0% found this document useful (0 votes)
45 views6 pages

Spectrochimica Acta Part A: Molecular and Biomolecular Spectros

Uploaded by

ROJITAS90
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
45 views6 pages

Spectrochimica Acta Part A: Molecular and Biomolecular Spectros

Uploaded by

ROJITAS90
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 6

Spectrochimica Acta Part A 77 (2010) 45–50

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

Quantum chemistry calculations of 3-Phenoxyphthalonitrile dye


sensitizer for solar cells
P. Senthil Kumar, K. Vasudevan, A. Prakasam, M. Geetha, P.M. Anbarasan ∗
Department of Physics, Periyar University, Salem 636 011, Tamilnadu, India

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

Article history: The geometries, electronic structures, polarizabilities, and hyperpolarizabilities of organic dye sensi-
Received 19 February 2010 tizer 3-Phenoxyphthalonitrile were studied based on Hartree–Fock (HF) and density functional theory
Received in revised form 1 April 2010 (DFT) using the hybrid functional B3LYP. Ultraviolet–visible (UV–vis) spectrum was investigated by time
Accepted 15 April 2010
dependent DFT (TD-DFT). Features of the electronic absorption spectrum in the visible and near-UV
regions were assigned based on TD-DFT calculations. The absorption bands are assigned to ␲ → ␲*
Keywords:
transitions. Calculated results suggest that the three excited states with the lowest excited energies
Dye sensitizer
in 3-Phenoxyphthalonitrile is due to photoinduced electron transfer processes. The interfacial electron
Density functional theory
Vibrational spectrum
transfer between semiconductor TiO2 electrode and dye sensitizer 3-Phenoxyphthalonitrile is due to
Electronic structure an electron injection process from excited dye to the semiconductor’s conduction band. The role of phe-
Absorption spectrum noxy group in 3-Phenoxyphthalonitrile in geometries, electronic structures, and spectral properties were
analyzed.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction relatively simple synthesis procedure, various structures and lower


cost in contrast to a ruthenium dye and the flexibility in molecular
The new technologies for direct solar energy conversion have tailoring of an organic sensitizer provides a large area to explore
gained more attention in the last few years. In particular, dye sen- [22–24]. Moreover, recently a rapid progress of organic dyes has
sitized solar cells (DSSCs) are promising in terms of efficiency and been witnessed reaching close to 10.0% efficiencies in combination
low cost [1–3]. The foremost feature of DSSC consists in a wide with a volatile acetonitrile-based electrolyte [25]. Phthalonitrile
band gap nanocrystalline film grafted with a quasi-monolayer of is an important class of high performance dyes, which are easily
dye molecules and submerged in a redox electrolyte. This ele- processable, and display good mechanical properties, outstanding
gant architecture can synchronously address two critical issues thermal and thermal-oxidative stability. Phthalonitrile dyes were
of employing organic materials for the photovoltaic applications: used for aerospace, marine, and electronic packaging applications.
(i) efficient charge generation from the Frenkel excitons and (ii) By thermal treatment of phthalonitrile derivatives at elevated tem-
long-lived electron–hole separation up to the millisecond time peratures (generally high up to 350 ◦ C) for an extended period of
domain. The latter attribute can often confer an almost quantitative time. In this paper the performance of 3-Phenoxyphthalonitrile
charge collection for several micrometer-thick active layers, even metal-free dye that can be used in DSSC is analyzed.
if the electron mobilities in nanostructured semiconducting films
are significantly lower than those in the bulk crystalline materi-
2. Experimental details
als. Benefited from systematic device engineering and continuous
material innovation, a state of the art DSSC with a ruthenium sen-
The compound 3-Phenoxyphthalonitrile was obtained from
sitizer has achieved a validated efficiency of 11.1% [4] measured
Sigma–Aldrich Chemical Company, USA, with a stated purity of
under the Air Mass 1.5 Global (AM1.5G) conditions. In view of the
greater than 99% and it was used as such without further purifi-
limited ruthenium resource and the heavy-metal toxicity, metal-
cation. The FT-Raman spectrum of 3-Phenoxyphthalonitrile has
free organic dyes have received surging research interest in recent
been recorded using 1064 nm line of Nd:YAG laser as excitation
years [5–21]. Because of their high molar absorption coefficient,
wavelength in the region 50–3500 cm−1 on a Brucker model IFS
66 V spectrophotometer. The FT-IR spectrum of this compound was
recorded in the region 400–4000 cm−1 on IFS 66 V spectropho-
∗ Corresponding author. Tel.: +91 0427 2345766/2345520; tometer using KBr pellet technique. The spectrum was recorded
fax: +91 0427 2345565/2345124. at room temperature, with scanning speed of 30 cm−1 min−1 and
E-mail address: [email protected] (P.M. Anbarasan). the spectral resolution of 2.0 cm−1 .

1386-1425/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.saa.2010.04.021
46 P.S. Kumar et al. / Spectrochimica Acta Part A 77 (2010) 45–50

3. Computational methods

The computations of the geometries, electronic structures,


polarizabilities and hyperpolarizabilities, vibrational frequencies
as well as electronic absorption spectrum for dye sensitizer 3-
Phenoxyphthalonitrile were done using Hartree–Fock (HF) and
density functional theory (DFT) with Gaussian 03 package [26].
The DFT was treated according to hybrid functional Becke’s
three parameter gradient-corrected exchange potential and the
Lee–Yang–Parr gradient-corrected correlation potential (B3LYP)
[27–29], and all calculations were performed without any sym-
metry constraints using polarized split-valence 6-311++G(d,p)
basis sets. The NBO analysis was performed using restricted
Hartree–Fock (RHF) with the same basis set. The electronic
absorption spectrum requires calculation of the allowed exci-
tations and oscillator strengths. These calculations were done
using time dependent DFT (TD-DFT) with the same basis sets
and exchange-correlation functional in vacuum and solution, and
the non-equilibrium version of the polarizable continuum model
(PCM) [30,31] was adopted for calculating the solvent effects. Fig. 2. The frontier molecular orbital energies and corresponding density of state
(DOS) spectrum of the dye 3-Phenoxyphthalonitrile.

4. Results and discussion


dye 3-Phenoxyphthalonitrile is shown in Fig. 2. The HOMO–LUMO
4.1. The geometric structure gap of the dye 3-Phenoxyphthalonitrile in vacuum is 4.47 eV.
While the calculated HOMO and LUMO energies of the bare
The optimized geometry of the 3-Phenoxyphthalonitrile is Ti38 O76 cluster as a model for nanocrystalline are −6.55 and
shown in Fig. 1, and the selected bond lengths, bond angles and −2.77 eV, respectively, resulting in a HOMO–LUMO gap of 3.78 eV,
dihedral angles are listed in supplementary Table S1. The crys- the lowest transition is reduced to 3.20 eV according to TD-DFT,
tal structure of the exact title compound is not available. The and this value is slightly smaller than typical band gap of TiO2
optimized structure can be compared with similar systems for nanoparticles with nm size [32]. Furthermore, the HOMO, LUMO
which crystal structures have been solved along with complete and HOMO–LUMO gap of (TiO2 )60 clusters is −7.52, −2.97, and
optimized bond lengths, bond angles and dihedral angles. The opti- 4.55 eV (B3LYP/VDZ), respectively [33]. Taking into account of the
mized bond lengths of C6–C16 and C8–C17 are 1.4261 and 1.4304 Å cluster size effects and the calculated HOMO, LUMO, HOMO–LUMO
respectively at B3LYP/6-311++G(d,p) and also well matched with gap of the dye 3-Phenoxyphthalonitrile, Ti38 O76 and (TiO2 )60 clus-
HF/6-311++G(d,p). ters, we can find that the HOMO energies of these dyes fall within
the TiO2 gap.
The above data also reveal the interfacial electron transfer
4.2. Electronic structures and charges
between semiconductor TiO2 electrode and the dye sensitizer 3-
Phenoxyphthalonitrile is electron injection processes from excited
Natural bond orbital (NBO) analysis was performed in order to
dye to the semiconductor conduction band. This is a kind of typical
analyze the charge populations of the dye 3-Phenoxyphthalonitrile.
interfacial electron transfer reaction [34].
Charge distributions in C, N and H atoms were observed because
of the different electro-negativity, the electrons transferred from
4.3. IR and Raman frequencies
C atoms to C, N atoms and C atoms to H. The natural charges of
different groups are the sum of every atomic natural charge in
Figs. 3 and 4 show the observed and calculated IR and Raman
the group. These data indicate that the cyanine and amide groups
spectra of 3-Phenoxyphthalonitrile respectively. Comparison of
are acceptors, while the acetic groups are donors, and the charges
the observed (FT-IR and FT-Raman) and calculated vibrational fre-
were transferred through chemical bonds. The frontier molecular
quencies of 3-Phenoxyphthalonitrile is shown in supplementary
orbitals (MO) energies and corresponding density of state of the
Table S2. Comparison of the frequencies calculated by HF and
B3LYP with experimental values reveals the overestimation of
the calculated vibrational modes due to neglect of anharmonic-
ity. A corrective vibrational scaling factor of 0.9613 to B3LYP
calculated frequencies and scaling factor of 0.8982 to HF cal-
culated frequencies were applied to account for anharmonicity.
Inclusion of electron correlation in density functional theory to
a certain extent makes the frequency values smaller in compari-
son with experimental values. Anyway notwithstanding the level
of calculations it is customary to scale down the calculated har-
monic frequencies in order to improve the agreement with the
experiment. The 3-Phenoxyphthalonitrile molecule give rise to
eleven C–C ring deformation, two C–C–N torsion, two C–C tor-
sion, seventeen C–C–N in plane bending, three C–C–N out of
plane bending, two C–H wagging, six C–H out of plane bend-
ing, one C–H torsion, nine C–H in plane bending, fourteen C≡N
symmetric stretching, three C N symmetric stretching, three C C
Fig. 1. Optimized geometrical structure of dye 3-Phenoxyphthalonitrile. stretching, one C C asymmetric stretching, two C≡N asymmet-
P.S. Kumar et al. / Spectrochimica Acta Part A 77 (2010) 45–50 47

Fig. 4. Observed and calculated FT-Raman spectra of 3-Phenoxyphthalonitrile.

3-Phenoxyphthalonitrile were calculated. Here, the polarizability


Fig. 3. Observed and calculated FT-IR spectra of 3-Phenoxyphthalonitrile.
and the first hyperpolarizabilities are computed as a numerical
derivative of the dipole moment using B3LYP/6-311++G(d,p). The
ric stretching and six C–H asymmetric stretching were assigned to definitions [36,37] for the isotropic polarizability is
B3LYP/6-311++G(d,p) scaling method. The strongest IR absorption 1
˛= 3 (˛XX + ˛YY + ˛ZZ )
for 3-Phenoxyphthalonitrile corresponds to the vibrational mode
48 near about 1301 cm−1 , which is corresponding to symmetri- The polarizability anisotropy invariant is
cal stretching mode of C≡N bond. The next stronger IR absorption
  12
is attributed to vibrational mode 56 near about 1609 cm−1 cor- (˛XX − ˛YY )2 + (˛YY − ˛ZZ )2 + (˛ZZ − ˛XX )2
responding to the stretching mode of C≡N bond. In the Raman ˛ =
2
spectrum, the strongest activity mode is the vibrational mode
60 near about 2342 cm−1 , which is corresponding to symmetrical and the average hyperpolarizability is
stretching mode of C≡N bond. The same vibrations computed by 1
 
ˇ = 5
ˇiiZ + ˇiZi + ˇZii
HF/6-311++G(d,p) and also show good agreement with experimen-
tal data. where ˛XX , ˛YY , and ˛ZZ are tensor components of polarizability;
ˇiiZ , ˇiZi , and ˇZii (i from X to Z) are tensor components of hyperpo-
4.4. Polarizability and hyperpolarizability larizability.
Tables 1 and 2 list the values of the polarizabilities and hyper-
Polarizabilities and hyperpolarizabilities characterize the polarizabilities of the dye 3-Phenoxyphthalonitrile. In addition to
response of a system in an applied electric field [35]. They deter- the individual tensor components of the polarizabilties and the
mine not only the strength of molecular interactions (long-range first hyperpolarizabilities, the isotropic polarizability, polarizabil-
intermolecular induction, dispersion forces, etc.) as well as the ity anisotropy invariant and hyperpolarizability are also calculated.
cross sections of different scattering and collision processes but The calculated isotropic polarizability of 3-Phenoxyphthalonitrile
also the nonlinear optical properties (NLO) of the system [36,37].
It has been found that the dye sensitizer hemicyanine system, Table 1
which has high NLO property, usually possesses high photo- Polarizability (˛) of the dye 3-Phenoxyphthalonitrile (in a.u.).
electric conversion performance [38]. In order to investigate the ˛XX ˛XY ˛YY ˛XZ ˛YZ ˛ZZ ˛ ˛
relationships among photocurrent generation, molecular struc-
−110.38 11.34 −103.88 0.76 0.0083 −91.54 101.93 16.57
tures and NLO, the polarizabilities and hyperpolarizabilities of
48 P.S. Kumar et al. / Spectrochimica Acta Part A 77 (2010) 45–50

Table 2
Hyperpolarizability (ˇ) of the dye 3-Phenoxyphthalonitrile (in a.u.).

ˇXXX ˇXXY ˇXYY ˇYYY ˇXXZ ˇXYZ ˇYYZ ˇXZZ ˇYZZ ˇZZZ ˇii

230.75 −50.29 5.63 −62.69 5.73 0.07 0.86 21.72 6.44 2.16 5.25

is 101.93 a.u. However, the calculated isotropic polarizability of Phenoxyphthalonitrile because of the agreement of line shape and
JK16, JK17, dye 1, dye 2, D5, DST and DSS is 759.9, 1015.5, 694.7, relative strength as compared with the vacuum and solvent.
785.7, 510.6, 611.2 and 802.9 a.u., respectively [39,40]. The above The HOMO–LUMO gap of 3-Phenoxyphthalonitrile in acetoni-
data indicate that the donor-conjugate ␲ bridge-acceptor (D-␲-A) trile at B3LYP/6-311++G(d,p) theory level is smaller than that in
chain-like dyes have stronger response for external electric field. vacuum. This fact indicates that the solvent effects stabilize the
Whereas, for dye sensitizers D5, DST, DSS, JK16, JK17, dye 1 and frontier orbitals of 3-Phenoxyphthalonitrile. In order to obtain the
dye 2, on the basis of the published photo-to-current conversion microscopic information about the electronic transitions, the cor-
efficiencies, the similarity and the difference of geometries, and the responding MO properties are checked. The absorption in visible
calculated isotropic polarizabilities, it is found that the longer the and near-UV regions is the most important region for photo-
length of the conjugate bridge in similar dyes, the larger the polar- to-current conversion, so only the twenty lowest singlet/singlet
izability of the dye molecule, and the lower the photo-to-current transitions of the absorption band in visible and near-UV regions
conversion efficiency. This may be due to the fact that the longer for 3-Phenoxyphthalonitrile are listed in Table 3. The data of Table 3
conjugate ␲ bridge enlarged the delocalization of electrons, thus it and Fig. 6 are based on the 6-311++G(d,p) results with solvent
enhanced the response of the external field, but the enlarged delo- effects involved.
calization may be not favorable to generate charge separated state This indicates that the transitions are photoinduced charge
effectively. So it induces the lower photo-to-current conversion transfer processes, thus the excitations generate charge separated
efficiency. states, which should favor the electron injection from the excited
dye to semiconductor surface. Commonly, the atom occupied by
4.5. Electronic absorption spectra and sensitized mechanism more densities of HOMO should have stronger ability for detaching
electrons, whereas the atom with more occupation of LUMO should
Electronic absorption spectra of 3-Phenoxyphthalonitrile in be easier to gain electrons. For 3-Aminophthalonitrile, the highest
vacuum and solvent were performed using TD-DFT(B3LYP)/6- occupied molecular orbital (HOMO) lying at −7.024 eV, is a delocal-
311++G(d,p) calculations, and the results are shown in Fig. 5. ized ␲ orbital. The HOMO-1, lying −7.56 eV below the HOMO, is a
It is observed that the absorption in the visible region is much delocalized ␲ orbital over the entire molecule. While the HOMO-2
weaker than that in the UV region for 3-Phenoxyphthalonitrile. and HOMO-3, lying −7.90, −8.04 eV below the HOMO, respectively,
The results of TD-DFT have an appreciable redshift in vacuum are ␲ orbitals that localized in benzene ring. Whereas, the low-
and acetonitrile solvent, and the degree of redshift in acetonitrile est unoccupied molecular orbital (LUMO), lying at −2.55 eV, is ␲*
solvent is more significant than that in vacuum. The discrep- orbital that localized. The LUMO+1, lying about −1.78 eV above the
ancy between vacuum and solvent effects in TD-DFT calculations
may result from two aspects. The first aspect is smaller gap of
materials which induces smaller excited energies. The other is sol-
vent effects. Experimental measurements of electronic absorptions
are usually performed in solution. Solvent, especially polar sol-
vent, could affect the geometry and electronic structure as well
as the properties of molecules through the long-range interaction
between solute molecule and solvent molecule. For these reasons
it is more difficult to make that the TD-DFT calculation is consis-
tent with quantitatively. Though the discrepancy exists, the TD-DFT
calculations are capable of describing the spectral features of 3-

Fig. 6. Isodensity plots (isodensity contour = 0.02 a.u.) of the frontier orbitals of 3-
Fig. 5. Calculated electronic absorption spectra of the dye 3-Phenoxyphthalonitrile. Phenoxyphthalonitrile and corresponding orbital energies (in eV).
P.S. Kumar et al. / Spectrochimica Acta Part A 77 (2010) 45–50 49

Table 3
Computed excitation energies, electronic transition configurations and oscillator strengths (f) for the optical transitions with f > 0.1 of the absorption bands in visible and
near-UV regions for the dye 3-Phenoxyphthalonitrile in acetonitrile.

State Configurations composition (corresponding transition orbitals) Excitation energy Oscillator


(eV/nm) strength (f)

1 0.69 (57 → 58) 3.71/333.9 0.0692


2 0.70 (56 → 58) 4.26/290.5 0.0176
3 0.27 (55 → 58) 0.63 (57 → 59) 4.50/275.4 0.0520
4 −0.26 (54 → 59) 0.57 (55 → 58) −0.27 (57 → 59) 4.61/268.7 0.0646
6 0.59 (54 → 58) 0.22 (55 → 59) −0.15 (56 → 59) 5.15/240.4 0.2021
7 −0.21 (56 → 59) −0.11 (56 → 60) 0.34 (56 → 61) 0.54 (57 → 60) 0.12 (57 → 61) 5.31/233.3 0.0191
8 −0.13 (54 → 58) 0.50 (55 → 59) 0.37 (57 → 61) 5.72/216.6 0.4645
9 0.37 (53 → 58) 0.49 (54 → 59) 0.16 (55 → 58) −0.12 (55 → 59) 5.92/209.4 0.2967
10 −0.11 (54 → 59) −0.28 (55 → 59) −0.13 (55 → 60) −0.32 (56 → 60) −0.16 (57 → 60) 0.37 (57 → 61) −0.18 (57 → 62) 5.94/208.5 0.1111
13 −0.17 (49 → 58) 0.54 (53 → 58) −0.30 (54 → 59) 6.22/199.0 0.3367
16 −0.29 (54 → 62) 0.12 (55 → 61) 0.18 (55 → 63) 0.12 (56 → 60) 0.53 (57 → 63) 6.46/191.7 0.0214
17 0.50 (55 → 61) 0.30 (56 → 60) 0.15 (56 → 61) 0.14 (57 → 61) −0.18 (57 → 63) 6.52/190.0 0.0693
18 0.21 (54 → 62) −0.12 (54 → 63) 0.43 (55 → 62) 0.12 (56 → 60) 0.38 (56 → 62) −0.12 (57 → 62) 6.60/187.7 0.0247
19 −0.11 (48 → 58) −0.15 (49 → 58) 0.12 (50 → 58) 0.60 (52 → 58) −0.10 (53 → 58) 0.14 (56 → 62) 6.63/186.0 0.0277

LUMO, is also a ␲* orbital that is similar to LUMO. The HOMO–LUMO results suggest that 3-Phenoxyphthalonitrile is a (D-␲-A) system.
gap of the dye 3-Aminophthalonitrile is 4.47 eV. The calculated isotropic polarizability of 3-Phenoxyphthalonitrile
The solar energy to electricity conversion efficiency () under is 101.93 a.u. The calculated polarizability anisotropy invariant of
AM1.5 white-light irradiation can be obtained from the following 3-Phenoxyphthalonitrile is 16.57 a.u. The hyperpolarizability of 3-
formula: Phenoxyphthalonitrile is 5.25 (in a.u.). The frequency of strongest
IR absorption for 3-Phenoxyphthalonitrile is 1301 cm−1 and the
Jsc [m A cm−2 ] × Voc [V] × ff
 (%) = × 100 frequency of strongest Raman activity for 3-Phenoxyphthalonitrile
I0 [m W cm−2 ]
is 2342 cm−1 . The electronic absorption spectral features in visi-
where I0 is the photon flux, Jsc is the short-circuit photocurrent ble and near-UV regions were assigned based on the qualitative
density, and Voc is the open-circuit photovoltage, and ff represents agreement to TD-DFT calculations. The absorptions are all ascribed
the fill factor [41]. At present, the Jsc , the Voc , and the ff are obtained to ␲ → ␲* transition. The three excited states with the low-
only by experiment, the relationship among these quantities and est excited energies of 3-Phenoxyphthalonitrile are photoinduced
the electronic structure of dye are still unknown. The analytical electron transfer processes that contributes sensitization of photo-
relationship between Voc and ELUMO may exist. According to the to-current conversion processes. The interfacial electron transfer
sensitized mechanism (electron injected from the excited dyes to between semiconductor TiO2 electrode and dye sensitizer 3-
the semiconductor conduction band) and single electron and single Phenoxyphthalonitrile is electron injection process from excited
state approximation, there is an energy relationship: dye as donor to the semiconductor conduction band. Based on the
analysis of geometries, electronic structures, and spectrum prop-
eVoc = ELUMO − ECB
erties of 3-Phenoxyphthalonitrile, the role of phenoxy group in
where ECB is the energy of the semiconductor’s conduction band phthalonitrile is as follows: it enlarged the distance between elec-
edge. So the Voc may be obtained applying the following formula: tron donor group and semiconductor surface, and decreased the
timescale of the electron injection rate, resulted in giving lower
ELUMO − ECB
Voc = conversion efficiency. This indicates that the choice of the appropri-
e
ate conjugate bridge in dye sensitizer is very important to improve
It induces that the higher the ELUMO , the larger the Voc . The the performance of DSSC.
results of organic dye sensitizer JK16 and JK17 [39], D-ST and D-SS
also proved the tendency [42] (JK16: LUMO = −2.73 eV, Voc = 0.74 V;
Acknowledgement
JK17: LUMO = −2.87 eV, Voc = 0.67 V; D-SS: LUMO = −2.91 eV,
Voc = 0.70 V; D-ST: LUMO = −2.83 eV, Voc = 0.73 V). Certainly, this
This work was partly financially supported by University Grants
formula expects further test by experiment and theoretical calcu-
Commission, Govt. of India, New Delhi, within the Major Research
lation. The Jsc is determined by two processes, one is the rate of
Project scheme under the approval-cum-sanction nos. F.No.34-
electron injection from the excited dyes to the conduction band
5\2008(SR) and 34-1/TN/08.
of semiconductor, and the other is the rate of redox between
the excited dyes and electrolyte. Electrolyte effect on the redox
processes is very complex, and it is not taken into account in the Appendix A. Supplementary data
present calculations. This indicates that most of the excited states
of 3-Phenoxyphthalonitrile have larger absorption coefficient, and Supplementary data associated with this article can be found, in
then with shorter lifetime for the excited states, so it results in the online version, at doi:10.1016/j.saa.2010.04.021.
the higher electron injection rate which leads to the larger Jsc of
3-Phenoxyphthalonitrile. On the basis of above analysis, it is clear References
that the 3-Phenoxyphthalonitrile has better performance in DSSC.
[1] B. O’Regan, M. Gratzel, Nature 353 (1991) 737–739.
[2] M. Gratzel, Nature 414 (2001) 338–344.
5. Conclusions
[3] N.G. Park, K. Kim, Phys. Status Solidi (a) 205 (2008) 1895–1904.
[4] Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide, L. Han, Jpn. J. Appl. Phys.
The geometries, electronic structures, polarizabilities, and 45 (2006) L638–L640.
hyperpolarizabilities of dye 3-Phenoxyphthalonitrile were stud- [5] K. Hara, M. Kurashige, Y. Dan-oh, C. Kasada, A. Shinpo, S. Suga, K. Sayama, H.
Arakawa, New J. Chem. 27 (2003) 783–785.
ied using HF and DFT with hybrid functional B3LYP, and the [6] T. Kitamura, M. Ikeda, K. Shigaki, T. Inoue, N.A. Anderson, X. Ai, T. Lian, S.
UV–Vis spectra were investigated using TD-DFT methods. The NBO Yanagida, Chem. Mater. 16 (2004) 1806–1812.
50 P.S. Kumar et al. / Spectrochimica Acta Part A 77 (2010) 45–50

[7] T. Horiuchi, H. Miura, K. Sumioka, S. Uchida, J. Am. Chem. Soc. 126 (2004) [26] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
12218–12219. J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar,
[8] W.M. Campbell, A.K. Burrell, D.L. Officer, K.W. Jolley, Coord. Chem. Rev. 248 J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson,
(2004) 1363–1379. H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T.
[9] K.R.J. Thomas, J.T. Lin, Y.C. Hsu, K.C. Ho, Chem. Commun. (2005) 4098–4100. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian,
[10] D.P. Hagberg, T. Edvinsson, T. Marinado, G. Boschloo, A. Hagfeldt, L. Sun, Chem. J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J.
Commun. (2006) 2245–2247. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A.
[11] S.L. Li, K.J. Jiang, K.F. Shao, L.M. Yang, Chem. Commun. (2006) 2792–2794. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels,
[12] N. Koumura, Z.S. Wang, S. Mori, M. Miyashita, E. Suzuki, K. Hara, J. Am. Chem. M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman,
Soc. 128 (2006) 14256–14257. J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A.
[13] S. Kim, J.W. Lee, S.O. Kang, J. Ko, J.H. Yum, S. Fantacci, F. De Angellis, D. Di Censo, Liashenko, P. Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham,
M.K. Nazeeruddin, M.J. Gratzel, J. Am. Chem. Soc. 128 (2006) 16701–16707. C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen,
[14] Z.S. Wang, Y. Cui, K. Hara, Y. Dan-oh, C. Kasada, A. Shinpo, Adv. Mater. 19 (2007) M.W. Wong, C. Gonzalez, J.A. Pople, Gaussian 03, Gaussian, Inc., Pittsburgh, PA,
1138–1141. 2003.
[15] T. Edvinsson, C. Li, N. Pschirer, J. Schoneboom, F. Eickemeyer, R. Sens, G. [27] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652.
Boschloo, A. Herrmann, K. Mullen, A. Hagfeldt, J. Phys. Chem. C 111 (2007) [28] B. Miehlich, A. Savin, H. Stoll, H. Preuss, Chem. Phys. Lett. 157 (1989) 200–206.
15137–15140. [29] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789.
[16] M. Wang, M. Xu, D. Shi, R. Li, F. Gao, G. Zhang, Z. Yi, R. Humphry-Baker, P. Wang, [30] V. Barone, M. Cossi, J. Phys. Chem. A 102 (1998) 1995–2001.
S.M. Zakeeruddin, M. Gratzel, Adv. Mater. 20 (2008) 4460–4463. [31] M. Cossi, N. Rega, G. Scalmani, V. Barone, J. Comput. Chem. 24 (2003) 669–681.
[17] D. Shi, N. Pootrakuchote, Z. Yi, M. Xu, S.M. Zakeeruddin, M. Gratzel, P.J. Wang, [32] M.K. Nazeeruddin, F. De Angelis, S. Fantacci, A. Selloni, G. Viscardi, P. Liska, S.
Phys. Chem. C 112 (2008) 17478–17484. Ito, B. Takeru, M. Gratzel, J. Am. Chem. Soc. 127 (2005) 16835–16847.
[18] G. Zhou, N. Pschirer, J.C. Schoneboom, F. Eickemeyer, M. Baumgarten, K. Mullen, [33] M.J. Lundqvist, M. Nilsing, P. Persson, S. Lunell, Int. J. Quant. Chem. 106 (2006)
Chem. Mater. 20 (2008) 1808–1815. 3214–3234.
[19] J.T. Lin, P.C. Chen, Y.S. Yen, Y.C. Hsu, H.H. Chou, M.C.P. Yeh, Org. Lett. 11 (2009) [34] D.F. Waston, G.J. Meyer, Annu. Rev. Phys. Chem. 56 (2005) 119–156.
97–100. [35] C.R. Zhang, H.S. Chen, G.H. Wang, Chem. Res. Chin. U 20 (2004) 640–646.
[20] G. Zhang, Y. Bai, R. Li, D. Shi, S. Wenger, S.M. Zakeeruddin, M. Gratzel, P. Wang, [36] Y. Sun, X. Chen, L. Sun, X. Guo, W. Lu, J. Chem. Phys. Lett. 381 (2003) 397–403.
Energy Environ. Sci. 2 (2009) 92–95. [37] O. Christiansen, J. Gauss, J.F. Stanton, J. Chem. Phys. Lett. 305 (1999) 147–155.
[21] M. Xu, S. Wenger, H. Bara, D. Shi, R. Li, Y. Zhou, S.M. Zakeeruddin, M. Gratzel, P. [38] Z.S. Wang, Y.Y. Huang, C.H. Huang, J. Zheng, H.M. Cheng, S.J. Tian, Synth. Met.
Wang, J. Phys. Chem. C 113 (2009) 2966–2973. 14 (2000) 201–207.
[22] X.H. Zhang, C. Li, W.B. Wang, X.X. Cheng, X.S. Wang, B.W. Zhang, J. Mater. Chem. [39] C.R. Zhang, Y.Z. Wu, Y.H. Chen, H.S. Chen, Acta Phys. Chim. Sin. 25 (2009) 53–60.
17 (2007) 642–649. [40] A. Seidl, A. Gorling, P. Vogl, J.A. Majewski, M. Levy, Phys. Rev. B 53 (1996)
[23] M. Liang, W. Xu, F. Cai, P. Chen, B. Peng, J. Chen, Z. Li, J. Phys. Chem. C 111 (2007) 3764–3774.
4465–4472. [41] K. Hara, T. Sato, R. Katoh, A. Furube, Y. Ohga, A. Shinpo, S. Suga, K. Sayama, H.
[24] W. Xu, B. Peng, J. Chen, M. Liang, F. Cai, J. Phys. Chem. C 112 (2008) 874–880. Sugihara, H. Arakawa, J. Phys. Chem. B 107 (2003) 597–606.
[25] S. Ito, H. Miura, S. Uchida, M. Takata, K. Sumioka, P. Liska, P. Comte, P. Pechy, [42] C.R. Zhang, Z.J. Liu, Y.H. Chen, H.S. Chen, Y.Z. Wu, L.H. Yuan, J. Mol. Struct. 899
M. Gratzel, Chem. Commun. (2008) 5194–5196. (2009) 86–93.

You might also like