Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Optical Materials 34 (2012) 1013–1018

Contents lists available at SciVerse ScienceDirect

Optical Materials
journal homepage: www.elsevier.com/locate/optmat

Two-photon absorption in oxazole derivatives: An experimental and quantum


chemical study
D.L. Silva a, L. De Boni b,⇑, D.S. Correa c, S.C.S. Costa d, A.A. Hidalgo d, S.C. Zilio b, S. Canuto a, C.R. Mendonca b
a
Instituto de Física, Universidade de São Paulo, Caixa Postal 66318, 05314-970 São Paulo, SP, Brazil
b
Instituto de Física de São Carlos, Universidade de São Paulo, Caixa Postal 369, 13560-970 São Carlos, SP, Brazil
c
Embrapa Instrumentação, Rua XV de Novembro 1452, Caixa Postal 741, 13560-970 São Carlos, SP, Brazil
d
Departamento de Física, Universidade Federal do Piauí, Av. Ministro Petrônio Portella s/n, Ininga 64.049-550, Teresina, PI, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Experimental and theoretical studies on the two-photon absorption properties of two oxazole deriva-
Received 10 August 2011 tives: 2,5-diphenyloxazole (PPO) and 2-(4-biphenylyl)-5-phenyl-1,3,4-oxadiazole (PBD) are presented.
Received in revised form 7 December 2011 The two-photon absorption cross-section spectra were determined by means of the Z-scan technique,
Accepted 15 December 2011
from 460 up to 650 nm, and reached peak values of 84 GM for PBD and 27 GM for PPO. Density Functional
Available online 18 January 2012
Theory and response function formalism are used to determine the molecular structures and the one- and
two-photon absorption properties and to assist in the interpretation of the experimental results. The
Keywords:
Polarizable Continuum Model in one-photon absorption calculations is used to estimate solvent effects.
Two-photon absorption
Quantum chemical calculation
Ó 2011 Elsevier B.V. All rights reserved.
Z-Scan technique

1. Introduction lowering, because of the use of longer excitation wavelengths


and the low absorption outside the focal volume. Materials pre-
Oxazole and Oxadiazole are related to a class of five-membered senting two-photon absorption have found applications in two-
nitrogen heterocyclic ring, where an oxygen and a nitrogen atoms photon polymerization [9–12], upconversion fluorescence [13–
are separated by one carbon. Oxazole derivatives are known to dis- 16], optical limiting [17–19] and photodynamic therapy [20,21].
play strong absorption and luminescence in the UV–Vis spectrum, The assessment of the electronic structure of molecules by one or
and can be used as electron transporting materials for optical de- two photon absorption spectroscopies and concomitant studies
vices, [1] being already used in applications of fluorescence micros- at the quantum chemistry level may reveal distinct molecular
copy, UV photodetectors, lasing dyes [2–4] and as active layer in charge transfer processes. These properties may help elucidating
organic photodiodes [5]. Also, the p-electron conjugation of the electronic transport in organic devices.
oxadiazole derivatives could be used to manufacture nonlinear Some results, depicting 2PA and three-photon absorption (3PA)
optical materials [6]. However, studies reporting multi-photon in oxazole derivatives, can be found in the literature. For example,
absorption spectrum of these compounds, for applications in mul- three-photon induced fluorescence have been reported by Gry-
ti-photon fluorescence microscopy and optical limiting are still czynski and co-workers [22] in 2,5-diphenyloxazole using a single
needed. Besides, characterizing the material’s multi-photon wavelength excitation (870 nm). Moreover, Lakowicz et al. [23]
absorption is also important to calculate the material figure of have presented two-photon absorption in 2,5-diphenyloxazole
merit, which is relevant when dealing with all-optical switches dissolved in different solvents. In that case, they observed a
[7,8]. fluorescence signal induced by an absorption of simultaneous
Simultaneous two-photon absorption (2PA) can be achieved by two-photons at 584 nm [22]. They noticed that the fluorescence
exposing the molecules to an intense light source. In this situation, signal presents the same shape as the one excited by one-photon.
an electron from the ground state is promoted to an excited state, Although both works have presented fluorescence anisotropy sig-
where each individual photon has half of the energy required for nals related to multi-photon absorption, quantification of the mul-
the optical transition. The interesting features achieved by 2PA in- ti-photon absorption cross-sections and the spectrum were not
clude spatial confinement of the absorption and scattering losses obtained for these molecules. In this way, we report here an
experimental and theoretical investigation on the two-photon
⇑ Corresponding author. Address: Instituto de Física de São Carlos, Universidade absorption (2PA) properties of two oxazole derivatives: 2-(4-bi-
de São Paulo, P.O. Box 369, 13560-970 São Carlos, SP, Brazil. Fax: +55 16 33738085. phenylyl)-5-phenyl-1,3,4-oxadiazole and 2,5-diphenyloxazole,
E-mail address: deboni@ifsc.usp.br (L. De Boni). PBD and PPO [24], respectively. The 2PA cross-section spectra from

0925-3467/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.optmat.2011.12.009
1014 D.L. Silva et al. / Optical Materials 34 (2012) 1013–1018

460 up to 650 nm of both compounds were obtained by means of delivers 150 fs pulses at 775 nm, with 1 kHz repetition rate. Such
the Z-scan technique, by which a threefold peak value was found pulses are employed to pump an optical parametric amplifier,
for PBD (84 GM) in comparison to PPO (27 GM). A theoretical study which provides 120 fs pulses in the spectral range from 460 to
employing quantum chemical calculations based on the Density 700 nm with the same repetition rate (1 KHz). To achieve a Gauss-
Functional Theory (DFT) [25,26] determined the equilibrium ian beam profile we use spatial filtering. The output signal was
molecular geometry and the electronic transitions of the mole- monitored using a silicon photo-detector coupled to a lock-in
cules. Computations employing the linear and quadratic response amplifier that integrates 1000 shots for each point of the Z-scan
functions [27] determined the one- and two-photon absorption trace. The laser pulse energies and beam waist sizes used in this
properties of the two oxazole derivatives and helped explaining experiment range from 70 to 150 nJ and 14 to 17 lm, respectively,
the experimental results. Moreover, the Time Dependent DFT according to the excitation wavelength.
(TDDFT) method [28–30] and the Polarizable Continuum Model
(PCM) [31–33] were employed in this study to estimate the effect 2.2. Computational details
of the solvent on the electronic transitions.
All the quantum chemical calculations were carried out using
2. Experimental and computational details the DFT model [25,26]. The equilibrium geometries of the mole-
cules were obtained by employing the hybrid B3LYP [35] [36] func-
2.1. Optical measurements tional and the 6-311G(d, p) basis set [37] as implemented in the
Gaussian 03 program [38]. From the optimized geometries, the
The chemical structures of PPO and PBD are displayed in Fig. 1. one- and two-photon transitions were determined by using the re-
Both were purchased from Sigma Aldrich, dissolved in dichloro- sponse functions formalism [27], within DFT as implemented in
methane, and placed in a 2-mm long fused silica cuvette for optical the DALTON program [39]. In this approach, the oscillator
measurements. strengths and two-photon transition probabilities are calculated
The linear absorption spectra of both compounds were obtained analytically as single residues of the linear and quadratic response
by a Cary 17 spectrometer, by using solutions with concentration in functions of the molecular electronic density, respectively.
the order of 105 mol L1. In order to investigate if molecular aggre- The two-photon absorption calculations were carried out in
gation occurred in the high concentration solution used in the non- vacuum and, in an effort to obtain an estimate of the spectral shift
linear experiment, we measured the linear absorption using a of the electronic transitions due to solvent effects, the one-photon
homemade silica cuvette with path length of the order of tens of mi- absorption calculations were also carried out in vacuum and in sol-
crons. No visual changes were observed in the bands position of the vent medium employing the TDDFT method [28–30] for the
high concentration sample, when compared with the low concen- absorption and the PCM [31–33] for obtaining the solvent effects,
tration one, indicating the non-occurrence of aggregation. The non- as implemented in the Gaussian 03 program [38]. All electronic
linear absorption spectrum of the compounds (with concentration transition calculations were carried out using the hybrid PBE0
in the order of 101 mol L1) were obtained by means of the Z-scan functional [40,41] and the 6-31+G(d) basis set [37].
technique [34]. This technique is able to provide the two-photon In the experimental technique used in this work, the two-pho-
absorption coefficient of a given material by moving the sample ton absorption is measured through the dissipation of the incident
through the focal plane of a Gaussian laser beam and monitoring light, which for a single beam 2PA experiment is twice the transi-
intensity changes in the far field. In the case of a two-photon tion rate. In the case, the 2PA cross-section (rgf) of a given excited
absorption material, the total absorption coefficient, now intensity state (in the two-photon resonant condition), for a degenerate pro-
dependent, can be given by a = a0 + bI, where I is the laser beam cess, is written as [42–44]
irradiance, a0 is the linear absorption coefficient, and b is the two-
photon absorption coefficient. As a result, b can be determined by 16p3 aa50 ðhxÞ2  
rgf ðxÞ ¼ d ð2Þ
monitoring changes in the transmittance while the sample is c pðCf =2Þ gf
scanned through the focal position. By integrating the transmitted
where a is the fine structure constant, a0 is the Bohr’s radium, c is the
power over time, and assuming a temporal Gaussian pulse, one
speed of light and hx is the energy of each photon absorbed (half of
can obtain the normalized energy transmittance according to [34]:
Z the transition energy). Cf is the damping constant describing full
1
1 2 width at half-maximum of the final state linewidth (assuming a
T ¼ pffiffiffiffi ln½1 þ q0 ðz; 0Þes ds ð1Þ
pq0 ðz; 0Þ 1 Lorentzian line-shape) and hdgfi is the two-photon transition proba-
bility for the transition from the ground state (g) to a final state (f).
where q0 ðz; tÞ ¼ bI0 ðtÞLð1 þ z2 =z20 Þ1 , L is the sample thickness, z0 the
The degenerate two-photon transition probability induced by a
Rayleigh length, z the sample position and I0 the pulse irradiance.
linearly polarized laser beam in molecules at liquid environments
The two-photon absorption b can be determined by fitting the
(isotropic media) is given by: [45,46]
experimental Z-scan curves using Eq. (1). By using r ¼ hmb=N,
where hm is the excitation photon energy, and N is the number of   1 X gf  gf   
dgf ¼ 2Saa Sbb þ 4Sgf gf
ab Sab ð3Þ
molecules per cm3, the 2PA cross-section, r, can be calculated. r 30 a;b
values are normally expressed in Goeppert–Mayer units, where,
1 GM = 1  1050 cm4 s photon1. The excitation source in our in which the subscripts a and b represent the Cartesian coordinates
experiment is a Ti:sapphire chirped pulse amplified system that and Sgf
ab is the two-photon matrix element defined as:

(a) (b)
N N
N

O
O

Fig. 1. Chemical structures of (a) 2,5-diphenyloxazole (PPO) and (b) 2-(4-biphenylyl)-5-phenyl-1,3,4-oxadiazole (PBD).
D.L. Silva et al. / Optical Materials 34 (2012) 1013–1018 1015

 
1 X hgj^e  l
^ a jkihkj^e  l
^ b jf i hgj^e  l
^ b jkihkj^e  l
^ a jf i two-photon wavelength (nm)
Sgf
ab ¼ þ ð4Þ
2
h k xgk  x xgk  x 400 500 600 700 800
1.0 30
for a single beam 2PA experiment [42–44]. (a)

2PA cross-section (GM)


normalized absorbance
0.5 15
3. Results and discussion

The optimized molecular geometry of 2,5-diphenyloxazole


0.0 0
(PPO) and 2-(4-biphenylyl)-5-phenyl-1,3,4-oxadiazole (PBD), ob-
1.0 100
tained by B3LYP/6-311G(d, p) calculations in vacuum, are dis-
played in Fig. 2. Both compounds present a virtually planar (b)
geometry, but PBD presents a longer conjugation length owing to
0.5 50
an extra phenyl ring located at one of its end and which is about
35° out of plane to avoid steric hindrance. It is worth pointing
out that, in addition to the conjugation length increase, the inclu-
sion of a nitrogen atom to the central ring of PBD also alters the 0.0 0
200 250 300 350 400
electron delocalization throughout its structure, making it more
one-photon wavelength (nm)
polarizable. The equilibrium geometries were also obtained in sol-
vent medium (dichloromethane) employing the PCM model. In Fig. 3. Normalized absorbance (solid line) and two-photon absorption cross-section
both cases, gas phase and solvent, the oxazole derivatives present spectra (circles) of PPO (a) and PBD (b) in dichloromethane. The solid line along the
a molecular conformation where the torsion angles among their experimental 2PA cross-section is a guide to the eye.
aromatics rings are minimal, the only exception being the extra
phenyl ring of PBD. Due to the slight influence of the calculated sol-
vent effect on the equilibrium geometries of the oxazole deriva- derivatives studied here can be accessed via both processes due to
tives investigated, the ones obtained in vacuum were adopted for their noncentrosymmetric character.
the electronic transition calculations. We observe in Fig. 3, especially for PBD (Fig. 3b), a subtle shoul-
To determine the two-photon absorption cross-section spectra, der in the 2PA spectrum (290–300 nm) in the vicinity of the peak of
we obtained Z-scan curves in the spectral range from 460 to the linear absorption band at 310 nm. The shift between the subtle
650 nm, which were fitted by means of Eq. (1). By performing shoulder in the 2PA spectrum and the peak in the 1PA spectrum
the same procedure for every spectral component, one can obtain most likely points out that distinct ground and first excited states
the two-photon absorption spectra of PPO and PBD in a large spec- vibrational modes are being coupled in the electronic transition to
tral range, which are displayed in Fig. 3a and b, respectively (cir- this excited state if the transition is induced via one- or two-pho-
cles). The thin solid line along the experimental data (circles) is a ton absorption. As a consequence, and as already reported in liter-
guide to the eye. The 2PA cross-section spectra, displayed in ature [50–53], the oscillator strength and the 2PA transition
Fig. 3 (circles) show that PPO and PBD present moderate values, probability of the transitions accessing this excited state are redis-
compared to other organic molecules reported in the literature tributed among its 0–0 and vibronic states in a slightly different
[18,47–49]. The 2PA cross-section peak value obtained for PPO way, what results in the spectral shift (1640 cm1) between the
and PBD are 27 and 84 GM, respectively. Since both oxazole deriv- subtle shoulder in the 2PA spectrum and the peak of the linear
atives have a virtually planar equilibrium geometry (Fig. 2), the absorption band. In addition, both molecules present a 2PA peak
higher 2PA cross-section peak of PBD in comparison to PPO can around 260 nm, nearly 50 nm blue-shifted from the 1PA peak.
be directly explained by its longer conjugation length, which To aid the interpretation of the experimental 2PA properties of
yielded a more pronounced charge delocalization and leads to an PPO and PBD, we have performed a theoretical study using the re-
enhancement of nonlinear absorption [18,47–49]. sponse functions formalism, to determine the electronic transi-
For molecules with center of symmetry (symmetric molecules), tions, either related to one- or two-photon absorption, of these
the selection rules for the one- and two-photon absorption pro- oxazole derivatives. Moreover, to estimate the solvent effect on
cesses are complementary, i.e., the rules state that one- and two- the electronic transitions TDDFT calculations were also performed
photon allowed transitions involve initial and final electronic employing the PCM model with dichloromethane as the solvent. As
states with opposite and the same parities, respectively. both PPO and PBD are polar molecules (with ground state dipole
Conversely, for molecules without center of symmetry (noncentro- moment of 1.49 D and 3.60 D, respectively) and dichloromethane
symmetric molecules) the parity of the electronic states is not is a nonprotic solvent, it is expected that the PCM model may give
well-defined and such selection rules are relaxed. Therefore, for a good estimate of the solvent effects on the calculated transition
molecules without center of symmetry transitions via 1PA and energies. The theoretical results for PPO and PBD, i.e., transitions
2PA can occur to the same final excited state. As it will be pointed energies (E), oscillator strengths (f) and 2PA transition probabilities
out by the theoretical results, the first excited state of the oxazole (P), are displayed in Tables 1 and 2 respectively.

Fig. 2. Optimized molecular geometries of PPO (a) and PBD (b) determined in vacuum by B3LYP/6-311G(d, p) calculations.
1016 D.L. Silva et al. / Optical Materials 34 (2012) 1013–1018

Table 1
Theoretical results of the one- and two-photon absorption calculations for PPO. 1PA: TD-DFT method and PCM model as implemented in Gaussian 03 package. 2PA: quadratic
response function within the DFT framework as implemented in Dalton program. All the transition calculations were carried out using the PBE0/6-31+G(d) model. Transition
energies (E) in electron-volts, oscillator strengths (f) and 2PA transition probabilities (prob.) in atomic units (a.u.). Also shown is the percentage contribution of each excitation
(column 2) to the electronic transitions reported.

Trans. Contributions 1PA (Vacuum) 1PA (PCM-Dichloromethane) 2PA (Vacuum)


E (eV) f E (eV) f E (eV) Prob. (a.u.)
1 HOMO ? LUMO, 50% 3.91 0.860 3.79 0.972 3.91 0.427E+03
(317 nm) (327 nm) (317 nm)
2 HOMO-2 ? LUMO, 0.7% 4.63 0.010 4.65 0.026 4.63 0.240E+04
HOMO ? LUMO + 1, 13.5% (268 nm) (267 nm) (268 nm)
HOMO ? LUMO + 2, 1.3%
HOMO ? LUMO + 3, 0.5%
3 HOMO-1 ? LUMO, 3.9% 4.73 0.003 4.73 0.002 4.73 0.102E+04
HOMO-2 ? LUMO, 0.2% (262 nm) (262 nm) (262 nm)
HOMO ? LUMO + 2, 8.8%
HOMO ? LUMO + 3, 1.0%
4 HOMO-3 ? LUMO, 0.2% 4.87 0.022 4.86 0.028 4.87 0.229E+05
HOMO-2 ? LUMO, 0.3%
HOMO-1 ? LUMO, 0.3%
HOMO ? LUMO + 1, 6.5%
HOMO ? LUMO + 1, 0.2%
HOMO ? LUMO + 3, 3.9% (254 nm) (255 nm) (254 nm)
5 HOMO-2 ? LUMO, 0.2% 5.30 0.035 5.27 0.037 5.30 0.216E+03
HOMO-1 ? LUMO, 11.5% (234 nm) (235 nm) (234 nm)
HOMO ? LUMO + 2, 1.3%
HOMO ? LUMO + 3, 2.9%
6 HOMO-3?LUMO, 0.4% 5.33 0.029 5.29 0.053 5.33 0.448E+04
HOMO-2 ? LUMO, 14.6% (233 nm) (234 nm) (233 nm)
HOMO-1 ? LUMO, 0.1%
HOMO ? LUMO + 2, 1.0%
HOMO ? LUMO + 3, 0.5%
7 HOMO ? LUMO + 4, 21.2% 5.45 0.001 5.41 0.190 5.45 0.419E+04
(227 nm) (229 nm) (227 nm)
8 HOMO-3 ? LUMO, 24.5% 5.48 0.125 5.57 0.002 5.48 0.584E+01
HOMO-2 ? LUMO, 0.2% (226 nm) (223 nm) (226 nm)
HOMO ? LUMO + 3, 0.3%

Table 2
Theoretical results of the one- and two-photon absorption calculations for PBD. 1PA: TD-DFT method and PCM model as implemented in Gaussian 03 package. 2PA: quadratic
response function within the DFT framework as implemented in Dalton program. All the transition calculations were carried out using the PBE0/6-31+G(d) model. Transition
energies (E) in electron-volts, oscillator strengths (f) and 2PA transition probabilities (prob.) in atomic units (a.u.). Also shown is the percentage contribution of each excitation
(column 2) to the electronic transitions reported.

Trans. Contributions 1PA (Vacuum) 1PA (PCM-Dichloromethane) 2PA (Vacuum)


E (eV) f E (eV) f E (eV) Prob. (a.u.)
1 HOMO ? LUMO, 48.0% 3.93 1.222 3.85 1.354 3.93 0.734E+03
(315 nm) (322 nm) (315 nm)
2 HOMO-4 ? LUMO, 3.9% 4.66 0.023 4.64 0.056 4.66 0.255E+05
HOMO ? LUMO + 1, 7.2% (266 nm) (267 nm) (266 nm)
HOMO ? LUMO + 2, 2.0%
3 HOMO-4 ? LUMO, 1.3% 4.73 0.005 4.70 0.027 4.73 0.354E+05
HOMO-1 ? LUMO, 8.0% (262 nm) (264 nm) (262 nm)
HOMO ? LUMO + 1, 1.4%
HOMO?LUMO + 2, 1.0%
4 HOMO-2 ? LUMO, 15.7% 4.83 0.023 4.75 0.049 4.83 0.310E+03
HOMO-2 ? LUMO + 1, 0.2% (257 nm) (261 nm) (257 nm)
HOMO-1 ? LUMO, 0.5%
HOMO-4 ? LUMO + 1, 0.2%
HOMO ? LUMO + 4, 0.5%
5 HOMO-4 ? LUMO, 0.3% 4.84 0.036 4.77 0.004 4.84 0.547E+04
HOMO-3 ? LUMO, 0.2% (256 nm) (260 nm) (256 nm)
HOMO-2 ? LUMO, 0.7%
HOMO-1 ? LUMO, 7,2%
HOMO ? LUMO + 1, 1.6%
6 HOMO-3 ? LUMO, 9.7% 4.85 0.003 4.82 0.005 4.85 0.320E+03
HOMO-3 ? LUMO + 1, 0.5% (255 nm) (257 nm) (255 nm)
HOMO-1 ? LUMO + 3, 0.2%
HOMO ? LUMO + 1, 0.7%
HOMO ? LUMO + 3, 1.0%
7 HOMO-5 ? LUMO, 46.1% 4.95 0.000 5.20 0.000 4.95 0.922E+02
(250 nm) (238 nm) (250 nm)
8 HOMO-4 ? LUMO, 0.3% 5.31 0.006 5.31 0.011 5.31 0.974E+02
HOMO-2 ? LUMO, 2.9% (234 nm) (233 nm) (234 nm)
HOMO ? LUMO + 2, 2.9%
HOMO ? LUMO + 4, 5.1%
D.L. Silva et al. / Optical Materials 34 (2012) 1013–1018 1017

As it can be seen in Table 1, the theoretical calculations deter- states. In fact, even if one decides to add the 2PA probabilities of
mined that the first strongly allowed 1PA transition of PPO occurs these almost degenerate transitions, considering in this way some
at 317 nm in vacuum and at 327 nm in dichloromethane solvent. superposition effect in this spectral region, it still seems that the
This result indicates that the excited state accessed by this transi- total probability is not enough to explain the experimental peak.
tion stabilizes more than the ground state of the molecule in the Therefore, we believe that the theoretical calculations most proba-
solvent medium and therefore the transition is red-shifted. How- bly underestimated the probabilities of such 2PA transitions at the
ever, the shift due to solvent effects (about 10 nm) can be consid- vicinity of the experimental peak for PBD. However, the experi-
ered small. mental measurements determined the 2PA cross-section of the ex-
The experimental data shows that the 1PA spectrum of PPO pre- cited states, which is proportional to the 2PA probabilities, but it is
sents a band centered at 312 nm and, therefore, we can conclude also proportional to the square of the transition energy and inver-
that the present theoretical calculation taking into account the sol- sely proportional to the linewidth of the excited state, Eq. (2). The
vent underestimates the energy of this strongly allowed 1PA tran- value of the linewidth of the excited states is an important point
sition in about 0.18 eV. The theoretical result for the energy of this here, but the absence of such information for the investigated oxa-
transition of PPO in vacuum is in good agreement with the exper- zole derivatives prevent us from doing a direct quantitative com-
imental data in dichloromethane. In general, considering the en- parison between experimental (cross-sections) and theoretical
ergy and oscillator strength of the 10 lowest electronic (probabilities) values. Finally, in general it is observed that the
transitions of PPO we can conclude that the solvent effects do 2PA probabilities determined by the theoretical calculations for
not critically affect the one-photon absorption of PPO, what PBD are higher than the ones provided for PPO. Due to the virtually
encourages the discussion of the experimental nonlinear spectrum planar equilibrium geometry of both molecules, the higher proba-
of PPO using the results of the 2PA calculation of PPO in vacuum. bilities of PBD might be directly explained by its longer conjuga-
The 2PA calculation determined that the four lowest electronic tion length, a feature that usually increases the nonlinearity of
states are allowed via two-photon transition and from an energy molecules [18,47–49].
analysis we can infer that those are excited states accessed along Fig. 4 presents the molecular orbitals involved in the low-lying
the spectral region investigated experimentally. The first state is electronic transitions, displayed in Tables 1 and 2. These molecular
only slightly allowed via 2PA and corresponds to the same state orbitals were determined from the theoretical calculations in vac-
strongly allowed via 1PA. The second and third 2PA allowed states, uum, but the PCM calculations showed that the molecular orbital
with corresponding transition energies of 4.63 eV (268 nm) and changes due to solvent effects are negligible. We observe that for
4.73 eV (262 nm) respectively, have small calculated intensities both compounds all these molecular orbitals are of p character.
for the 1PA. The fourth state, corresponding to a weak transition The first electronic transition of both compounds is described
via 1PA, is the strongest two-photon allowed state and the energy by an excitation from the highest occupied molecular orbital
of the corresponding transition is in excellent agreement with the
position of the peak of the experimental 2PA spectrum (at 512 nm)
of PPO molecule. Finally, it is worth noting that the calculated
changes in the energy of the transitions for these last three two-
photon allowed states due to solvent effects are negligible.
From the results of the 1PA calculations for PBD, Table 2, one
can observe that the first strongly allowed 1PA transition occurs
at 315 nm in vacuum and at 322 nm in dichloromethane solvent,
while the experimental linear absorption spectrum presents an in-
tense band at 307 nm. The solvent effects slightly shift this transi-
tion to the lower energy region (red-shift), as also observed for PPO
molecule. From an analysis of the energy and oscillator strength of
all one-photon transitions determined we can conclude that the
solvent effects do not critically affect the one-photon absorption
of PBD molecule. In other words, in addition to the slight changes
of the transition energies and the small changes of the oscillator
strengths, we also note that the character of the electronic transi-
tions and the molecular orbitals involved in each of them are not
critically affect by solvent effects. This is confirmed analyzing the
nature of the transitions, presented on Table 2.
The 2PA calculations for PBD determined seven electronic states
along the spectral region investigated experimentally, being the six
lowest states important to understand the nonlinear spectrum. The
first excited state, strongly allowed via 1PA, is only slightly allowed
via 2PA. The second and third excited states are strongly two-pho-
ton allowed and separated by only 0.07 eV. Following the theoret-
ical calculations, we determined three excited states almost
degenerate. The fourth, fifth and sixth excited states are two-pho-
ton allowed and the energy of the transitions to these excited
states are in excellent agreement with the position of the peak of
the experimental 2PA spectrum (at 514 nm) of PBD. Up to this en-
ergy region, and except for the first transition, all the remaining
transitions are insensitive to solvent effects.
The 2PA probabilities (Table 2) of the three almost degenerate
transitions of PBD, which are at the region of the experimental Fig. 4. Molecular orbitals involved in the excitations describing the electronic
peak, are smaller than the probabilities of the second and third transitions of the oxazole derivatives.
1018 D.L. Silva et al. / Optical Materials 34 (2012) 1013–1018

(HOMO) to the lowest unoccupied molecular orbital (LUMO), Ta- [6] G. Hughes, D. Kreher, C.S. Wang, A.S. Batsanov, et al., Org. Biomol. Chem. 2
(2004) 3363–3367.
bles 1 and 2, and therefore correspond to p ? p transitions. For
[7] F. Yoshino, S. Polyakov, G.I. Stegeman, Appl. Phys. Lett. 84 (2004) 5362–5364.
PPO and PBD, most of the following transitions are described by [8] I. Fuks-Janczarek, J. Luc, B. Sahraoui, F. Dumur, et al., J. Phys. Chem. B 109
excitations involving a rich combination of molecular orbitals. (2005) 10179–10183.
From the analysis of the molecular orbitals involved in the calcu- [9] S. Kawata, H.B. Sun, T. Tanaka, K. Takada, Nature 412 (2001) 697–698.
[10] W. Haske, V.W. Chen, J.M. Hales, W.T. Dong, et al., Opt. Express 15 (2007)
lated spectra (Tables 1 and 2 and Fig. 4), it can be seen that the cal- 3426–3436.
culated transitions for both PPO and PBD are of p ? p character. [11] C.R. Mendonca, D.S. Correa, F. Marlow, T. Voss, et al., Appl. Phys. Lett. 95 (2009)
This characteristic may be used to understand the small red shift 113309.
[12] A. Ovsianikov, J. Viertl, B. Chichkov, M. Oubaha, et al., ACS Nano 2 (2008) 2257–
of the transitions in the solvent environment. However, some of 2262.
the excitations present some charge transfer, as it is the case of [13] D.S. Correa, L. De Boni, D.T. Balogh, C.R. Mendonca, Adv. Mater. 19 (2007)
the HOMO-1 ? LUMO and HOMO-2 ? LUMO of PPO. However, 2653–2656.
[14] I. Cohanoschi, K.D. Belfield, F.E. Hernandez, Chem. Phys. Lett. 406 (2005) 462–
as these two excitations contribute to the same transition energy 466.
they partially cancel each other, and the excited states may be [15] G.S. He, P.P. Markowicz, T.C. Lin, P.N. Prasad, Nature 415 (2002) 767–770.
interpreted as simple p ? p states. Similar analysis can be also [16] X.M. Wang, P. Yang, W.L. Jiang, G.B. Xu, et al., Opt. Mater. 27 (2005) 1163–
1170.
made for PBD. [17] K.A. Belfield, M.V. Bondar, F.E. Hernandez, O.V. Przhonska, J. Phys. Chem. C 112
(2008) 5618–5622.
4. Conclusions [18] S.L. Oliveira, D.S. Correa, L. Misoguti, C.J.L. Constantino, et al., Adv. Mater. 17
(2005) 1890–1893.
[19] J.E. Ehrlich, X.L. Wu, I.Y.S. Lee, Z.Y. Hu, et al., Opt. Lett. 22 (1997) 1843–1845.
We determined the two-photon absorption cross-section spec- [20] J.D. Bhawalkar, G.S. He, P.N. Prasad, Rep. Prog. Phys. 59 (1996) 1041–1070.
tra of two oxazole derivatives, PPO and PBD, from 460 to 650 nm. [21] J. Liu, Y.W. Zhao, J.Q. Zhao, A.D. Xia, et al., J. Photochem. Photobiol. B-Biol. 68
(2002) 156–164.
Although both compounds display moderate values of 2PA cross- [22] I. Gryczynski, H. Malak, J.R. Lakowicz, Chem. Phys. Lett. 245 (1995) 30–35.
section, PBD presents a higher 2PA cross-section (84 GM) com- [23] J.R. Lakowicz, I. Gryczynski, Z. Gryczynski, E. Danielsen, et al., J. Phys. Chem. 96
pared to PPO (27 GM) which is a direct outcome of its larger con- (1992) 3000–3006.
[24] F.N. Hayes, B.S. Rogers, D.G. Ott, J. Am. Chem. Soc. 77 (1955) 1850–1852.
jugation length provided by an extra phenyl ring, since both [25] P. Hohenberg, W. Kohn, Phys. Rev. B 136 (1964) B864-B871.
compounds have a virtually planar equilibrium geometry, as [26] W. Kohn, L. J. Sham, Physical Review 140 (1965) A1133-A1138.
pointed out by quantum chemical computations. The theoretical [27] P. Salek, O. Vahtras, J.D. Guo, Y. Luo, et al., Chem. Phys. Lett. 374 (2003) 446–
452.
calculations, within the Density Functional Theory model, yielded
[28] R. Bauernschmitt, R. Ahlrichs, Chem. Phys. Lett. 256 (1996) 454–464.
one- and two-photon transition properties in accordance with [29] R.E. Stratmann, G.E. Scuseria, M.J. Frisch, J. Chem. Phys. 109 (1998) 8218–8224.
the experimental results and identified the set of active electronic [30] M.E. Casida, C. Jamorski, K.C. Casida, D.R. Salahub, J. Chem. Phys. 108 (1998)
4439–4449.
states along the spectral region investigated. Besides, analyses con-
[31] S. Miertus, E. Scrocco, J. Tomasi, Chem. Phys. 55 (1981) 117–129.
sidering the molecular orbitals of the compounds determined the [32] B. Mennucci, J. Tomasi, J. Chem. Phys. 106 (1997) 5151–5158.
p ? p character of all electronic transitions. Finally, the solvent ef- [33] V. Barone, M. Cossi, J. Tomasi, J. Chem. Phys. 107 (1997) 3210–3221.
fects on the calculated electronic transitions were estimated using [34] M. Sheik-Bahae, A.A. Said, T.H. Wei, D.J. Hagan, et al., IEEE J. Quantum Electron.
26 (1990) 760–769.
the Polarizable Continuum Model and seen to be small for the low- [35] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652.
est lying transition and negligible for the remaining excited states [36] C.T. Lee, W.T. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789.
for both oxazole derivatives. [37] M.J. Frisch, J.A. Pople, J.S. Binkley, J. Chem. Phys. 80 (1984) 3265–3269.
[38] M.J. Frisch, H.B. Schlegel, G.E. Scuseria, M.A. Robb, et al., Gaussian 03, in: P.
Pittsburgh (Ed.), Gaussian, Pittsburgh, PA, 2003.
Acknowledgments [39] DALTON a molecular electronic structure program. Release 2.0, 2005. <http://
www.kjemi.uio.no/software/dalton/dalton.html>.
[40] C. Adamo, G.E. Scuseria, V. Barone, J. Chem. Phys. 111 (1999) 2889–2899.
We thank financial support from Fapesp (Fundação de Amparo [41] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865–3868.
a Pesquisa do Estado de São Paulo), CNPq (Conselho Nacional de [42] R.W. Boyd: Nonlinear optics, Academic Press, Boston, 1992.
Desenvolvimento Científico), CAPES (Coordenação de Aperfeiçoa- [43] K. Ohta, K. Kamada, J. Chem. Phys. 124 (2006).
[44] K. Ohta, L. Antonov, S. Yamada, K. Kamada, J. Chem. Phys. 127 (2007).
mento de Pessoal de Nível Superior), FINEP (Financiadora de Estu-
[45] W.M. McClain, J. Chem. Phys. 55 (1971) 2789–2790.
dos e Projetos), FAPEPI (Fundação de Amparo à Pesquisa do Estado [46] P.R. Monson, W.M. McClain, J. Chem. Phys. 53 (1970) 29–30.
do Piauí), UFPI (Universidade Federal do Piauí), from Brazil. [47] M. Albota, D. Beljonne, J.L. Bredas, J.E. Ehrlich, et al., Science 281 (1998) 1653–
1656.
[48] K.D. Belfield, A.R. Morales, B.S. Kang, J.M. Hales, et al., Chem. Mat. 16 (2004)
References 4634–4641.
[49] J.M. Hales, J. Matichak, S. Barlow, S. Ohira, et al., Science 327 (2010) 1485–
[1] K.Y. Kwon, X. Lin, G. Pawin, K. Wong, et al., Langmuir 22 (2006) 857–859. 1488.
[2] C.U. Murade, V. Subramaniam, C. Otto, M.L. Bennink, Biophys. J. 97 (2009) 835– [50] P.R. Callis, Ann. Rev. Phys. Chem. 48 (1997) 271–297.
843. [51] D. Beljonne, J.L. Bredas, M. Cha, W.E. Torruellas, et al., J. Chem. Phys. 103 (1995)
[3] Q. Dai, X.Q. Zhang, Opt. Express 18 (2010) 11821–11826. 7834–7843.
[4] A.N. Fletcher, R.A. Henry, R.F. Kubin, R.A. Hollins, Opt. Commun. 48 (1984) [52] S. Ohira, I. Rudra, K. Schmidt, S. Barlow, et al., Chem.-Eur. J. 14 (2008) 11082–
352–356. 11091.
[5] S.C.S. Costa, M.A.L. Reis, J. Del Nero, L.S. Roman, Revista de Propriedade [53] E. Kamarchik, A.I. Krylov, J. Phys. Chem. Lett. 2 (2011) 488–492.
Industrial 1943 (2008) 87.

You might also like