You are on page 1of 7

Journal of Luminescence 205 (2019) 7–13

Contents lists available at ScienceDirect

Journal of Luminescence
journal homepage: www.elsevier.com/locate/jlumin

The effects of dye aggregation on the performance of organic dyes in dye- T


sensitized solar cells: From static model to molecular dynamics simulation
⁎ ⁎
Weiyi Zhang, Yue Zhang, Huishuang Su, Xinrui Zhu, Li Wang , Jinglai Zhang
College of Chemistry and Chemical Engineering, Henan University, Kaifeng, Henan 475004, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: The properties of three D1-π-D2-A dyes with dithiafulvenyl as donor (D1), phenothiazine as ancillary donor
Dye-sensitized solar cells (D2), cyanoacrylic acid as acceptor along with different π groups, diphenyl for 1, dithienyl for 2, and thio-
Dye-TiO2 adsorbed system pheneyl-benzothiadiazole for 3, are studied. Besides the isolated dyes, the monomeric and dimeric adsorptions
Aggregation are also studied by combination of first principle and molecular dynamic simulations. There is no distinct dif-
Electronic coupling
ference among three dyes on the basis of only isolated dyes. When the dye-TiO2 adsorption is considered, 3 is
Molecular dynamics
excluded because of its inferior items. However, it is still a difficult task to differentiate 1 and 2 even if the
aggregation effect is included. The fluctuation of 1 is larger than that of 2 during the dynamic simulation
resulting in the larger electronic coupling. The aggregation, especially for the aggregation during the dynamic
simulation, is necessary to evaluate the performance of organic dye. Additionally, a new D-π-A dye 4 with the
identical donor, π-bridge and acceptor is designed to compare with 2. 2 is totally superior to 4 indicating that
suitable groups and appropriate configuration are both important to enhance the performance of organic dye.

1. Introduction along with cyanoacrylic acid as acceptor linked through different π-


bridges [5]. After that, the design of D-π-A sensitizers for DSSCs is ac-
As the important means for conversion solar radiation to electricity celerated by employment of DTF as donor.
directly, dye-sensitized solar cells (DSSCs) have attracted extensive Phenothiazine (PTZ) derivatives are also promising donors in or-
interests due to easy fabrication, high efficiency, and low costs [1]. ganic sensitizers in DSSCs owing to their excellent electron donating
Further improvement of the efficiency is the prerequisite to accomplish abilities. Their butterfly configurations would impede molecular ag-
the large scale applications. Although optimization of each component gregation resulting in better electron separation. Lots of D-π-A organic
would improve the overall performance, the sensitizers play the most dyes have been developed with PTZ derivatives as donor [6].
essential role in increasing the power conversion efficiency [2]. Since both DTF and PTZ are potential candidates for high-efficiency
Metal-free organic dyes have been the powerful alternatives for dyes, Jia et al. have constructed new organic dyes, WD15 (See Fig. S1),
polypyridyl Ru(II) complex and zinc porphyrin dyes because of their by introduction of DTF and PTZ units as building blocks simultaneously
cheap raw materials, facile synthesis, and easily tunable structures [3]. achieving the best efficiency of 6.63% [6]. Besides the chemical prop-
D-π-A (donor-π-acceptor) is the most common configuration for organic erties of components, the configuration is also the other critical item to
dyes, which is favorable for intramolecular charge transfer leading to a affect the performance. On the basis of WD15, different π groups are
high efficiency. inserted between DTF and PTZ to develop new organic dyes with D1-π-
As a smaller version of the fulvene family, dithiafulvenyl (DTF) is D2-A configuration [7].
considered as a good choice to be terminal electron-donating group in A great variety of components as well as configurations results in
D-π-A system for DSSC applications [4]. As a strong electron donor, it substantial organic dyes. To develop organic dyes with desired prop-
facilitates the electron injection from excited state dye molecules to the erties, it is necessary to build an efficient structure-property relation-
conduction band of semiconductor. Its non-planar molecular config- ship. The performance of DTF-C sensitizers (See Fig. S1) is quantita-
uration is also helpful to suppress the aggregation and the recombina- tively evaluated [5]. Moreover, theoretical studies have been included
tion of injected electrons with the redox couple [4]. In 2012, DTF is in most of experimental investigations to uncover the electron dis-
firstly taken as donor to synthesize organic dyes using D-π-A approach tribution and variation of energy level of frontier molecular orbital


Corresponding authors.
E-mail addresses: chemwangl@henu.edu.cn (L. Wang), zhangjinglai@henu.edu.cn (J. Zhang).

https://doi.org/10.1016/j.jlumin.2018.08.084
Received 15 May 2018; Received in revised form 3 August 2018; Accepted 30 August 2018
Available online 31 August 2018
0022-2313/ © 2018 Elsevier B.V. All rights reserved.
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

absorption shapes and the absorption bands simulated by CAM-B3LYP


method are more accurate and reasonable. Therefore, the TD-CAM-
B3LYP functional was employed to study the absorption spectra for
other two dyes. All aforementioned calculations were carried out by the
Gaussian 09 software [18].
All periodic DFT calculations were carried out by the plane-wave
code VASP (Vienna ab initio simulation package) [19–21] with the
generalized gradient approximation (GGA) using Perdew-Burke-Ern-
zerhof (PBE) exchange-correlation function, supplemented with Grim-
me's D2 dispersion correction [22,23]. The energy cutoff was set to be
400 eV and the optimization would stop when the force on each atom
was smaller than 0.1 eV/Å. Both monomeric adsorption form and π-π
dimeric adsorption form were considered. To gain insight into the in-
terfacial properties between dye and TiO2, the 4 × 6 × 6 and 6 × 8 × 6
TiO2 anatase (101) supercells for monomer-adsorption and π-π dimeric-
adsorption, respectively, were prepared as the adsorption surface. A
vacuum buffer space of 21, 23, and 44 Å was added in x-, y-, and z-
direction for monomer-adsorption, and 31, 31, and 40 Å was added in
x-, y-, and z-direction for π-π dimeric-adsorption, respectively. Fur-
thermore, the densities of states (DOS) and projected density of states
(PDOS) were calculated at the same level to deeply analyze the ad-
sorption properties.
For the dyes adsorbed on the TiO2 surface, we carried out molecular
dynamics simulations for 10 ps with an integration time step of 1 fs at
T = 298 K in the constant volume constant internal energy (NVE en-
semble). The dynamic simulation was implemented in the DFTB+ 1.3
program package [24]. The SK-parameters of mio-1–1 [25–27] for C, N,
O, H, and S and tiorg-0–1 [28] for Ti were employed.

3. Results and discussion

3.1. Isolated dyes

Three investigated dyes have the similar structural parameters since


the only difference for them is the π-bridge, which is tabulated in Table
S1. The corresponding energy levels of FMOs are shown in Fig. 1. In the
one hand, the lowest unoccupied molecular orbital (LUMO) energy of
the dye is higher than the energy level of TiO2 conduction band
Scheme 1. The sketch map structures of all dyes. (-4.0 eV) [29], which would provide enough force for electron injec-
tion. In the other hand, the highest occupied molecular orbital (HOMO)
(FMO). However, the influence of aggregation is not included to con- energy of the dye is more stable than the Nernst potential of iodine/
sider the performance of organic dyes, which is an important item triiodide redox electrolytes (-4.8 eV) [30], which ensures the oxidized
especially for PTZ and DTF units. sensitizers to accept the electrons from the electrolytes to accomplish
On the basis of 1, other two organic dyes, 2, and 3, are theoretically the whole cell circuit. When the diphenyl group is replaced by dithienyl
designed with dithienyl and thiopheneyl-benzothiadiazole as π groups or thiopheneyl-benzothiadiazole, the LUMO energy level is stabilized
(See Scheme 1) to explore the influence of π group. Besides the isolated and the HOMO energy level is inverse. As a result, the HOMO-LUMO
dyes and dyes adsorbed on TiO2 surface, the influence of dye ag- energy gap is narrowed, which is helpful to enhance the light har-
gregation on the performance is also investigated. Furthermore, the vesting.
dynamic property is also calculated with the goal to build more precise
structure-property relationship.

2. Computational details

All ground state geometries of the isolated dyes were optimized by


Becke's three-parameter nonlocal-exchange functional with the non-
local correlation of Lee-Yang-Parr method (B3LYP) [8,9] with the
6–31 G(d,p) basis sets [10]. The positive vibrational frequencies testi-
fied that the optimized structures were local minima on their respective
potential energy surfaces. Based on above results, the time-dependent
density functional theory (TD-DFT) was utilized to simulate the ab-
sorption spectra. A series of different functionals, B3LYP, CAM-B3LYP
[11], LC-BLYP [12,13], M06-2 ×[14], and PBE0 [15] were performed
for dye 1 with the polarized continuum model (PCM) [16,17] in tet-
rahydrofuran (THF) medium to determine the more suitable functional Fig. 1. Energy diagram of HOMO and LUMO for all dyes calculated at the CAM-
(See Fig. S2). As compared with the experimental result, both the B3LYP/6–31 G(d,p) level of theory, TiO2, and the electrolyte.

8
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

build a reliable structure-property relationship, the dye-TiO2 adsorbed


system should be studied rather than only the isolated dyes.

3.3. Adsorbed dyes onto the TiO2

When the dye is adsorbed on the TiO2 surface, the energy gap would
be varied. Correspondingly, some related properties would also be
changed. It is necessary to study the dye-TiO2 adsorbed system, which
is more close to the real environment than the isolated dye. The bi-
dentate bridging adsorption model via the carboxyl group is employed
along with one proton transferred to nearby surface oxygen. The TiO2
(101) surface is employed since it is the most stable and popular surface
[38]. The optimized adsorption structures simulated by the first prin-
ciple are plotted in Fig. 3. The adsorption has negligible effect on the
structures of dye. The adsorption energy (Eads ) of 3 is more positive
than that of 1 and 2 (See Table 2) indicating the weaker binding, which
Fig. 2. The UV/Vis absorption spectra of all dyes calculated in tetrahydrofuran
is unfavorable for the electron injection. The Eads is defined by the
solvent at the CAM-B3LYP/6–31 G(d,p) level of theory.
following relationship: as Eads = Edye + TiO2 − (Edye + ETiO2) , where Edye ,
ETiO2 , and Edye + TiO2 refer to the total energies of the isolated dye, TiO2,
On the basis of the optimized geometries, the simulated absorption and dye-TiO2 complex, respectively.
spectra are presented in Fig. 2. The maximum absorption wavelength After adsorption, the energy gap is greatly varied as compared with
(λmax) of 1 (389 nm) is in good agreement with the experimental result the isolated dyes. On the basis of the optimized dye-TiO2 adsorbed
(382 nm). The λmax(s) of 2 and 3 are both red-shifted as compared with system, DOS and PDOS for the pure and adsorbed dyes are shown in
that of 1 suggesting the wider light harvesting region. Although the Fig. 4 along with those of TiO2 surface. After adsorption, the LUMO of
λmax of 3 is larger than that of 2, the absorption strength of former is adsorbed dye moves up to more positive values. In contrast, the CB of
smaller. According to the following equation [31]: TiO2 shifts towards the inverse direction. As a result, the energy gap
LHE = 1 − 10 −f (1) between the LUMO of adsorbed dye and the CB of TiO2 is enlarged as
compared with that of isolated dye. However, 3-TiO2 adsorption system
the light harvesting efficiency (LHE) is proportional to the oscillator has the smallest energy gap between the LUMO of adsorbed dye and the
strength of the dye at the λmax. The weaker absorption strength in- CB of TiO2 with the value of 0.99 eV, which is not beneficial for the
dicates the less power conversion efficiency (PCE). electron injection. To further confirm the inferior performance of
electron from the excited dye to the CB of TiO2, the electron injection
3.2. Overall efficiency time (in fs) is calculated [39–41]. 3 has the longer electron injection
time of 1.66 fs than other two dyes with the values of 1.23 fs for 1 and
The overall conversion efficiency (η) is related with Jsc, Voc, and FF: 1.21 fs for 2 indicating the worst electron injection efficiency.
For isolated dyes, 3 still has some advantageous to improve the PCE
Voc Jsc including the maximum absorption wavelength and the smallest λtotal.
η = FF
Is (2) When the dye-TiO2 adsorption system is considered, 3 is firstly ex-
in which Jsc is short-circuit current density, Voc is open-circuit voltage, cluded due to all inferior items. There is an interesting thing that the
and FF is fill factor. Normally, η is determined by both Jsc and Voc, since performance of 1-TiO2 is better or comparable as compared to that of 2-
FF is an incidental factor. It is impossible to calculate the accurate va- TiO2 for all employed factors. However, the difference between them is
lues of Jsc and Voc, therefore, some related items are developed to too small to differentiate them. Therefore, other factors should be fur-
qualitatively determine the PCE including ΔGinject, λtotal, and LHE. Their ther considered.
detailed definitions please refer to references 31–36. The corresponding
values are listed in Table 1. The 2 has the most negative ΔGinject in- 3.4. Comparison with D-π-A configuration dye
dicating the more sufficient force to promote electron injection
from excited dyes to the conduction band (CB) of TiO2. At the Although 1 and 2 have the different π groups, they have the same
same time, it also has the larger λtotal suggesting the smaller charge performance as far. If the same π group is employed in different con-
transfer rate constant (k) on the basis of the equation of figurations, the performance would be kept or varied? A new D-π-A dye
k = V 2 (π /ℏ2kB Tλtotal )1/2 exp(−λtotal /4kB T ) [37]. The contradictory re- 4 is designed with the same donor, π group, and acceptor to construct 2.
sults would also be suitable for other two dyes 1 and 3. Moreover, 3 has The corresponding results are also shown in Table 1 and Table 2. For
the smallest Voc, which is also unfavorable for the improvement of PCE. the isolated 4, it has the more positive ΔGinject, the larger λtotal, and the
In general, it is difficult to make a clear sequence for the PCE of three smaller LHE than 2 indicating the less PCE. Even if the adsorbed system
dyes since no one has all superiority. Even if a relative sequence is is considered, 4 still has the more positive adsorbed energy, the smaller
made, the difference between various items is too small to distinguish energy gap between the LUMO of dye and the CB of TiO2, and the
them. Some of them are even in an acceptable computational error. To longer electron injection time than 2. The 4-TiO2 system has the most

Table 1
Estimated electrochemical parameters for the dyes.
Dye dye
EOX (eV) dye *
EOX (eV) ΔGinject(eV) LHE Voc(eV) λh(eV) λe(eV) λtotal(eV)

1 6.34 3.15 − 0.85 0.86 2.72 0.32 0.31 0.63


2 5.99 3.07 − 0.93 0.99 2.69 0.39 0.30 0.68
3 6.07 3.54 − 0.46 0.92 2.37 0.33 0.24 0.57
4 6.31 3.79 − 0.21 0.97 2.23 0.33 0.42 0.75

9
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

Table 2
The calculated adsorption energy (Eads), H-L energy gaps before (Eg1) and after
adsorption (Eg2) on TiO2 surface, energy difference between the CB of TiO2 and
the LUMO of dye before (ΔE1) and after adsorption (ΔE2) on TiO2 surface and
electron injection time τinj for monomer.
Dye Eads (eV) Eg1 (eV) Eg2 (eV) ΔE1 (eV) ΔE2 (eV) τinj (fs)

Monomer-1 − 4.46 0.55 0.70 0.12 1.24 1.23


Monomer-2 − 3.80 0.16 0.60 0.32 1.18 1.21
Monomer-3 − 2.49 0.41 0.43 0.20 0.99 1.66
Monomer-4 − 1.96 0.51 0.56 0.14 1.09 1.50

Fig. 3. The graph is the adsorption configuration for dye-TiO2 complexes of


Monomer-1, Monomer-2, and Monomer-3, respectively.

positive adsorbed energy, which is even worse than 3-TiO2 system.


However, aforementioned other two items of dye 4 are still better than
those of dye 3. It is testified that the suitable group would play more Fig. 4. Calculated total density of states (DOS) and projected density of states
important role in enhancing PCE especially when it is combined with (PDOS) for clean TiO2 surface and interfaces of TiO2 adsorbed with 1-monomer
appropriate configuration. (a), 2-monomde (b), and 3-monomer (c).

10
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

Table 3
The calculated adsorption energy (Eads), H-L energy gap before (Eg1) and after
adsorption (Eg2) on TiO2 surface, energy difference between the CB of TiO2 and
the LUMO of dye before (ΔE1) and after adsorption (ΔE2) on TiO2 surface, and
the electron injection time τinj for π-π dimeric architecture.
Dye Eads (eV) Eg1 (eV) Eg2 (eV) ΔE1 (eV) ΔE2 (eV) τinj (fs)

D-1 − 37.77 0.40 0.58 0.21 1.09 1.29


D-2 − 36.09 0.08 0.44 0.37 1.04 0.68

3.5. Dyes aggregation

Besides to improve the electron injection and transfer, the reduction


of deactivation process is the other pathway to enhance the overall
performance indirectly. Aggregation is an inevitable deactivation
pathway, which should be carefully considered. Although some re-
searches have reported it previously, it is rarely to be included in the
study to design new dyes.
After the π-π dimeric architecture is adsorbed on the TiO2 (101)
surface, the adsorption energy for 2 with the value of −36.09 eV is still
more positive than that for 1 (See Table 3), which is consistent with the
monomer-adsorption. Moreover, the differences between CB of TiO2
and LUMO of adsorption dye are 1.09 eV for D-1 (π-π dimeric archi-
tecture of 1) and 1.04 eV for D-2 (π-π dimeric architecture of 2). There
is no great difference between monomeric and dimeric adsorption for
not only the absolute value but also the difference between 1 and 2.
However, the injection time of D-2 (0.68 fs) is less than of D-1 (1.29 fs).
In general, there is not big difference between monomeric and dimeric
adsorption.
There is electronic coupling between stacking monomers, which is
detrimental for the electron injection into the TiO2 resulting in the less
Fig. 5. Evolution of the total electronic energy of the systems as a function of
PCE. The electronic coupling is calculated by the direct method, which
the simulation time.
has been widely employed in previous literature [42,43]. The direct
method is defined via the relationship:

0, sitea
Vab = ϕLUMO 0 0, site b 0, sitea 0, site b
/ HOMO F ϕLUMO / HOMO = ϕLUMO / HOMO h core ϕLUMO / HOMO

+∑l (occ ) ⎛⎜ ϕLUMO


0, sitea 0 0, site b − 0, sitea 0, site b 0 0 ⎞
/ HOMO ϕl ϕLUMO / HOMO ϕLUMO / HOMO ϕLUMO / HOMO ϕl ϕl ⎟
⎝ ⎠
(3)

where Vab is the charge transfer integral for the electron/hole and
0, sitea 0, site b
ϕLUMO / HOMO and ϕLUMO / HOMO represent the LUMOs/HOMOs of two ad-
jacent molecules a and b when no intermolecular interaction is pre-
sented. F° is the Fock operator and its density matrix is constructed from
orbitals of two adjacent molecules [44]. The calculated Vab for 2 is
larger than that for 1. All aforementioned items are based on the static
model, while the dye would be distorted. Therefore, it is necessary to
consider them by dynamic model.
As shown in Fig. 5, there is small fluctuation in the total simulated
10 ps except for in the initial 0.1 ps. Initially, the adsorbed dyes are all
perpendicular to the TiO2 (101) surface. Then, there is noticeable bend
Fig. 6. Intermolecular electronic couplings of 10 selected snapshots in dy-
between dye and TiO2 (101) surface and between individual component
namics. D-1 and D-2.
of dye during the dynamics simulations. Three snapshots at 0, 5, and
10 ps are extracted to illustrate the variation of structures (inset of
Fig. 5). The electronic coupling is greatly affected by the distortion of factors. However, it is still hard to differentiate 1 and 2 because of the
dye. Eleven electronic couplings are calculated with an interval of one similar performance. This problem is still kept even if the dimeric ad-
ps, which are plotted in Fig. 6. In the initial or final status, the elec- sorption model is employed. It is resolved when the dynamic properties
tronic coupling of 2 is larger than that of 1. However, the extent of is considered. The adsorption model and dynamic simulation are ne-
fluctuation for 2 is smaller than that of 1. The average electronic cou- cessary to build reliable structure-property relationship for dyes applied
pling of 2 is smaller than that of 1, which agrees well with the less in DSSCs.
electron injection time of 2, especially for the dimeric adsorption.
Three models are considered in this work to evaluate the perfor- 4. Conclusion
mance of dye. For the isolated dye, it is impossible to screen out the dye
with the better PCE since they have the contradictory items in quali- On the basis of reported 1, other two D1-π-D2-A dyes, 2 and 3, are
tative comparison. Next, when the monomeric adsorption is studied, 3 theoretical designed with different π groups. Their performance is
is excluded because of the inferior factors in almost every determined qualitatively compared by three different models to screen out the more

11
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

efficient dye. When only isolated dye is studied, no one would stand out class functionals and 12 other functionals, Theor. Chem. Acc. 120 (2008) 215–241.
since no one has all superiority. After adsorption, 3 has the less per- [15] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
simple, Phys. Rev. Lett. 77 (1996) 3865–3868.
formance than other two dyes with the inferior items. However, 1 and 2 [16] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Ab initio study of ionic solutions by a
have the similar adsorption energy, energy gap between the CB of TiO2 polarizable continuum dielectric model, Chem. Phys. Lett. 286 (1998) 253–260.
and the LUMO of dye, and injection time. The corresponding results are [17] E. Cancès, B. Mennucci, J. Tomasi, A. New, integral equation formalism for the
polarizable continuum model: theoretical background and applications to isotropic
kept even when the dimeric adsorption is considered. However, the and anisotropic dielectrics, J. Chem. Phys. 107 (1997) 3032–3041.
distortion of dye 1 is greater than that of dye 2 resulting in the larger [18] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
average electronic coupling during the simulated time. The aggregation G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada,
plays a more important role in distinguishing the performance of dif- M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
ferent dyes. Moreover, the dynamic model should be considered rather O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro,
than only the static model. Another D-π-A dye, 4, is designed with the M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, T. Keith,
R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar,
same donor, π group, and acceptor with 2. However, its performance is
J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross,
much less than 2. The suitable combination of configuration and π V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev,
group would reach the final goal. A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma,
V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich,
A.D. Daniels, O. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian
Acknowledgements 09, Revision C.01, Gaussian, Inc., Wallingford CT, 2010.
[19] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy cal-
We thank the National Supercomputing Center in Shenzhen culations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186.
[20] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals
(Shenzhen Cloud Computing Center) for providing computational re- and semiconductors using a plane-wave basis set, Comp. Mater. Sci. 6 (1996)
sources and software. This work was supported by the National Natural 15–50.
Science Foundation of China (21476061, 21503069, 21676071), [21] G. Kresse, J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B 47
(1993) 558–561.
Program for He’nan Innovative Research Team in University [22] R. Elmér, M. Berg, L. Carlén, B. Jakobsson, B. Norén, A. Oskarsson, G. Ericsson,
(15IRTSTHN005). J. Julien, T.F. Thorsteinsen, M. Guttormsen, G. Løvhøiden, V. Bellini, E. Grosse,
C. Müntz, P. Senger, L. Westerberg, K+ emission in symmetric heavy ion reactions
at subthreshold energies, Phys. Rev. Lett. 77 (1996) 4884–4886.
Appendix A. Supporting information [23] T. Bučko, J. Hafner, S. Lebègue, J.G. Ángyán, Improved description of the structure
of molecular and layered crystals: ab initio DFT calculations with van der waals
Supplementary data associated with this article can be found in the corrections, J. Phys. Chem. A 114 (2010) 11814–11824.
[24] B. Aradi, B. Hourahine, T. Frauenheim, DFTB+, a sparse matrix-based im-
online version at https://doi.org/10.1016/j.jlumin.2018.08.084.
plementation of the DFTB method, J. Phys. Chem. A 111 (2007) 5678–5684.
[25] M. Elstner, D. Porezag, G. Jungnickel, J. Elsner, M. Haugk, T. Frauenheim, S. Suhai,
References G. Seifert, Self-consistent-charge density-functional tight-binding method for si-
mulations of complex materials properties, Phys. Rev. B 58 (1998) 7260–7268.
[26] T.A. Niehaus, M. Elstner, T. Frauenheim, S. Suhai, Application of an approximate
[1] J. Gong, K. Sumathy, Q. Qiao, Z. Zhou, Review on dye-sensitized solar cells (DSSCs): density-functional method to sulfur containing compounds, J. Mol. Struct.
advanced techniques and research trends, Renew. Sust. Energ. Rev. 68 (2017) (Theochem) 541 (2001) 185–194.
234–246. [27] M. Gaus, Q. Cui, M. Elstner, DFTB3: extension of the self-consistent-charge density-
[2] Y.K. Eom, J.Y. Hong, J. Kim, H.K. Kim, Triphenylamine-based organic sensitizers functional tight-binding method (SCC-DFTB), J. Chem. Theory Comput. 7 (2011)
with π-spacer structural engineering for dye-sensitized solar cells: synthesis, theo- 931–948.
retical calculations, molecular spectroscopy and structure-property-performance [28] G. Dolgonos, B. Aradi, N.H. Moreira, T. Frauenheim, An improved self-consistent-
relationships, Dyes Pigments 136 (2017) 496–504. charge density-functional tight-binding (SCC-DFTB) set of parameters for simula-
[3] Y. Wu, M. Marszalek, S.M. Zakeeruddin, Q. Zhang, H. Tian, M. Grätzel, W. Zhu, tion of bulk and molecular systems involving titanium, J. Chem. Theory Comput. 6
High-conversion-efficiency organic dye-sensitized solar cells: molecular engineering (2010) 266–278.
on D-A-π-A featured organic indoline dyes, Energy Environ. Sci. 5 (2012) [29] R. Yan, X. Qian, Y. Jiang, Y. He, Y. Hang, L. Hou, Ethynylene-linked planar rigid
8261–8272. organic dyes based on indeno[1,2-b]indole for efficient dye-sensitized solar cells,
[4] N. Duvva, U. Chilakamarthi, L. Giribabu, Recent developments in tetrathiafulvalene Dyes Pigments 141 (2017) 93–102.
and dithiafulvalene based metal-free organic sensitizers for dye-sensitized solar [30] R. Maragani, R. Misra, M.S. Roy, M.K. Singh, G.D. Sharma, D-π-A)2-π-D-A type
cells: a mini-review, Sustain. Energy Fuels 1 (2017) 678–688. ferrocenyl bisthiazole linked triphenylamine based molecular systems for DSSC:
[5] K. Guo, K. Yan, X. Lu, Y. Qiu, Z. Liu, J. Sun, F. Yan, W. Guo, S. Yang, Dithiafulvenyl synthesis, experimental and theoretical performance studies, Phys. Chem. Chem.
unit as a new donor for high-efficiency dye-sensitized solar cells: synthesis and Phys. 19 (2017) 8925–8933.
demonstration of a family of metal-free organic sensitizers, Org. Lett. 14 (2012) [31] H.S. Nalwa, Handbook of Advanced Electronic and Photonic Materials and Devices:
2214–2217. Semiconductors 1 Academic Press, 2001.
[6] Z. Wan, C. Jia, Y. Wang, J. Luo, X. Yao, Significant improvement of phenothiazine [32] R. Katoh, A. Furube, T. Yoshihara, K. Hara, G. Fujihashi, S. Takano, S. Murata,
organic dye-sensitized solar cell performance using dithiafulvenyl unit as additional H. Arakawa, M. Tachiya, Efficiencies of electron injection from excited N3 dye into
donor, Org. Electron. 27 (2015) 107–113. nanocrystalline semiconductor (ZrO2, TiO2, ZnO, Nb2O5, SnO2, In2O3) films, J.
[7] X. Zhang, F. Gou, D. Zhao, J. Shi, H. Gao, Z. Zhu, H. Jing, π-Spacer effect in di- Phys. Chem. B 108 (2004) 4818–4822.
thiafulvenyl-π-phenothiazine dyes for dye-sensitized solar cells, J. Power Sources [33] A. Datta, S. Mohakud, S.K. Pati, Electron and hole mobilities in polymorphs of
324 (2016) 484–491. benzene and naphthalene: role of intermolecular interactions, J. Chem. Phys. 126
[8] A.D. Becke, Density-functional thermochemistry. III. The role of exact exchange, J. (2007) 144710–144717.
Chem. Phys. 98 (1993) 5648–5652. [34] A. Datta, S. Mohakud, S.K. Pati, Comparing the electron and hole mobilities in the α
[9] C. Lee, W. Yang, R.G. Parr, Development of the colle-salvetti correlation-energy and β phases of perylene: role of π-stacking, J. Mater. Chem. 17 (2007) 1933–1938.
formula into a functional of the electron density, Phys. Rev. B 37 (1988) 785–789. [35] V. Mohan, A. Datta, Structures and electronic properties of Si-substituted benzenes
[10] P.M.W. Gill, B.G. Johnson, J.A. Pople, M.J. Frisch, The performance of the Becke- and their transition-metal complexes, J. Phys. Chem. Lett. 1 (2010) 136–140.
Lee-Yang-Parr (B-LYP) density functional theory with various basis sets, Chem. [36] S. Mohakud, S.K. Pati, Large carrier mobilities in octathio[8]circulene crystals: a
Phys. Lett. 197 (1992) 499–505. theoretical study, J. Mater. Chem. 19 (2009) 4356–4361.
[11] T. Yanai, D.P. Tew, N.C. Handy, A new hybrid exchange-correlation functional [37] R.A. Marcus, Electron transfer reactions in chemistry. Theory and experiment, Rev.
using the Coulomb-Attenuating method (CAM-B3LYP), Chem. Phys. Lett. 393 Mod. Phys. 65 (1993) 599–610.
(2004) 51–57. [38] M. Pastore, F. De Angelis, Computational modelling of TiO2 surfaces sensitized by
[12] H. Iikura, T. Tsuneda, T. Yanai, K. Hirao, A long-range correction scheme for organic dyes with different anchoring groups: adsorption modes, electronic struc-
generalized-gradient-approximation exchange functionals, J. Chem. Phys. 115 ture and implication for electron injection/recombination, Phys. Chem. Chem.
(2001) 3540–3544. Phys. 14 (2012) 920–928.
[13] Y. Tawada, T. Tsuneda, S. Yanagisawa, A long-range-corrected time-dependent [39] P. Persson, M.J. Lundqvist, R. Ernstorfer, W.A. Goddard III, F. Willing, Quantum
density functional theory, J. Chem. Phys. 120 (2004) 8425–8433. chemical calculations of the influence of anchor-cum-spacer groups on femtosecond
[14] Y. Zhao, D.G. Truhlar, The M06 suite of density functionals for main group ther- electron transfer times in dye-sensitized semiconductor nanocrystals, J. Chem.
mochemistry, thermochemical kinetics, noncovalent interactions, excited states, Theory Comput. 2 (2006) 441–451.
and transition elements: two new functionals and systematic testing of four M06- [40] F. Labat, I. Ciofini, H.P. Hratchian, M. Frisch, K. Raghavachari, C. Adamo, First

12
W. Zhang et al. Journal of Luminescence 205 (2019) 7–13

principles modeling of eosin-loaded ZnO films: a step toward the understanding of Nanotechnology 18 (2007) 424029.
dye-sensitized solar cell performances, J. Am. Chem. Soc. 131 (2009) [43] S. Yin, Y. Yi, Q. Li, G. Yu, Y. Liu, Z. Shuai, Balanced carrier transports of electrons
14290–14298. and holes in Silole-based compoundss-a theoretical study, J. Phys. Chem. A 110
[41] F. Labat, T. Le Bahers, I. Ciofini, C. Adamo, First-principles modeling of dye-sen- (2006) 7138–7143.
sitized solar cells: challenges and perspectives, Accounts Chem. Res. 45 (2012) [44] G. Nan, L. Wang, X. Yang, Z. Shuai, Y. Zhao, Charge transfer rates in organic
1268–1277. semiconductors beyond first-order perturbation: from weak to strong coupling re-
[42] X. Yang, Q. Li, Z. Shuai, Theoretical modelling of carrier transports in molecular gimes, J. Chem. Phys. 130 (2009) 024704.
semiconductors: molecular design of triphenylamine dimer systems,

13

You might also like