/mol, in agreement with theoretical estimates. It is C.-P.; Kedziora, G.; Khalliulin, R. Z.; Klunzinger, P.; Lee, A. M.; Lee,
also found that the red and blue-shifted bands appear and change M. S.; Liang, W.; Lotan, I.; Nair, N.; Peters, B.; Proynov, E. I.; Pieniazek,
P. A.; Rhee, Y. M.; Ritchie, J.; Rosta, E.; Sherrill, C. D.; Simmonett, A. C.;
similarly as a function of temperature. This indicates that the Subotnik, J. E.; Woodcock, H. L. III; Zhang, W.; Bell, A. T.; Chakraborty,
dimerization process takes place before crystallization and A. K.; Chipman, D. M.; Keil, F. J.; Warshel, A.; Hehre, W. J.; Schaefer,
creates simultaneously the blue and the red-shifted bands in the H. F. III; Kong, J.; Krylov, A. I.; Gill, P. M. W.; Head-Gordon, M. Phys.
spectrum. The DHO model was found to be a useful and reliable Chem. Chem. Phys. 2006, 8, 3172-3191.
(19) Livshits, E.; Baer, R. Phys. Chem. Chem. Phys. 2007, 9, 2932.
tool for the monomer absorption at all temperatures, but not (20) Stein, T.; Kronik, L.; Baer, R. J. Chem. Phys. 2009, 131, 244119.
for the dimer. (21) Stein, T.; Kronik, L.; Baer, R. J. Am. Chem. Soc. 2009, 131, 2818.
The dimerization process occurs mainly in the temperature (22) Hobza, P.; Sponer, J.; Reschel, T. J. Comput. Chem. 1995, 16, 1315.
(23) Mishra, B. K.; Sathyamurthy, N. J. Phys. Chem. A 2004, 109, 6.
range between 298 and 195 K. Upon further cooling to 83 K (24) Miertus, S.; Scrocco, E.; Tomasi, J. Chem. Phys. 1981, 55, 117.
crystallization of the solvent sets in and inhibits free movement (25) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995.
of the solute. This prevents further dimerization on one hand, (26) Myers, A. B.; Mathies, R. A.; Tannor, D. J.; Heller, E. J. J. Chem.
and in addition hinders the torsional motion of the monomer, Phys. 1982, 77, 3857.
(27) Resonance Raman Intensities: A Probe of Excited-State Structure
which is required in order to reach to the geometry of conical and Dynamics; Myers, A. B., Mathies, R. A., Eds.; Wiley: New York, 1987;
interaction, making the fluorescence process more efficient then Vol. 2.
the IC. The new analysis corrects τf to be 2.9 × 10-13 s, smaller (28) Strickler, S. J.; Berg, R. A. J. Chem. Phys. 1962, 37, 814.
(29) Sodupe, M.; Bertran, J.; Rodriguez-Santiago, L.; Baerends, E. J. J.
than in polar solvent shown in ref 1, strengthening the main Phys. Chem. A 1998, 103, 166.
assumption that in nonpolar solvent internal conversion via the (30) Murov, S. L.; Carmichae, I.; Hug, G. L. Handbook of Photochem-
CI is more efficient. istry, 2 ed.; Marcel Dekker, Inc.: New York, Basel, Hong Kong, 1993.
(31) Lohr, A.; Grune, M.; Wurthner, F. Chem.sEur. J. 2009, 15, 3691.
(32) Rosch, U.; Yao, S.; Wortmann, R.; Wurthner, F. Angew. Chem.
Acknowledgment. A. K. thanks the Ministry of Science and Inter. Ed. 2006, 45, 7026.
Technology, for a Levi Eshkol Ph.D. Scholarship. The Minerva (33) Boaz, H.; Rollefson, G. K. J. Am. Chem. Soc. 1950, 72, 3435.
10574 J. Phys. Chem. A, Vol. 114, No. 39, 2010 Kahan et al.

(34) Dabkowska, I.; Jurecka, P.; Hobza, P. J. Chem. Phys. 2005, 122, (38) Abramczyk, H.; Paradowska-Moszkowska, K. J. Chem. Phys. 2001,
204322. 115, 11221.
(35) Mandanici, A.; Cutroni, M.; Triolo, A.; Rodriguez-Mora, V.; (39) Chibisov, A. High Energ. Chem. 2007, 41, 200.
Ramos, M. A. J. Chem. Phys. 2006, 125, 054514. (40) Strambi, A.; Durbeej, B. J. Phys. Chem. B. 2009, 113, 5311.
(36) Angell, C. A.; Sare, J. M.; Sare, E. J. J. Phys. Chem. 1978, 82, (41) Goerigk, L.; Grimme, S. ChemPhysChem 2008, 9, 2467.
2622.
(37) Abramczyk, H.; Paradowska-Moszkowska, K.; Wiosna, G. J. Chem.
Phys. 2003, 118, 4169. JP1047375

View publication stats

You might also like