You are on page 1of 5

Nanoscale

View Article Online


COMMUNICATION View Journal | View Issue
Published on 23 September 2016. Downloaded by PUSAN NATIONAL UNIVERSITY on 12/9/2019 3:05:39 AM.

A strategy to improve the efficiency of hole


Cite this: Nanoscale, 2016, 8, 17752
transporting materials: introduction of a highly
Received 2nd August 2016, symmetrical core†
Accepted 22nd September 2016
DOI: 10.1039/c6nr06116h Wei-Jie Chi, Ping-Ping Sun and Ze-Sheng Li*
www.rsc.org/nanoscale

The electronic, optical and hole transport properties of three new cost and poor hole mobility, which limits its practical appli-
hole transporting materials (HTMs) with a planar core have been cations. Therefore, it is imperative to design and synthesize
investigated by using density functional theory and Marcus theory. more efficient and economical alternatives for successful com-
A reliable semi-rational formula was adopted to calculate the mercialization of PSCs. Recently, a rich body of HTMs have
highest occupied molecular orbital (HOMO) levels of new HTMs. been reported in combination with perovskites from polymers
The results showed that the HOMO levels of new HTMs were to organic small molecules. For the latter, different cores have
0.07–0.30 eV lower than those of Spiro-OMeTAD, and their been used in the state-of-the-art HTMs, such as 3,4-ethylene-
absorption peaks appeared outside or close to the visible region dioxythiophene,9 truxene,10 benzotrithiophene,11 1,3,4-oxa-
and overlapped slightly with the absorption band of perovskites. diazole,12 fluorene-9,9′-xanthene,13 triazine,14,15 azulene,16
Moreover, the Stokes shifts of the designing molecules were calcu- thiophene,17 and so on.
lated to lie in a range from 72 to 124 nm, meaning that they could In this study, we present a new family of easily designed
undergo large geometrical changes on excitation. More impor- HTMs based on an anthrax-[1,2-b:4,3-b′:5,6-b″:8,7-b′′′]tetrathio-
tantly, the hole mobility of new HTMs (0.099–0.27 cm2 v−1 s−1) phene (ATT)18 as a core endowed with p-methoxydiphenyl-
was approximately two orders of magnitude higher than that of amine (ATT-1), p-methoxytriphenylamine (ATT-2), and N,N-bis
Spiro-OMeTAD (0.0056 cm2 v−1 s−1) due to strong hole coupling (4-methoxyphenyl)-amino-N-diphenyl (ATT-3). The ATT core
from a face-to-face packing pattern. Our results indicated that shows a D2 symmetry, and the coplanar structure of ATT is
planar core-based HTMs could become potential candidates to foreseen to promote an effective π–π intermolecular inter-
replace the widely established Spiro-OMeTAD. action, which could eventually lead to efficient hole transport
behavior. All new designing molecules are shown in Fig. 1. The
Since the first report in 2009,1 organic–inorganic halide molecular neutral ground-state geometries are optimized by
perovskite solar cells (PSCs) have attracted much attention due density functional theory with B3LYP functional (UB3LYP for
to their low material cost, simple configuration, and excellent cations) and 6-31G(d,p) basis set in dichloromethane solvent,
photovoltaic performance. In subsequent years, devices based
on such materials have shown an unparalleled increase of
power conversion efficiency (PCE) from 3.8% to 22.1%.2–4
In PSCs, a hole-transporting material (HTM) plays an
important role in facilitating hole transport from the valence
band of perovskites to the back contact and impeding back
electron transfer as well.5 Until now, small organic molecule
2,2-7.7-tetrakis(N,N′-di-para-methoxy-phenyl-amine) 9,9′-spiro-
bifluorene (Spiro-OMeTAD) has been widely utilized for the
fabrication of PSCs, exhibiting a high PCE surpassing 20%.6–8
However, Spiro-OMeTAD suffers from the expensive synthetic

Key Laboratory of Cluster Science of Ministry of Education, Beijing Key Laboratory of


Photoelectronic/Electrophotonic Conversion Materials, School of Chemistry, Beijing Fig. 1 The chemical structures of ATT-1, ATT-2, and ATT-3. R denotes
Institute of Technology, Beijing 100081, China. E-mail: zeshengli@bit.edu.cn p-methoxydiphenylamine, p-methoxytriphenylamine, and N,N-bis(4-
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ methoxyphenyl)-amino-N-diphenyl for ATT-1, ATT-2, and ATT-3,
c6nr06116h respectively.

17752 | Nanoscale, 2016, 8, 17752–17756 This journal is © The Royal Society of Chemistry 2016
View Article Online

Nanoscale Communication

based on the conductor-like polarizable continuum model Additionally, PSC based-new HTMs may also yield large short-
(C-PCM), as implemented in the Gaussian 09 program.19 circuit current density, which have a superlinear increase as
In order to maximize the hole injection processes and open the HOMO levels of HTMs decrease.22
circuit voltage of the device, it is essential to have a good For LUMO levels, it is found that their value gradually
energy alignment between the HTM and the perovskite. The decreases as the size of the electron donating moiety increases.
highest occupied molecular orbital (HOMO) levels of all the Moreover, the calculated LUMO levels are −2.22, −2.67, and
Published on 23 September 2016. Downloaded by PUSAN NATIONAL UNIVERSITY on 12/9/2019 3:05:39 AM.

investigated molecules are calculated by using a reliable semi- −2.75 eV for ATT-1, ATT-2, and ATT-3, respectively. As a conse-
rational formula.20 The lowest unoccupied molecular orbital quence, we believe that this larger energy barrier (1.16–1.90 eV)
(LUMO) level can be estimated by adding its excitation energy between the conduction band of perovskites (such as
to the HOMO level.20 Fig. 2 displays HOMO and LUMO distri- CH3NH3PbI3, −3.91 eV) and the LUMO of HTMs would block
butions and energy levels of HTMs. As shown in Fig. 2, the the electrons efficiently from perovskites to the metal electrodes
HOMOs are fully delocalized within the whole molecules for and thus largely reduce the recombination chance.
Spiro-OMeTAD, ATT-1, and ATT-2, while the HOMO of ATT-3 is Taking into consideration the appropriate HOMO and
mostly localized across the four N,N-bis(4-methoxyphenyl)- LUMO levels, the three new ATT derivatives seem to be good
amino-N-diphenyl units. In terms of LUMOs, all investigated candidates as HTMs to improve the performance of PSCs.
molecules show similar distribution diagrams, which is Except suitable energy levels, transparency in the visible
mainly localized on the central Spiro-unit or ATT core. The cal- region of the electromagnetic spectrum and large Stokes shift
culated HOMO levels given in Fig. 2 indicate that an additional are also typical features for a good HTM. To characterize the
phenyl ring between the ATT core and an electron donating optical properties of the investigated molecules, we performed
moiety stabilized HOMO levels by 140 meV and 90 meV, time dependent-DFT calculations on the ground-state and
moving from ATT-1 to ATT-2 and from ATT-2 to ATT-3, respect- excited-state structures optimized in dichloromethane. The
ively. The result implies that the size of the electron donating excited-state geometries of new HTMs are optimized at the
moiety plays an important role in adjusting the HOMO levels CAM-B3LYP/6-31G(d,p) level. The absorption properties are
of materials, and a larger electron donating moiety provides evaluated with the MPW1K functional and 6-31G(d,p) basis
lower HOMO levels. The tendency can also be found in an set. The method is reasonable and capable of evaluating the
experimental study.21 Furthermore, the HOMO levels are calcu- absorption properties of phenylamine-type molecules.23,24 The
lated as −5.16, −5.30, and −5.39 eV for ATT-1, ATT-2, and maxima absorption peak position (vertical solid lines with
ATT-3, respectively, suggesting that the HOMO levels of new arrow, λabs), maxima emission peak position (vertical dotted
designing molecules are more stabilized compared with that lines with arrow, λemi), and Stokes shifts (horizontal dotted
of Spiro-OMeTAD. On the other hand, the HOMO levels of ATT lines with arrow) of the investigated molecules are shown in
derivatives are higher than the valence band maximum of Fig. 3. As shown in Fig. 3, the calculated maxima absorption
CH3NH3PbI3 (−5.43 eV), indicating that they have an efficient peak of Spiro-OMeTAD is at 346 nm, which is close to the
hole extraction capability. Hence, when used as new HTMs in experimental value with 383 nm.25 For new HTMs, we find that
the PSCs, higher open circuit voltages are expected as the the maxima absorption peaks of ATT derivatives range from
open-circuit voltage is relative to the difference between the 364 to 420 nm, which are shown to be red-shifted by 18–74 nm
quasi-Fermi levels of TiO2 and the HOMO level of HTM, a
bigger difference implies a higher open-circuit voltage.

Fig. 3 The maxima absorption peak (MAP, vertical solid lines with
arrow), maxima emission peak (MEP, vertical dotted lines with arrow),
Fig. 2 Calculated frontier molecular orbital distributions and energy and Stokes shifts (horizontal dotted lines with arrow) of the investigated
alignments of HTMs and CH3NH3PbI3. molecules at the MPW1 K/6-31G(d,p) level.

This journal is © The Royal Society of Chemistry 2016 Nanoscale, 2016, 8, 17752–17756 | 17753
View Article Online

Communication Nanoscale

compared with that of Spiro-OMeTAD. The results imply that ue(HTM1) = 6.00 × 10−6 cm2 v−1 s−1, and ue(OMeTAD-TPA) =
in all cases the absorption takes place outside or close to the 1.08 × 10−4 cm2 v−1 s−1) are of the same order of magnitude. It
visible region and overlaps slightly with the absorption band is obvious to conclude that the simplified method is reason-
of perovskites. The maxima emission peaks are at 488, 492, able and capable of evaluating the hole mobility of organic
and 481 nm for ATT-1, ATT-2, and ATT-3, respectively, which is materials. Su’s study33 also supports our conclusion above,
91–102 nm red-shifted compared with the maxima emission of their calculation pointed out that the carrier mobility obtained
Published on 23 September 2016. Downloaded by PUSAN NATIONAL UNIVERSITY on 12/9/2019 3:05:39 AM.

Spiro-OMeTAD with 390 nm (experimental value with 419 nm). by the optimized geometry of the major pathway is in good
According to the maxima absorption peak and maxima emis- agreement with the experimental value. The stable dimers of
sion peak, the Stokes shift can be calculated by using the fol- ATT derivatives are shown in Fig. S1 of the ESI.† The hole
lowing expression: Stokes shift = λemi − λabs. mobility of ATT derivatives and Spiro-OMeTAD is shown in
Stokes shift can reflect the degree of structural deformation Fig. 4 and Table S2 of the ESI.† As shown in Fig. 4, the hole
between the ground state and excited state geometries. A mobilities of ATT-1 (0.13 cm2 v−1 s−1), ATT-2 (0.27 cm2 v−1 s−1),
larger Stokes shift is beneficial for pore-filling of HTM.9,26 As and ATT-3 (0.099 cm2 v−1 s−1) are approximately 2 orders of
shown in Fig. 3, the Stokes shift of Spiro-OMeTAD is 44 nm, magnitude higher compared with Spiro-OMeTAD (0.0056
which is in good agreement with the experimental value of cm2 v−1 s−1). This can be explained by the fact that the ATT
33 nm.27 Stokes shifts follow the trend ATT-1 (124 nm) > ATT-3 derivatives with a planar core have a large transfer integral (see
(75 nm) > ATT-2 (72 nm) > Spiro-OMeTAD (44 nm). The result Table S2†) between adjacent dimers due to short centroid to
suggests that new designing ATT derivatives as HTMs are centroid distances and strong π–π interactions. A previous
better for increasing the performance of PSCs compared with study also has proved that the planar and fully conjugated
Spiro-OMeTAD. Interestingly, we also find that although Spiro- molecule enables investigated materials to have ordered
OMeTAD and ATT-1 have the same arms as p-methoxydiphenyl- assemblies for better charge transport in the film state.10
amine, the Stokes shift of ATT-1 is larger than that of Spiro- Additionally, sulfur–sulfur interactions can also strongly affect
OMeTAD by 80 nm. The result allows us to conclude that the the packing model of molecules and are beneficial to enhance
planar core is more beneficial to increasing the Stokes shift of the charge transport rate, which is consistent with a previous
the HTM compared with the Spiro-core since the Spiro-unit work that the corresponding S⋯S distances in stacked dimers
ensures a great rigidity. are inversely proportional to hole mobility.34 The higher hole
Many studies28,29 have pointed out that the faster hole mobility of new designing molecules allows them to function
mobility is responsible for larger short-circuit current density as effective HTMs without the need of an extra doping process,
and higher fill factor of PSC. Hence, obtaining reliable hole which significantly simplifies device fabrication and improves
mobility is an important step to develop efficient HTM. the stability of PSC.
Marcus theory and the Einstein relationship are employed to For new designing materials, the solubility and stability
calculate the hole mobility of HTMs. The Einstein relationship should also be considered. The solubility of a material can be
has been widely used to analyze the carrier mobility of organic predicted by using solvation free energy (ΔGsolv). The defi-
molecules, and reliable results can be reproduced, especially nition of solvation free energy is the difference in free energy
for planar organic molecules. For example, Shi et al. calculated of the solute in the solution and gas phases.35 The more nega-
the carrier transport properties of tetrathiafulvalene deriva- tive solvation free energy is, the more soluble the material is.
tives, and they found that the calculated electron and hole In the present work, the solvent is dichloromethane and the
mobilities agree reasonably well with the experimental one for solvation free energy is calculated according to the Coote’s
DB-TTF and DN-TTF.30 In the present work, the hole transport report.35 The calculated ΔGsolv value of Spiro-OMeTAD is
pathways are simplified from three-dimension to one-dimen-
sion. It should be pointed out that the one-dimensional
pathway selected has the lowest energy of dimers in all poss-
ible hole transport pathways. A previous study20 has shown
that the more stable dimers have larger hole transfer integrals.
The hole transport pathway with large hole transfer integral
determines the hole mobility of a molecule. The compu-
tational methods of obtaining stable dimers and calculating
molecular hole mobility are listed in the ESI.† To further deter-
mine the reliability of the simplified method, we take Spiro-
OMeTAD, tris[N,N-bis(4-methoxyphenyl)-N-biphenyl]amine
(OMeTPA-TPA),31 and HTM132 as models to calculate their
hole mobility. The relevant data are listed in Table S1.† The
result shows that the predicted (uc(Spiro-OMeTAD) = 5.56 ×
10−3 cm2 v−1 s−1, uc(HTM1) = 1.33 × 10−6 cm2 v−1 s−1, and
uc(OMeTAD-TPA) = 5.75 × 10−4 cm2 v−1 s−1) and experimental
hole mobility (ue(Spiro-OMeTAD) = 1.30 × 10−3 cm2 v−1 s−1, Fig. 4 Calculated hole mobility of new HTMs.

17754 | Nanoscale, 2016, 8, 17752–17756 This journal is © The Royal Society of Chemistry 2016
View Article Online

Nanoscale Communication

−23.9 kcal mol−1, the ΔGsolv values of ATT-1, ATT-2, and ATT-3 Conclusions
are −23.1, −27.2, and −31.8 kcal mol−1, respectively. The result
shows that ATT-1 has similar solubility with Spiro-OMeTAD, In conclusion, three new HTMs were designed by introducing
while ATT-2 and ATT-3 possess higher solubility than Spiro- p-methoxydiphenylamine, p-methoxytriphenylamine, and N,N-
OMeTAD. The stability of an organic semiconductor can be bis(4-methoxyphenyl)-amino-N-diphenyl units onto the D2
estimated by absolute hardness (η),36 which is calculated using ATT-core. The frontier molecular orbital energies and optical
Published on 23 September 2016. Downloaded by PUSAN NATIONAL UNIVERSITY on 12/9/2019 3:05:39 AM.

the operational definition given by: η = (IP − EA)/2.37 EA rep- properties were studied by using density functional theory and
resents the adiabatic electron affinity and IP denotes the adia- time dependent density functional theory. Results suggested
batic ionization potential. We calculate the absolute hardness that three ATT derivatives have deeper energy levels and larger
of Spiro-OMeTAD and ATT derivatives, and the relevant data Stokes shifts compared to Spiro-OMeTAD. The planar and fully
are listed in Table S3 of the ESI.† As shown in Table S3,† ATT conjugated D2 HTMs resulted in more excellent hole mobility
derivatives have approximate η values (1.23–1.25 eV), which than Spiro-OMeTAD, the hole mobility of new HTMs
implies that they have similar stability. Moreover, we also find (0.099–0.27 cm2 v−1 s−1) was approximately 2 orders of magni-
that the η values of ATT derivatives are slightly smaller com- tude higher compared to Spiro-OMeTAD (0.0056 cm2 v−1 s−1).
pared with that of Spiro-OMeTAD (1.62 eV). The difference According to the above data, we believe that ATT derivatives
between the η values of Spiro-OMeTAD and ATT derivatives can be used as efficient candidates to replace the state-of-the-
only is about 0.38 eV. The result indicates that the stability of art Spiro-OMeTAD. Our work opens up the potential incorpo-
ATT derivatives is comparable to that of Spiro-OMeTAD. ration of the planar core into HTMs to explore high-efficiency
From an experimental viewpoint, we believe that newly PSCs in the future.
designed HTMs can be synthesized, and the conjectural syn-
thetic pathways are given in Fig. 5 according to previous
reports.11,15 As shown in Fig. 5, p-methoxydiphenylamine can Acknowledgements
be introduced by the Buchwald–Hartwig amination reaction to
obtain ATT-1. Three p-methoxytriphenylamine units can be This work is financially supported by the Major State Basic
covalently linked to the ATT core by a Suzuki cross-coupling Research Development Programs of China (2011CBA00701)
reaction to obtain ATT-2. The ATT-3 can be synthesized by the and the National Natural Science Foundation of China
Stille coupling reaction of N,N-bis(4-methoxyphenyl)-4′-(tri- (21473010, 21303007).
methylstannyl)biphenyl-4-amine with bromination of ATT.

Notes and references


1 A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am.
Chem. Soc., 2009, 131, 6050–6051.
2 H. Zhou, Q. Chen, G. Li, S. Luo, T.-b. Song, H.-S. Duan,
Z. Hong, J. You, Y. Liu and Y. Yang, Science, 2014, 345,
542–546.
3 W. S. Yang, J. H. Noh, N. J. Jeon, Y. C. Kim, S. Ryu, J. Seo
and S. I. Seok, Science, 2015, 348, 1234–1237.
4 W. Li, J. Fan, J. Li, Y. Mai and L. Wang, J. Am. Chem. Soc.,
2015, 137, 10399–10405.
5 F. Zhang, C. Yi, P. Wei, X. Bi, J. Luo, G. Jacopin, S. Wang,
X. Li, Y. Xiao, S. M. Zakeeruddin and M. Grätzel, Adv.
Energy Mater., 2016, 6, 1600401.
6 D. Bi, W. Tress, M. I. Dar, P. Gao, J. Luo, C. Renevier,
K. Schenk, A. Abate, F. Giordano, J.-P. Correa Baena,
J.-D. Decoppet, S. M. Zakeeruddin, M. K. Nazeeruddin,
M. Grätzel and A. Hagfeldt, Sci. Adv., 2016, 2, e1501170.
7 N. Ahn, D.-Y. Son, I.-H. Jang, S. M. Kang, M. Choi and
N.-G. Park, J. Am. Chem. Soc., 2015, 137, 8696–8699.
8 M. Saliba, T. Matsui, J.-Y. Seo, K. Domanski, J.-P. Correa-
Baena, M. K. Nazeeruddin, S. M. Zakeeruddin, W. Tress,
A. Abate, A. Hagfeldt and M. Gratzel, Energy Environ. Sci.,
2016, 9, 1989–1997.
9 H. Li, K. Fu, A. Hagfeldt, M. Grätzel, S. G. Mhaisalkar and
A. C. Grimsdale, Angew. Chem., Int. Ed., 2014, 53,
Fig. 5 Conjectural synthetic pathways for new HTMs. 4085–4088.

This journal is © The Royal Society of Chemistry 2016 Nanoscale, 2016, 8, 17752–17756 | 17755
View Article Online

Communication Nanoscale

10 C. Huang, W. Fu, C.-Z. Li, Z. Zhang, W. Qiu, M. Shi, 22 M. Planells, A. Abate, D. J. Hollman, S. D. Stranks,
P. Heremans, A. K. Y. Jen and H. Chen, J. Am. Chem. Soc., V. Bharti, J. Gaur, D. Mohanty, S. Chand, H. J. Snaith and
2016, 138, 2528–2531. N. Robertson, J. Mater. Chem. A, 2013, 1, 6949–6960.
11 A. Molina-Ontoria, I. Zimmermann, I. Garcia-Benito, 23 C.-F. Du, L. Jiang, L. Sun, N.-Y. Huang and W.-Q. Deng,
P. Gratia, C. Roldán-Carmona, S. Aghazada, M. Graetzel, RSC Adv., 2015, 5, 37574–37580.
M. K. Nazeeruddin and N. Martín, Angew. Chem., Int. Ed., 24 Z.-Z. Sun, Q.-S. Li, P.-P. Sun and Z.-S. Li, J. Power Sources,
Published on 23 September 2016. Downloaded by PUSAN NATIONAL UNIVERSITY on 12/9/2019 3:05:39 AM.

2016, 55, 6270–6274. 2015, 276, 230–237.


12 S. Carli, J. P. C. Baena, G. Marianetti, N. Marchetti, 25 G. Gong, N. Zhao, D. Ni, J. Chen, Y. Shen, M. Wang and
M. Lessi, A. Abate, S. Caramori, M. Grätzel, F. Bellina, G. Tu, J. Mater. Chem. A, 2016, 4, 3661–3666.
C. A. Bignozzi and A. Hagfeldt, ChemSusChem, 2016, 9, 26 D. Alberga, G. F. Mangiatordi, F. Labat, I. Ciofini,
657–661. O. Nicolotti, G. Lattanzi and C. Adamo, J. Phys. Chem. C,
13 B. Xu, D. Bi, Y. Hua, P. Liu, M. Cheng, M. Gratzel, L. Kloo, 2015, 119, 23890–23898.
A. Hagfeldt and L. Sun, Energy Environ. Sci., 2016, 9, 27 N. Arora, S. Orlandi, M. I. Dar, S. Aghazada, G. Jacopin,
873–877. M. Cavazzini, E. Mosconi, P. Gratia, F. De Angelis, G. Pozzi,
14 K. Lim, M.-S. Kang, Y. Myung, J.-H. Seo, P. Banerjee, M. Graetzel and M. K. Nazeeruddin, ACS Energy Lett., 2016,
T. J. Marks and J. Ko, J. Mater. Chem. A, 2016, 4, 1, 107–112.
1186–1190. 28 Y. Hua, B. Xu, P. Liu, H. Chen, H. Tian, M. Cheng, L. Kloo
15 K. Do, H. Choi, K. Lim, H. Jo, J. W. Cho, M. K. Nazeeruddin and L. Sun, Chem. Sci., 2016, 7, 2633–2638.
and J. Ko, Chem. Commun., 2014, 50, 10971–10974. 29 Y.-D. Lin, B.-Y. Ke, K.-M. Lee, S. H. Chang, K.-H. Wang,
16 H. Nishimura, N. Ishida, A. Shimazaki, A. Wakamiya, S.-H. Huang, C.-G. Wu, P.-T. Chou, S. Jhulki, J. N. Moorthy,
A. Saeki, L. T. Scott and Y. Murata, J. Am. Chem. Soc., 2015, Y. J. Chang, K.-L. Liau, H.-C. Chung, C.-Y. Liu, S.-S. Sun
137, 15656–15659. and T. J. Chow, ChemSusChem, 2016, 9, 274–279.
17 H. Li, K. Fu, P. P. Boix, L. H. Wong, A. Hagfeldt, M. Grätzel, 30 H.-X. Li, R.-h. Zheng and Q. Shi, Phys. Chem. Chem. Phys.,
S. G. Mhaisalkar and A. C. Grimsdale, ChemSusChem, 2014, 2011, 13, 5642–5650.
7, 3420–3425. 31 H. Choi, S. Paek, N. Lim, Y. H. Lee, M. K. Nazeeruddin and
18 W.-J. Liu, Y. Zhou, Y. Ma, Y. Cao, J. Wang and J. Pei, Org. J. Ko, Chem. – Eur. J., 2014, 20, 10894–10899.
Lett., 2007, 9, 4187–4190. 32 M. Franckevicius, A. Mishra, F. Kreuzer, J. Luo, S. M.
19 M. Frisch, G. Trucks, H. B. Schlegel, G. Scuseria, M. Robb, Zakeeruddin and M. Gratzel, Mater. Horiz., 2015, 2, 613–618.
J. Cheeseman, G. Scalmani, V. Barone, B. Mennucci and 33 Y. Geng, S.-X. Wu, H.-B. Li, X.-D. Tang, Y. Wu, Z.-M. Su and
G. E. Petersson, Gaussian 09, revision D 01, Gaussian Inc., Y. Liao, J. Mater. Chem., 2011, 21, 15558–15566.
Wallingford, CT, 2009. 34 W.-J. Chi, Q.-S. Li and Z.-S. Li, Synth. Met., 2016, 211,
20 W.-J. Chi, Q.-S. Li and Z.-S. Li, Nanoscale, 2016, 8, 6146– 107–114.
6154. 35 J. Ho, A. Klamt and M. L. Coote, J. Phys. Chem. A, 2010,
21 A. Krishna, D. Sabba, H. Li, J. Yin, P. P. Boix, C. Soci, 114, 13442–13444.
S. G. Mhaisalkar and A. C. Grimsdale, Chem. Sci., 2014, 5, 36 R. G. Pearson, J. Am. Chem. Soc., 1985, 107, 6801–6806.
2702–2709. 37 M. S. Stark, J. Phys. Chem. A, 1997, 101, 8296–8301.

17756 | Nanoscale, 2016, 8, 17752–17756 This journal is © The Royal Society of Chemistry 2016

You might also like