310-JMES-2033-2015-El Assyry
310-JMES-2033-2015-El Assyry
310-JMES-2033-2015-El Assyry
ISSN : 2028-2508
CODEN: JMESCN
Abstract
B3LYP/6-311G(d,p) quantum calculations are performed to optimize geometries and obtain properties depending
on the electronic and photovoltaic for some quinoxalinone derivatives. Novel twelve organic donor-π-acceptor
dyes (D-π-A), used for dye-sensitized solar cells (DSSC), based on quinoxalin-2(1H)-one were studied by density
functional theory (DFT) and time dependant DFT (TDDFT) approaches to shed light on how the π-conjugation
order influence the performance of the dyes. The electron acceptor group was 2-cyanoacrylic for all dyes whereas
the electron-donor unit varied and the influence was investigated. The theoretical results have shown that TDDFT
calculations using the Coulomb attenuating method CAM-B3LYP with the polarized split-valence 6-311G(d,p)
basis sets and the polarizable continuum model (PCM) were reasonably capable of predicting the excitation
energies, the absorption and the emission spectra of the molecules. The HOMO and LUMO energy levels of
these dyes can ensure a positive effect on the process of electron injection and dye regeneration. The trend of the
calculated HOMO–LUMO gaps nicely compares with the spectral data. Key parameters in close connection with
the short-circuit current density (Jsc), including light-harvesting efficiency (LHE), injection driving force (ΔGinject)
and total reorganization energy (λtotal) were discussed. In addition, the estimated values of open-circuit
photovoltage (Voc) for these dyes were presented. The calculated results of these dyes reveal that the Q4 dye can
be used as a potential sensitizer for TiO2 nanocrystalline solar cells due to its best electronic and optical
properties and good photovoltaic parameters.
1. Introduction
Considering the environmental issue and renewable resources, more and more attentions are paid to the solar
energy utilizations. Dye-sensitized solar cells (DSSC) have attracted an ever-increasing attention in scientific
research and in practical applications since the first report by O’Regan and Grätzel in 1991, because of its
potential advantages, such as low cost and highly efficient conversion of sunlight into electricity [1-3]. In
particular, these DSSC are composed of a wide band gap semi-conductor (typically TiO2) sensitized with
molecular dyes, able to capture light in the visible region of the spectrum, and a redox electrolyte (typically
Iodide/triiodide I-/I-3) [4-6]. In DSSC, incoming light causes electronic excitations of the dye sensitizers leading
to electrons injection to the conduction band of nanocrystalline metal oxide. Then, the dyes regain electrons from
redox couple in electrolyte solution [1]. In general, a power conversion efficiency dye sensitizer has the
following characteristics: the highest occupied molecular orbital (HOMO) energy must be located below the
HOMO energy of electrolyte to accept the electron from a redox electrolyte pair (I-/I-3), and the lowest
unoccupied molecular orbital (LUMO) should have a higher energy than that of the conduction band of semi-
conductor (TiO2). The sensitive dyes play an important role for DSSC to gain the higher solar-to-electricity
conversion efficiency that has been an active research subject, recently [7-9]. In this context, two different kinds
2612
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
of dye sensitizers are used in DSSC device: metal organic complexes and metal-free organic dyes. The metal
organic complexes (e.g. the Ruthenium dyes such as N719 and N3) [10, 11] are the most promising sensitizers
for DSSC and show a record power conversion efficiency of 11% under air mass (AM) 1,5 irradiation [12].
However, due to the limited availability and environment issues of the noble-metal Ruthenium, metal-free
organic dyes have been developed as alternatives. Therefore, the metal free organic dye sensitizers such as
coumarin [13, 14], indoline [15], triphenylamines [16-18], perylene and fluorene [19, 20], thienopyrazine [21]
have also been developed and exhibited satisfactory performance. The most extensively studied organic dyes
usually adopt the donor-π spacer-acceptor (D-π-A) structural motif, which exhibit several advantages: high molar
extinction coefficients, low cost of production, and an extraordinary diversity [22]. In this structure, the
intramolecular charge transfer (ICT) from D to A at the photoexcitation will inject the photoelectron into the
conduction band of the semi-conductor through the electron accepting group at the anchoring unit. By changing
the electron donor, acceptor, and/or -π-spacer group, the HOMO and LUMO energy levels are affected [23].
The quinoxalinone is rigid, coplanar and electron-accepting fused heterocycle due to the electron-withdrawing
nitrogen of imine (C=N) and giving rise to a highly extended π-electron system [24]. Recent researches showed
that the quinoxalinone system is a promising candidate as a core unit for high performance semi-conductors [25].
quinoxaline-based acceptors and thiophene- and pyrrole-based donors to create alternating D-A oligomers,
working towards low band gap, ambipolar materials for OFET applications [25]. Recently, the study of synthesis,
structural and energetic properties of quinoxalines has been done by our group [26-33]. More recently, some
quinoxaline-based polymers were synthesized for polymer solar cell applications, and power conversion
efficiency (PCE) up to 6–7% was reported [34-37]. Relative to quinoxaline-based polymers, quinoxaline-based
small molecules photovoltaic materials remain rare [38-41]. Recently, a theoretical study of photovoltaic
properties on a series of D-π-D molecules with quinoxalinone as π-spacer and different number of donor unit as
photoactive components of organic bulk heterojunction (BHJ) solar cells; was reported [41]. Despite the large
body of work outlined above, to the best of our knowledge the employment of quinoxaline as photosensitizers for
DSSC remains rare [42].
Nowadays, theoretical quantum calculations have been effective tools in the field of chemistry because they can
be used to rationalize the properties of known chemical compounds and also predict those of unknown ones to
guide observed experimental synthesis [43]. In contrast to experimental results for metal-free organic dyes, the
theoretical investigations are still limited. Only few research groups have studied the electronic structures and
photophysical properties of dye sensitizers [42-44], and intra molecular electron dynamic process between dyes
and TiO2 nanocrystalline [45, 46]. In this context, the quinoxalinone-based organic dyes Qi (i=1–12) whose
chemical structures are shown in Fig.1 were synthesized according to the method described in the literature [47].
The central quinoxalinone was paired through conjugation to an heterocyclic or phenylene as donor groups and
linked to substituent as acceptor/anchoring group forming D-π-A architecture.
In this work, the electronic structure and optical absorption properties of twelve dye sensitizers (Fig.1) were
calculated by using DFT and TDDFT. Based upon the calculated results, we analyzed the role of different
electron-donor groups in the tuning the geometries, electronic structures and optical properties. Also, we aimed to
see the sensitizer donor effects on the open circuit photovoltage (Voc) and the short-circuit current density (Jsc) of
the cell through discussing the key factors affecting Voc and Jsc with the goal of finding potential sensitizers for
use in DSSC.
2. Theoretical methodology
Theoretical background
The power conversion efficiency (η) is mainly determined by the short-circuit current density (Jsc) and the open
circuit photovoltage (Voc). The η can be expressed by the following equation [48]:
𝑉 𝐽
𝜂 = 𝐹𝐹 𝑜𝑐 𝑠𝑐 (1)
𝑃𝑖𝑛𝑐
Where FF is the fill factor and Pinc is the incident solar power on the cell. From this expression, Jsc, Voc, and FF
are only obtained by the experiment. The relationship among these values and electronic structures for the dyes is
still unknown. Herein, we aimed to see the sensitizer donor effects on the Voc and Jsc of the cell through
discussing the key factors affecting them.
For the short-circuit current density Jsc in DSSC, it is determined as:
𝐽𝑠𝑐 = 𝜆
𝐿𝐻𝐸 𝜆 𝛷𝑖𝑛𝑗𝑒𝑐𝑡 𝜂𝑐𝑜𝑙𝑙𝑒𝑐𝑡 𝑑𝜆 (2)
2613
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
where 𝐿𝐻𝐸 𝜆 is the light harvesting efficiency. 𝛷𝑖𝑛𝑗𝑒𝑐𝑡 is the electron injection efficiency, and 𝜂𝑐𝑜𝑙𝑙𝑒𝑐𝑡 is the
charge collection efficiency. For the same DSSC with only different dyes, it is reasonable to assume that the
𝜂𝑐𝑜𝑙𝑙𝑒𝑐𝑡 is a constant. As a result, to shed light onto the relationship between the Jsc and η theoretically, we
investigated the LHE, 𝛷𝑖𝑛𝑗𝑒𝑐𝑡 and total reorganization energy (λtotal). From Eq. (2), to obtain a high Jsc, the
efficient sensitizers applied in DSSC should have a large LHE, which can be expressed as [49]:
LHE = 1 – 10-f (3)
where f is the oscillator strength of the dye associate to the wavelength λmax in the equation. We noticed that the
larger oscillating strength obtained, the higher light-harvesting efficiency will have. At the same time, a large
𝛷𝑖𝑛𝑗𝑒𝑐𝑡 based on Eq. (2) could also guarantee a high Jsc, which is related to the injection driving force ΔGinject and
evaluated by [49]: ΔGinject = Edye* + ECB (4)
dye*
where E is the oxidation potential energy of the dye in the excited state and ECB is the reduction potential of
the conduction band of TiO2 in Eq. (4), respectively. There, we use in this work ECB = –4.0 eV for TiO2 [50],
which is widely used in some papers [49, 51, 52], and the Edye* can be estimated [51-53] by:
Edye* = Edye – E00 (5)
dye
where E is the oxidation potential energy of the dye in the ground state, while E00 is an electronic vertical
transition energy corresponding to the λmax. It is generally accepted that there are two schemes to evaluate the
ΔGinject, that is relaxed and unrelaxed paths. The previous works concluded that the calculation with the unrelax
path is reliable [48, 52, 53]. Thus, the electron injection from excited state of dye to the TiO2 (CB) is determined
by the unrelax path in our investigation.
Additionally, as mentioned in Eq. (2), the small total reorganization energy (λtotal) which contains the hole and
electron reorganization energy could enhance the Jsc. Namely, the smaller λtotal value obtained, the faster charge-
carrier transport rates will be [49]. So we computed the hole and the electron reorganization energy (λh and λe)
according to the following formula [54]:
𝜆𝑖 = 𝐸0± − 𝐸±± + 𝐸±0 − 𝐸𝟢 (6)
where 𝐸0 is the energy of the cation or anion calculated with the optimized structure of the neutral molecule, 𝐸±±
±
is the energy of the cation or anion calculated with the optimized cation or anion structure, 𝐸±0 is the energy of the
neutral molecule calculated at the cationic or anionic state, and the 𝐸𝟢 is the energy of the neutral molecule at
ground state.
To analyze the relationship between Voc (Open-circuit photovoltage) and the energy of LUMO (ELUMO) of the
dyes based on electron injection from LUMO to the conduction band (ECB) of the semi-conductor TiO2 the Voc (in
eV) can be approximately estimated by the analytical relationship [53]:
Voc = ELUMO – ECB (7)
Computational methods
All the calculations were performed with the Gaussian 09 packages [55]. The density functional theory (DFT)
with Becke’s three parameter functional and the Lee–Yang–Parr functional (B3LYP) [56-58] and 6-311G(d,p)
basis set was employed to investigate the structure optimization of the ground state of the dyes in gas phase.
Frequency calculations were performed at the same level of theory as geometry optimization to confirm that all
the optimized geometries are located at the lowest point of the potential energy surface. In a recent work, Tretiak
and Magyar [59] have demonstrated for a series of D-π-A systems that a good description of the charge transfer
states can be achieved when a large fraction of HF exchange is used. A newly designed functional, the long rang
Coulomb attenuating method (CAM-B3LYP) considered long-range interactions by comprising 19% of HF and
81% of B88 exchange at short-range and 65% of HF plus 35% of B88 at long-range [60]. Furthermore, The
CAM-B3LYP has been applied and was reasonably capable of predicting the excitation energies and the
absorption spectra of the D-π-A molecules [61-62]. Therefore, the vertical excitation energy and electronic
absorption spectra were simulated using TD-CAM-B3LYP method in this work. The inclusion of the solvent
effect in theoretical calculations is important when seeking to reproduce or predict the experimental spectra with
a reasonable accuracy. The iodure/iodide couple is used as regenerator in DSSC implying that the solar cells
work in solvent phase. Polarizable continuum model (PCM) [63] has emerged in the last two decades as the most
effective tool to treat bulk solvent effects for both the ground and excited-states. In this paper, the integral
equation formalism polarizable continuum model (IEF-PCM) [64, 65] was chosen in excitation energy
calculations. The cationic and anionic states of dyes were optimized at the B3LYP/6-311G(d,p) level to calculate
the total reorganization energies (λtotal).
2614
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
Q1 Q2
Q3 Q4
Q5 Q6
Q7 Q8
Q9 Q10
Q11 Q12
Table 1. Selected bond lengths (Å) parameters of all dyes in the ground (S0) and excited (S1) states.
Dye d1 d2 d3
S0 S1 S0 S1 S0 S1
Q1 1.0846 1.0823 1.0824 1.0802 1.0834 1.0815
Q2 1.5400 1.5263 1.0700 1.0672 1.0700 1.0681
Q3 1.5400 1.5247 1.0700 1.0687 1.0700 1.0679
Q4 1.4509 1.4726 1.0827 1.0762 1.0833 1.0671
Q5 1.5400 1.5236 1.0699 1.0652 1.0699 1.0661
Q6 1.0844 1.0823 1.0801 1.0768 1.8266 1.8117
Q7 1.0846 1.0817 1.5400 1.5254 1.0834 1.0816
Q8 1.0846 1.0815 1.0700 1.0674 1.4700 1.4534
Q9 1.5399 1.5248 1.0700 1.0683 1.5400 1.5252
Q10 1.5399 1.5254 1.0700 1.0678 1.4700 1.4669
Q11 1.5399 1.5257 1.7600 1.7435 1.7600 1.7435
Q12 1.5399 1.5257 1.5400 1.5260 1.5400 1.5260
Molecular orbitals
To afford deeper insight about the dependence of the electronic properties on the molecular structure, the analysis
of the energy levels of the frontier molecular orbitals (HOMO and LUMO) and the related energy gap (Eg) of
these dyes, is provided in Fig. 3. The electron-donating ability of the electron-donor in D-π-A dyes has the
tendency to influence the electrochemical properties. A D-π-A dye with a stronger electron-donating group
should give a high HOMO as compared to that with a weaker electron-donor. We have investigated the electron-
donor effect on the electronic properties by using different donor groups. According to the analysis of HOMO,
the results of these dyes are in order: Q11>Q8>Q4>Q6>Q12>Q7>Q2>Q9>Q10>Q1>Q3>Q5. The Q11 contains
the strongest electron-donor group (dichloride) since it has the highest HOMO (-8.2179 eV). Dyes Q3 and Q5
with calculated HOMO energy levels -8.8437, -8.8437 eV, respectively, have a weak contribution in electron-
donor ability due to the fact that they contain a benzene ring in the electron-donor moiety. As followed from Fig.
3, the calculated LUMO level for all sensitizers are relatively unaffected by the changes in molecular structure,
due to the inclusion of same electron acceptor group (C=O) in these sensitizers, which is less influenced by the
change of the donor group. The LUMO energy levels of all dyes are much higher than that of TiO2 conduction
band edge (ca. -4.0 eV) [50]. Moreover, molecules in excited states have a strong ability to inject electrons into
TiO2 electrodes. The HOMO of all dyes is lower than that of I-/I-3 (ca. -4.8 eV) [73], therefore, these molecules
that lose electrons could be restored by getting electrons from electrolyte. Thus, electron injection of excited
molecules and, subsequently, regeneration the oxidized species is energetically permitted. This allows the
application of the dyes in DSSC.
HOMO
LUMO
-3,0 -3.069 -3.123 -3.151 -3.232 -3.178 -3.151
-3.314 -3.205 -3.314 -3.314
Energy (ev)
-3,5 -3.559
-3.967
-4,0
-4,5
2.340 2.748 2.830 2.884 3.020 3.102 3.136 3.136 3.156 3.183 3.210 3.238
-5,0
-5,5
-6,0 -5.817
-5.953
-6,5 -6.307 -6.334 -6.307 -6.307 -6.416 -6.389 -6.389
-6.443 -6.443 -6.443
-7,0
Q4 Q11 Q8 Q5 Q7 Q6 Q1 Q3 Q12 Q2 Q9 Q10
Fig. 3. Schematic energy diagram of all dyes.
2617
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
The energy gap for Qi (i = 1–12) was obtained by the differences of HOMO and LUMO energy levels using
B3LYP/6-311G(d,p) and the results are listed in Fig. 3. The order of the Eg is: Q4<Q11<Q8<Q5<Q7<Q6
<Q1<Q3<Q12<Q2<Q9<Q10. With the HOMO–LUMO gap decrease, more photons at the longer-wavelength
side would be absorbed to excite the electrons into the unoccupied molecular orbital, which increases the short-
circuit current density and further enhances the conversion efficiency of the corresponding solar cell. The ranges
of Eg are about 2.34–3.24 eV; we may conclude that these dyes have the potential to be employed in the DSSC
application. It is found that increasing more in Q8, Q11 and Q5 gradually decreased the Eg to around 0.39, 0.44
and 0.35 eV compared to Q9, Q2 and Q10 respectively. For dyes with long conjugation length Q8–Q6, it was
shown clearly that inserting the benzene and chlorine unit in the dyes Q3 and Q6 increase the Eg comparing to
Q8–Q7 which contain heteroatom ring in the electron-donor group.
Optical properties
To gain insight of the optical property and electronic transition, the excitation energy and UV-Vis absorption
spectra for the singlet–singlet transition of all sensitized dyes were simulated using TDDFT with CAM-B3LYP
functional in chloroform solution [74, 75]. Based on the previous works, the chloroform was used as solvent in
the UV-Vis absorption spectra on the quinoxaline based molecules [25, 37, 40]. The computed vertical excited
singlet states, transitions energies and oscillator strength of all sensitized dyes in solvent media are tabulated in
Table 2. The simulated absorption spectra of the studied compounds obtained at the IEF-PCM/TD-CAM-
B3LYP/6-311G(d,p) level is shown in Fig. 4. The spectra show similar profile for all dyes; they present a main
intense band at higher energies from 435 to 515 nm. The most intense contribution to the main band is an
excitation from the HOMO to LUMO orbital in solvent as the first singlet excitation. As showing in Table 2, the
effects of the solvent on the excitation energies of all titled compounds were found stabilized by 0.2 eV as mean
averaged error, while small red-shift (~28 nm) was observed on the maximal wavelength (λmax). Similar effects
were detected on the oscillator strength (~0.18). On the other hand, the position (related to the gap between
HOMO and LUMO levels) and the width of the first band in the spectrum are the two first parameters that can be
related to the dye efficiency, since the absorption shift to lower energies favors the light harvesting process.
Herein, the first vertical excitation energies (Eve) of the dyes are in decreasing order:
Q10>Q9>Q2>Q12>Q3>Q1>Q6>Q7>Q5>Q8>Q11 >Q4 showing that there is a bathochromic shift when passing
from Q10 to Q4. Compared with Q2, the absorption spectra of Q9 and Q10 show slight blue-shift with less
decreased oscillator strength, due probably to the electronegativity of the heteroatom in the electron donor
groups. The absorption spectra of Q3, Q4 and Q5 present the main peak at 503.17, 574.91 and 553.95 nm,
respectively (Table 2), which are clearly shifted to longer wavelengths relative to those of the corresponding
derivatives Q1–Q2 due to the extended π-conjugation. Moreover, the main peak of Q8, Q9, Q10 and Q11 appears
red-shifted in the spectrum compared with the corresponding Q1, Q7 and Q2 by increasing the methyl unit. The
absorption maxima of Q6 and Q12 are found at shorter wavelength than those of Q3–Q5 due to the lower HOMO
level of benzene ring. All results of absorption spectra are in good agreement with the energy levels and band gap
discussed above.
2,2
oscillator strenght
Q7 Q5 Q Q
2,0 Q8 11 4 Q1
Q12Q3Q1 Q2
1,8
Q6
Q9Q10Q2 Q3
1,6 Q4
1,4 Q5
Q6
1,2 Q7
1,0 Q8
Q9
0,8 Q10
0,6 Q11
Q12
0,4
0,2
0,0
-0,2
280 320 360 400 440 480 520 560 600 640 680 720 760
wavelength (nm)
Fig. 4. Simulated absorption spectra of all dyes.
2618
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
In order to study the emission photoluminescence properties of the studied compounds Qi (i=1–12), the adiabatic
emission spectra were obtained using the optimized geometry of the first excited singlet state at the
TDDFT/CAM-B3LYP/6-311G(d,p) level. The simulated fluorescence wavelengths with the strongest oscillator
are presented in Table 2. According to the absorption and emission data, the values of Stokes shift (SS) for all
dyes were obtained. The emission spectra arising from S1 state is assigned to π*→π and LUMO→HOMO
transition character for all molecules. Through analyzing the transition configuration of the fluorescence, we
found that the calculated fluorescence is just the reverse process of the lowest lying absorption. Moreover, the
observed redshifted emission of the photoluminescence (PL) spectra in order: Q9<Q10<Q2<Q12<Q3<Q1
<Q6<Q7<Q5<Q8<Q11<Q4 when passing from Q9 to Q4 is in reasonable agreement with the obtained results of
absorption. Furthermore, the Stokes shift of these dyes is found to be in the range 58.8 and 105.5 nm. The Q4
emitted at higher wavelength (574.91 nm) with strongest intensity (f = 1.985), and was also characterized by
larger Stokes shift (105.5 nm, Table 2). These encouraging optical properties suggest that Q4 with styryl
electron-donor will be the best candidate in the DSSC devices.
Table 2. Emission spectra data for all dyes obtained with PCM-CAM-B3LYP/6-311G(d.p).
Dye E00 Eve EHOMO ELUMO ΔE λmax f SS ET µ
(eV) (eV) (eV) (eV) (eV) (nm) (nm) (eV) (Debye)
Q1 3.19 2.70 -6.443 -3.314 3.136 505.36 1.752 84.3 -13346.63 3.8922
Q2 3.23 2.84 -6.416 -3.232 3.183 480.73 1.601 64.1 -14410.89 3.1594
Q3 3.20 2.71 -6.443 -3.314 3.136 503.17 1.755 83.2 -20663.49 3.4040
Q4 2.89 2.40 -6.307 -3.967 2.340 574.91 1.985 105.5 -21694.51 2.4341
Q5 3.00 2.48 -6.443 -3.559 2.884 553.95 1.960 103.0 -34140.03 6.4885
Q6 3.17 2.66 -6.307 -3.205 3.102 515.54 1.716 90.3 -25792.91 1.4646
Q7 3.11 2.57 -6.334 -3.314 3.020 533.07 1.964 100.1 -14410.70 4.2558
Q8 2.91 2.44 -5.953 -3.123 2.830 563.40 1.839 96.9 -14844.70 7.0044
Q9 3.24 2.85 -6.389 -3.178 3.210 473.92 1.541 58.8 -15474.94 3.7563
Q10 3.32 2.88 -6.389 -3.151 3.238 479.96 1.587 76.3 -15908.43 3.6954
Q11 2.90 2.41 -5.817 -3.069 2.748 572.23 1.968 104.9 -39303.24 1.1978
Q12 3.21 2.72 -6.307 -3.151 3.156 500.15 1.763 81.6 -16484.54 4.1417
To confirm this result, we determined the dipole moment vectors in a three-dimensional representation of the
quinoxalin-2(1H)-one and its eleven derivatives (fig. 5). Dipole moment μ vectors computed at the RHF/STO-3G
level displayed in figure 4 lend further support to this image; µ Vectors are referred to the standard orientation of
every molecule, i.e. the nuclear charge center for the molecule is at the origin of coordinates.
Y
2 Q6
Q10 Q3
Q5
Q2
Q4
Q7
0
-4 -2 0 2 4 6 8
X
Q9 Q12
Q11
Q1
-2 Q8
Fig. 5. Dipole moments (Debye) in a three-dimensional representation, computed at the RHF/STO-3G level of
theory for quinoxalin-2(1H)-one derivatives in Fig. 1.
2619
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
The dipole moment for Q1 (3.892 D) is increased when more pchlorostyryl and nitro are added away from
position 3 and 7 in Q5 (6.488 D) and Q8 (7.004 D) with an orientation (+54° and +13°) and decreased when they
are replaced in Q7, Q2 and Q3 with an week orientation, and even effect at summer noted in the compound Q11
and Q12 with a small orientation (~12°). However, the variations of μ in these seven cases is considerably
smaller than that found in Q6, Q9, Q10 and Q4 where the presence of electronegative chlorine in the opposite
side with respect to carbonyl not only reduces dramatically the dipole moment but also changes the vector
orientation with large angle (-140° in Q4) (Fig. 5). The information provided by NPA charges as well as this
change with regard to μ are in agreement with the known weak resonance effect of chlorine, already known [76-
78], as compared with its inductive effect.
Photovoltaic properties
The electronic injection free energy ΔGinject, ground Edye and excited Edye* state oxidation potentials computed for
the dyes Qi are represented in Table 3. Based on Koopman’s theorem, ground state oxidation potential energy is
related to ionization potential energy.
Edye can be estimated as negative EHOMO [79]. Edye* is calculated based on Eq. (5). Edye* of all dyes is increasing
order: Q4<Q11<Q12<Q3=Q8<Q1<Q9<Q2=Q6<Q5<Q7<Q10. It shows that the most convenient oxidizing
species is Q4 while Q10 is the worst. All ΔGinject estimated from Eq. (4) is in negative value for all sensitizers,
thus the electron injection from the dye to TiO2 is spontaneous.
As seen from Table 3 and Fig. 6(a) the calculated ΔGinject are decreased in the order: Q4>Q11
>Q12>Q3=Q8>Q1>Q9>Q2=Q6>Q5>Q7>Q10. It shows that Q4 has the largest ΔGinject value while Q10 has the
smallest. Another factor related to efficiency of DSSC is the performance of the dyes responsible of the incident
light. Based on the LHE of the dyes, the value has to be as high as possible to maximize the photocurrent
response. The LHE values for all dyes are in narrow range 0.971–0.989 (Table 3), but increase slightly with
increasing the conjugation length (Fig. 6(b)). This means that all the sensitizers give similar photocurrent.
As we know, besides the reaction free energy, the reorganization energy λtotal could also affect the kinetics of
electron injection. So, the calculated λtotal is also important to analyze the relationship between the electronic
structure and the Jsc. As mentioned in section ‘Theoretical background’, the small λtotal which contains the hole
and electron reorganization energy could enhance the Jsc. As seen from Table 3 and Fig. 6(c) the calculated λtotal
of all dyes are increased in the order: Q4<Q11<Q8<Q5< Q6<Q7<Q9<Q2<Q1<Q3<Q12<Q10. It shows that dye
Q4 possesses the smallest total reorganization energy while dye Q10 has the largest. As a result, dye Q4 exhibits
a favorable Jsc due to the relative similar LHE, larger ΔGinject and smaller λtotal. At the same time, ΔGinject and λtotal
are more important to govern the Jsc mostly. We know that besides the short-circuit current density Jsc the overall
power conversion efficiency η also could be influenced by the open-circuit voltage (Voc). Therefore, between two
dyes with similar structures, the electron injection would be more efficient for that dye with the higher excited
state related to the semi-conductor conduction band edge (i.e. higher Voc). Based on Eq. (7), it was found that Voc
of all dyes (Table 3) is in the range 1.469–2.367 eV and in decreasing order:
Q4>Q5>Q1=Q3=Q7>Q2>Q6>Q9>Q10=Q12>Q8>Q11. It shows that Q4 and Q5 have the larger Voc values than
other dyes, while Q8 and Q11 have the smallest. As a consequence of the above data, we could draw a
conclusion that the large LHE, ΔGinject, Voc as well as small λtotal could have a high efficiency. Thus, the
2620
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
performance of DSSC sensitized by dye Q4 might be superior to the other dyes, due to its favorable
performances of the above factors based on our computed results.
1,10 (b)
(a)
0,990 1,05
0,988
1,00
0,986
0,984 0,95
(ev)
0,982 0,90
inject
0,980
LHE
0,85
-G
0,978
0,976 0,80
0,974
0,75
0,972
0,70
0,970
(c)
0,64 (d)
1,0 d e m o d e m o d e m o d e m o d e m o
0,62
d e m o d e m o d e m o d e m o d e m o
0,60 0,8
d e m o d e m o d e m o d e m o d e m o
0,58
total (ev)
0,6
Voc (eV)
0,56 d e m o d e m o d e m o d e m o d e m o
0,54 0,4
d e m o d e m o d e m o d e m o d e m o
0,52
0,2 d e m o d e m o d e m o d e m o d e m o
0,50
0,48 0,0 d e m o d e m o d e m o d e m o d e m o
Fig. 6. Critical parameters influenced Jsc along of investigated sensitizers: (a) the light-harvesting efficiency, (b)
the electronic injection free energy, (c) the reorganization energy λtotal and (d) the open-circuit voltage Voc.
Conclusion
In this study, the electronic structure and optical absorption properties of twelve selected quinoxalin-2(1H)-one
dye sensitizers were investigated by using DFT and TDDFT. Based upon on the calculated results, we have
analyzed the role of different electron-donor groups in the tuning the geometries, electronic structures and optical
properties. Also, we have aimed to see the sensitizer donor effects on the Voc and Jsc of the cell through discussing
the key factors affecting Voc and Jsc with the goal of finding potential sensitizers for use in DSSC. It can be
concluded that this class of selected quinoxalin-2(1H)-one dyes shows a good photophysical properties related to
DSSC use but in different outstanding properties. The Q4 dye possessing styryl donor group was found to be the
best photosensitizer for use in DSSC, with respect to the other dyes, as the calculation results show its good
oxidation potential energy and electron injection force that lie over the ECB of TiO2 and under reduction potential
energy of the electrolyte. This theoretical approach presents a guiding tool to the synthesis process and helps to
understand the structure–properties relationship of these new systems.
Acknowledgments-This work has been supported by the Hassan II Academy of Science and Technology. The authors are grateful to
Professor Mohamed Addou (Dean of the Faculty of Science and Technology in Tangier-Morocco) for the explication of solar cell
applications.
References
1. Grätzel M., Nature., 414 (2001) 338-344.
2. Hagfeldt A., Grätzel M., Acc. Chem. Res., 33 (2000) 269-277.
3. O’Regan B., Grätzel M., Nature., 353 (1991) 737-740.
4. Kamat P.V., Haria M., Hotchandani S., J. Phys. Chem. B, 108 (2004) 5166-5170.
2621
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
5. Bisquert J., Cahen D., Hodes G., Ruehle S., Zaban A., J. Phys. Chem. B, 108 (2004) 8106-8118.
6. Furube A., Katoh R., Yoshihara T., Hara K., Murata S., Arakawa H., Tachiya M., J. Phys. Chem. B, 108
(2004) 12583-12592.
7. Tian Z., Huang M., Zhao B., Huang H., Feng X., Nie Y., Shen P., Tan S., Dyes Pigm, 87 (2010) 181-187.
8. Matsui M., Ito A., Kotani M., Kubota Y., Funabiki K., Jin J., Yoshida T., Minoura H., Miura H., Dyes Pigm,
80 (2009) 233-238.
9. Ma X., Wu W., Zhang Q., Guo F., Meng F., Hua J., Dyes Pigm. 82 (2009) 353-359.
10. Nazeeruddin M.K., Kay A., Rodicio I., Humphry-Baker R., Mueller E., Liska P., Vlachopoulos N., Gratzel
M., J. Am. Chem. Soc. 115 (1993) 6382-6390.
11. Nazeeruddin M.K., Zakeeruddin S.M., Humphry-Baker R., Jirousek M., Liska P., Vlachopoulos N., Shklover
V., Fischer C.H., Gratzel M., Inorg. Chem., 38 (1999) 6298-6305.
12. Nogueira A.F., Longo C., De Paoli M.A., Coord. Chem. Rev. 248 (2004) 1455-1468.
13. Wang S.Z., Cui Y., Hara K., Dan-Oh Y., Kasada C., Shinpo A., Adv. Mater. 19 (2007) 1138-1141.
14. Wong B.M., Codaro J.G., J. Chem. Phys. 129 (2008) 214703-214711.
15. Horiuchi T., Miura H., Sumioka K., Uchid S., J. Am. Chem. Soc. 126 (2004) 12218-12219.
16. Chunyang J., Zhongquan W., Jiaqiang Z., Zi L., Xiaojun Y., Yu S., Spectrochim. Acta, Part A, 86 (2012) 387.
17. Zeng W.D., Cao Y.M., Bai Y., Wang Y.H., Shi Y.S., Wang P., Chem. Mater., 22 (2010) 1915-1925.
18. Li G., Zhou Y., Cao X., Bao P., Jiang K., Lin Y., Chem. Commun., 16 (2009) 2201-2203.
19. Ferrere S., Zaban A., J. Phys. Chem. B, 101 (1997) 4490-4493.
20. Ferrere S., Gregg B., New J. Chem., 26 (2002) 1155-1160.
21. Bourass M., Touimi Benjelloun A., Hamidi M., Benzakour M., Mcharfi M., Sfaira M., Serein-Spirauc F.,
Lère-Portec J.P., Sotiropoulos J.M., Bouachrine M., J. Saudi. Chem. Soc., DOI:10.1016/j.jscs.2013.01.003.
22. Casanova D., Chem. Phys. Chem., 12 (2011) 2979-2988.
23. Xu J., Wang L., Liang G., Bai Z., Wang L., Xu W., Shen X., Spectrochim. Acta, Part A, 78 (2011) 287-293.
24. Lin Y., Fan H., Li Y., Zhan X., Adv. Mater., 24 (2012) 3087-3106.
25. Dong Wook Chang, Seo-Jin Ko, Jin Young Kim, Liming Dai, Jong-Beom Baek, Synth. Met., 162 (2012)
1169.
26. El Assyry A., Benali B., Lakhrissi B., El Faydy M., Ebn Touhami M., Touir R., Touil M., Res. Chem.
Intermed., 41 (2015) 3419-3431.
27. Ben Hmamou D., Salghi R., Zarrouk A., Zarrok H., Hammouti B., Al-Deyab S. S., El Assyry A., Benchat N.,
Bouachrine M., Int. J. Electrochem. Sci., 8 (2013) 11526-11545.
28. Benhiba F., ELaoufir Y., Belayachi M., Zarrok H., El Assyry A., Zarrouk A., Hammouti B., Ebenso E. E.,
Guenbour A., Al Deyab S. S. and Oudda H., Der Pharm. Lett., 6 (2014) 306-318.
29. Larouj M., ELaoufir Y., Serrar H., El Assyry A., Galai M., Zarrouk A., Hammouti B., Guenbour A., El
Midaoui A., Boukhriss S., Ebn Touhami M. and Oudda H., Der Pharm. Lett., 6 (2014) 324-334.
30. El Assyry A., Benali B., Boucetta A., Lakhrissi B., J. Struct. Chem., 55 (2014) 38-44.
31. Tayebi H., Bourazmi H., Himmi B., El Assyry A., Ramli Y., Zarrouk A., Geunbour A., Hammouti B., Der
Pharm. Chem., 6 (2014) 220-234.
32. Benali B., Lazar Z., Boucetta A., El Assyry A., Lakhrissi B., Massoui M., Jermoumi C., P. Negrier, J. M.
Leger, Mondieig D., Spectrosc. Lett., 41 (2008) 64-71.
33. El Assyry A., Benali B., Jdaa R., Lakhrissi B., Touil M., Zarrouk A., J. Mater. Environ. Sci., 5 (2014) 1434.
34. Qiang Peng, Xiangju Liu, Yuancheng Qin, Jun Xu, Mingjun Liaand, Liming Dai, J. Mater. Chem., 21 (2011)
7714-7722.
35. Hsieh-Chih Chen, Ying-Hsiao Chen, Chung-Hao Liu, Yen-Hao Hsu, Yun-Chen Chien, Wei-Ti Chuang,
Chih-Yang Cheng, Chien-Liang Liu, Shang-Wei Chou, Shih-Huang Tung, Pi-Tai Chou, Polym.Chem., 4
(2013) 3411-3418.
36. Yoonkyoo Lee, Thomas P. Russell, Won Ho Jo, Org. Electron., 11 (2010) 846-853.
37. Fu Y., Cha H., Song S., Lee G.-Y., Park C. E., Park T., J. Polym. Sci., Part A: Polym. Chem., 51 (2013) 372.
38. Velusamy M., Huang J.-H., Hsu Y.-C., Chou H.-H., Ho K.-C., Wu P.-L., Chang W.-H., Lin J. T., Chu C.-W.,
Org. Lett., 11 (2009) 4898-4901.
39. Dhananjaya Kekuda, Jen-Shien Huang, Marappan Velusamy, Jiann T. Lin, Chih Wei Chu, Sol. Energy
Mater. Sol. Cells, 94 (2010) 1767-1771.
40. Lu X., Feng Q., Lan T., Zhou G., and Wang Z. S., Chem. Mater., 24 (2012) 3179-3187.
41. Chang D. W., Lee H. J., Kim J. H., Park S. Y., Park S.-M., Dai L., Baek J.-B., Org. Lett., 13 (2011) 3880.
2622
J. Mater. Environ. Sci. 6 (9) (2015) 2612-2623 El Assyry et al.
ISSN : 2028-2508
CODEN: JMESCN
42. Dessì A., Consiglio G.B., Calamante M., Reginato G., Mordini A., Peruzzini M., Taddei M., Sinicropi A.,
Parisi M. L., Biani F. F., Basosi R., Mori R., Spatola M., Bruzzi M., Zani L., Eur. J. Org. Chem. 2013 (2013)
1916-1928.
43. Belghiti N., Bennani M., Hamidi M., Bouzzine S.M., Bouachrine M., Afr. J. Pure. Appl. Chem. 6 (2012) 164.
44. Zhang C.R., Liu Z.J., Chen Y.H., Chen H. S., Wu Y.Z., Yuan L.H., J. Mol. Struct. (Theochem), 899 (2009) 86.
45. Duncan W.R., Prezhdo O.V., J. Am. Chem. Soc., 130 (2008) 9756-9762.
46. Guo Z., Liang W.Z., Zhao Y., Chen G.H., J. Phys. Chem. C, 112 (2008) 16655-16662.
47. Ramli Y., Benzeid H., Bouhfid R., Kandri Rodi Y., Ferfra S., Essassi E.M., St. Cerc. St. CICBIA, 11 (2010) 67.
48. Narayan M.R., Renew. Sust. Energy. Rev. 16 (2012) 208-215.
49. Zhang Z.L., Zou L.Y., Ren A.M., Liu Y.F., Feng J.K., Sun C.C., Dyes Pigm. 96 (2013) 349-363.
50. Asbury J.B., Wang Y.Q., Hao E., Ghosh H., Lian T., Res. Chem. Intermed. 27 (2001) 393-406.
51. Zhang J., Li H.B., Sun S.L., Geng Y., Wu Y., Su Z.M., J. Mater. Chem. 22 (2012) 568-576.
52. Ding W.L., Wang D.M., Geng Z.Y., Zhao X.L., Xu W.B., Dyes Pigm. 98 (2013) 125-135.
53. Sang-aroon W., Saekow S., Amornkitbamrung V., J. Photochem. Photobiol. A, 236 (2012) 35-40.
54. Balanay M.P., Kim D.H., J. Mol. Struct. (Theochem), 910 (2009) 20-26.
55. Frisch M.J., Trucks G.W., Schlegel H.B., Scuseria G.E., Robb M.A., Cheeseman J.R., et al., Gaussian 09,
revision A.02, Gaussian, Inc., Pittsburgh, PA, 2009.
56. Becke A.D., J. Chem. Phys., 98 (1993) 1372-1377.
57. El Assyry A., Benali B., Boucetta A., Lakhrissi B., J. Mater. Environ. Sci., 5 (2014) 1860-1867.
58. Lee C., Yang W., Parr R.G., Reviews B, 37 (1988) 785-789.
59. Magyar R.J., Tretiak S., J. Chem. Theory. Comput. 3 (2007) 976-987.
60. Xue Y., An L., Zheng Y., Zhang L., Gong X., Qian Y., Liu Y., Comput. Theor. Chem., 981 (2012) 90-99.
61. Camino B., De-La-Pierre M., Ferrari A.M., J. Mol. Struct. 1046 (2013) 116-123.
62. Irfan A., Jin R., Al-Sehemi A.G., Asiri A.M., Spectrochim. Acta, Part A, 110 (2013) 60-66.
63. Tomasi J., Mennucci B., Cammi R., Chem. Rev., 105 (2005) 2999-3093.
64. Cossi M., Barone V., J. Chem. Phys., 115 (2001) 4708-4717.
65. Adamo C., Barone V., Chem. Phys. Lett., 330 (2000) 152-160.
66. Zhang X.H., Wang L.Y., Zhai G.H., Wen Z.Y., Zhang Z.X., J. Mol. Struct., 881 (2008) 117-122.
67. Mondieig D., Negrier Ph., Leger J.M., Lakhrissi L., El Assyry A., Lakhrissi B., Essassi E.M., Benali B.,
Boucetta A., Russ. J. Phys. Chem. A, 89 (2015) 807-811.
68. Négrier Ph., Mondieig D., Léger J. M., Benali B., Lazar Z., Boucetta A., Elassyry A., Lakhrissi B., Jermoumi
C., Massoui M., Anal. Sci., 22 (2006) 175-176.
69. Benali B., Lazar Z., Elblidi K., Lakhrissi B., Massoui M., Elassyry A., Cazeau-Dubroca C., J. Mol. Liq., 128
(2006) 42-45.
70. El assyry A., Benali B., Boucetta A., Lazar Z., Lakhrissi B., Massoui M., Mondieig D., Spectrosc. Lett., 42
(2009) 203-209.
71. Benali B., El Assyry A., Boucetta A., Lazar Z., Lakhrissi B., Massoui M., Mondieig D., Spectrosc. Lett., 40
(2007) 893-901.
72. Elassyry A., Benali B., Lazar Z., Elblidi K., Lakhrissi B., Massoui M., Mondieig D., J. Mol. Liq., 128 (2006)
46.
73. Hagfeldt A., Grätzel M., Chem. Rev., 95 (1995) 49-68.
74. Jacquemin D., Perpete E.A., CiofiniI I., Adamo C., Acc. Chem. Res., 42 (2009) 326-334.
75. Adamo C., Jacquemin D., Chem. Soc. Rev., 42 (2013) 845-856.
76. El Assyry A., Benali B., Boucetta A., Mondieig D., Res. Chem. Intermed., 40 (2014) 1043-1052.
77. Lazar Z., Benali B., Elblidi K., Zenkouar M., Lakhrissi B., Massoui M., Kabouchi B., Cazeau-Dubroca C., J.
Mol. Liq., 106 (2003) 89-95.
78. El Assyry A., Benali B., Res. Chem. Intermed., 40 (2014) 627-636.
79. Pearson R.G., Inorg. Chem., 27 (1988) 734-740.
(2015) ; http://www.jmaterenvironsci.com
2623