Fused Fluorenylindolenine-Donor-Based Unsymmetrical Squaraine Dyes For Dye-Sensitized Solar Cells

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Research Article

Cite This: ACS Appl. Mater. Interfaces 2018, 10, 26335−26347 www.acsami.org

Fused Fluorenylindolenine-Donor-Based Unsymmetrical Squaraine


Dyes for Dye-Sensitized Solar Cells
Rajesh Bisht,†,§ Vediappan Sudhakar,‡,§ Munavvar Fairoos Mele Kavungathodi,† Neeta Karjule,†,§
and Jayaraj Nithyanandhan*,†,§

Physical and Materials Chemistry Division and ‡Polymer Science and Engineering Division, CSIR-National Chemical Laboratory,
CSIR-Network of Institutes for Solar Energy, Dr. Homi Bhabha Road, Pune 411008, India
§
Academy of Scientific and Innovative Research (AcSIR), New Delhi 110025, India
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via BEN GURION UNIV OF THE NEGEV on September 11, 2019 at 07:28:09 (UTC).

ABSTRACT: A series of four unsymmetrical squaraine dyes, XSQ1−4, were


synthesized using a fused fluorenylindolenine-based donor unit for dye-sensitized
solar cells (DSSCs). The fused structure of fluorenylindolenine helped in moving
the absorption toward the near-infrared (NIR) region, and the two sp3-C centers
available on this donor were utilized to incorporate out-of-plane alkyl chains in
opposite directions to control the dye−dye interactions on the TiO2 surface.
High extinction coefficient (ε ≥ 105 M−1 cm−1) for absorbing NIR photons and
suitable highest occupied molecular orbital and lowest unoccupied molecular
orbital energy levels with respect to the conduction band of TiO2 and electrolyte
for charge injection and dye regeneration processes, respectively, make these dyes
potential sensitizers for DSSCs. Introduction of branched alkyl groups in the π-
framework helped in controlling dye aggregation to reduce exciton quenching and
assisted in TiO2 surface passivation to avoid the charge recombination process.
Furthermore, having a naphthyl group on the indole part of the anchoring group
containing segment helped to red-shift the absorption spectrum of dyes 15 nm toward the NIR region (XSQ3−4). Among all of
the dyes under investigation, XSQ2 gave the best photovoltaic performance, having a short-circuit current density (JSC) of 13.99
mA cm−2, open-circuit voltage (VOC) of 0.66 V, and a fill factor (ff) of 0.71, with a device performance (η) of 6.57%.
Electrochemical impedance spectroscopy revealed higher electron lifetime on TiO2 for XSQ2, which helps to avoid the charge
recombination process.
KEYWORDS: squaraine dyes, fluorenylindolenine, NIR absorption, H- and J-type aggregation, out-of-plane alkyl groups,
dye-sensitized solar cells

■ INTRODUCTION
Dye-sensitized solar cells (DSSCs) have shown potential as a
sensitizers, it is desirable to have a broad absorption with a
high extinction coefficient (ε). The sensitizers based on
photovoltaic technology for future demand of clean and ruthenium−polypyridyl complexes remained the champion
renewable energy because of being cheaper and easy to fabricate sensitizers for almost a decade, and efficiency up to 11.8% was
compared with silicon solar cells.1 The working principle of achieved.4 The highest efficiency of 13% for organometallic dyes
DSSCs involves (i) a light-harvesting sensitizer, which promotes was reported for zinc-porphyrin-based dye SM315 where the
an electron to the excited state upon absorption of light, (ii) a donor−π−acceptor (D−π−A) configuration was used, with
metal-oxide-based semiconductor, which accepts the excited benzothiadiazole as an acceptor, to enhance the performance of
electron and transfers it to the external circuit, (iii) an electrolyte the DSSC.5 Although these metal complexes showed broad
that regenerates the oxidized dye, and (iv) the oxidized spectral coverage and high efficiency, low absorptivity at near-
electrolyte collects the electron from the external circuit at the infrared (NIR) wavelengths and difficulty in synthesis have led
counter electrode.2 The extensive research during the last two the attention toward metal-free sensitizers as alternatives.
and a half decades on DSSCs helped to understand charge Several D−π−A type of metal-free sensitizers have also been
transport and recombination processes occurring at various explored and proved to be highly efficient in terms of
interfaces that provided guidelines to design anode materials, performance and ease of synthesis.6 In these dyes, electron-
sensitizers, electrolytes, and counter electrodes of the device for rich moieties like triphenylamine derivatives are used as donors
the improved performance.1,3 The light-harvesting sensitizer
plays an important role in the conversion efficiency of the Received: June 13, 2018
DSSCs, and efforts have been devoted to developing metal- Accepted: July 17, 2018
based and metal-free efficient sensitizers. For efficient Published: July 17, 2018

© 2018 American Chemical Society 26335 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Figure 1. (a) Design of steric- and electronic-effect-incorporated unsymmetrical squaraine dyes, (b) design of a fused fluorenylindolenine-based
donor, and (c) design of unsymmetrical squaraine dyes XSQ1−4.

and electron-withdrawing groups like cyanoacetic acid are used squaraine dye showed an efficiency of 4.5% (VOC = 0.603 V and
as acceptors. They show high molar extinction coefficients due JSC = 10.5 mA cm−2) because the asymmetric structure led to an
to facile intramolecular charge transfer (ICT), which occurs efficient unidirectional charge injection to TiO2.33 Although
within the molecule because of the push−pull structure.7,8 Such squaraine dyes show very good absorption in the far-red region,
a structure provides the advantage of easy tweaking and the exciton quenching by aggregation limits its overall
modification of the electronic nature of the π-bridge. photovoltaic performance in DSSCs.34,35 The π−π interaction
Furthermore, the D−A−π−A configuration helped to synthe- between the molecules leads to aggregation of dyes on TiO2,
size low-band-gap dyes to extend the absorption in the visible which is a common phenomenon and has a variable effect on the
and far-red region of the solar spectrum.9 Such a configuration performance of DSSCs.36−39 The periodic availability of the 5-
used additional acceptors like benzthiadiazole,10,11 diketopyrro- coordinated Ti site on the anatase {101} facet provides dye-
lopyrrole,12,13 benzotriazole,14,15 benzoxadiazole,16,17 quinoxa- anchoring sites with close proximity. Such a monolayer of dyes
line,18,19 etc. between the conventional donor and the π-bridge. on the TiO2 surface allows the molecular dipoles to interact in a
The efficiency of ≥10% was obtained with a diketopyrrolo- head-to-head (H−H) fashion, which causes H-type aggregates
pyrrole-based dye employing the [Co(bpy)3]3+/2+ electrolyte.20 and can be observed as blue-shifted absorption in the UV−vis
Among the metal-free dyes, ≈14% device performance has been spectrum with respect to the monomer, whereas the head-to-tail
achieved by D−π−A dyes through the cosensitization (H−T) interaction leads to a J-type aggregate, which shows red-
approach.21 shifted absorption.39−42 Several attempts have been made to
In DSSCs, near-infrared-absorbing organic dyes are critically improve the efficiency of squaraine dyes by increasing the
needed as they are capable of generating high current density, absorption in the NIR region. The fused π-extended heterocyclic
which can improve the performance of the cell.22,23 Although structures such as benz[e]indole,43−45 benzo[c,d]indole,46,47
several sensitizers have been developed with broad absorption, and quinoline48−50 have been used as strong donors in place of
there are limited chromophores that absorb in the NIR regions indole to shift the absorption to NIR through strong ICT. These
such as phthalocyanines,24 porphyrins,25−29 and polymethine30 structures could not convert the increased NIR absorption into
dyes. Squaraine dyes, a subclass of polymethine dyes, are among better performance due to either aggregation or mismatched
the few chromophores having the potential to absorb in this highest occupied molecular orbital (HOMO)−lowest unoccu-
region.31 Indole-based squaraine dyes are one of the most pied molecular orbital (LUMO) levels. In another approach,
common designs of efficient squaraine dyes where the indole conjugation is increased by attaching a π-spacer to squaraine dye
moiety is condensed with squaric acid.32 Unsymmetrical to increase the length of the molecule, which increased the
squaraine dyes typically perform better than symmetrical ones absorption toward a longer wavelength.51−55 Remarkably, an
as the unidirectional electron flow proved to be beneficial for improved efficiency of 7.3% was obtained with the incorporation
better electron injection to TiO2. The first unsymmetrical of cyclopentadithiophene as a π-spacer in squaraine dye, which
26336 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Scheme 1. (a) Synthesis of Fluorenylindolium Salts 6a and 6b, (b) Semisquaric Acid 8, (c) Semisquaric Acid 12, and (d)
Unsymmetrical Squaraine Dyes (XSQ1−4)

26337 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Figure 2. (a) Absorption profile of XSQ dyes in CH2Cl2 solvent, (b) normalized fluorescence spectra excited at 600 nm, (c) absorption spectra on
TiO2, and (d) light-harvesting efficiency (LHE) of the XSQ dyes.

utilized out-of-plane hexyl chains to control aggregation as well


as enhance the absorption in the NIR region through extended
■ RESULTS AND DISCUSSION
Synthesis and Characterization. The synthetic scheme
conjugation.54
for dyes XSQ1−4 is represented in Scheme 1. Precursors 9,9-di-
The results indicate that both electronic and steric factors are
n-octyl-2-bromofluorene64 (1), 7,65 10,45 and 465 were
important in dye design, and certain modification in structural
synthesized following the previously reported protocols. To
features of the dyes may lead to the significant improvement in
attach the hydrazine functionality, precursor 1 was treated with
cell efficiency. An unsymmetrical squaraine dye, DTS-CA, which
n-BuLi at −78 °C followed by addition of dibutylazadicarbox-
has a squaraine unit linked with a silolodithiophene (DTS) π-
ylate to give tert-BOC-protected hydrazine derivative 2.
spacer possessing out-of-plane ethylhexyl chains, exhibits an
efficiency of 8.9%.51 In our recent work, benzodithiophene Compound 2 was unmasked by treatment with hydrochloric
(BDT) bearing two ethylhexyl branched chains was used as a π- acid to afford hydrazine derivative 3 in good yield.
spacer with squaraine dye, which produced the efficiency of The hydrazine derivative was refluxed with 4 (branched
6.72% without a coadsorbent.56 Incorporation of an out-of- methyl ketone) in acetic acid to give fluorenylindolenine 5 in
plane alkyl chain on the sp3-center in the indole unit of squaraine moderate yield, which was then heated with methyl iodide and
dyes helped to effectively control dye aggregation and charge hexyl iodide in a sealed tube to afford the corresponding
recombination (η = 9.1%).57 The approach of appending alkyl fluorenylindolium salts 6a and 6b, respectively. The precursors
groups to modulate the charge recombination process has been for the anchoring group containing indoline derivatives 761 and
used to boost the performance of porphyrin58−60 and 949 were synthesized and further functionalized to afford the
phthalocyanine dyes61−63 as well. required semisquaric acid derivatives 8, and 12, respectively
On the basis of these findings, we attempted to increase the (Scheme 1). Semisquarate 7 was treated with 2 N HCl to
absorption of the squaraine dyes toward the NIR region and also provide the required semisquaric acid 8 in quantitative yield.
include out-of-plane branched alkyl chains to control The carboxybenz[e]indole 949 derivative was reacted with
aggregation. To this end, we designed and synthesized a fused methyl iodide to give N-methylated carboxybenz[e] indolium
donor where a fluorene moiety is fused with indolenine to give a salt 10, which was then reacted with dibutyl squarate. The
fluorenylindolenine donor. This donor can accommodate two squarate intermediate was then reacted with 2 N HCl to provide
out-of-plane branched alkyl chains and one in-plane linear chain the required semisquaric acid 12. Semisquaric acid 8 was
on the nitrogen, which can help in controlling the aggregation of refluxed with 6a in a n-BuOH/PhMe mixture under azeotropic
dyes on TiO2 as well as extending conjugation toward NIR distillation of water employing a Dean−Stark apparatus to give
(Figure 1). squaraine dyes XSQ1, whereas dye XSQ2 was obtained by
A series of four dyes, XSQ1−4, were synthesized using this reacting 8 with 6b. Similarly, dyes XSQ3 and XSQ4 were
donor. XSQ1 and XSQ2 contained indole units toward the obtained by treating semisquaric acid 12 with 6a and 6b,
anchoring side with an N-methylated and N-hexylated respectively. All of the dyes were completely characterized, and
fluorenylindolenine donor, respectively. XSQ3 and XSQ4 their structural and electronic properties were thoroughly
contained a benz[e]indole unit toward the anchoring side with investigated by NMR spectroscopic and mass spectrometric
an N-methylated and N-hexylated donor, respectively. techniques.
26338 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Table 1. Photophysical and Electrochemical Properties of XSQ Dyes


λmax/CH2Cl2 λmax/em (nm) in λmax/TiO2 Eg/DFT EHOMO E0−0 ELUMO
dyes (nm) CH2Cl2 (nm) Amax/monomerAmax ε × 105 (M−1 cm−1)
dimer
(eV) (V vs NHE) (eV) (V vs NHE)
XSQ1 665 673 625/670 0.67 2.2 2.19 0.78 1.85 −1.09
XSQ2 666 674 628/672 0.60 3.0 2.19 0.75 1.85 −1.10
XSQ3 679 686 635/684 0.70 2.4 2.18 0.67 1.82 −1.15
XSQ4 681 687 637/686 0.69 3.5 2.17 0.66 1.81 −1.15

Figure 3. (a) Cyclic voltammogram of XSQ dyes measured in CH2Cl2 (platinum wire as the working electrode, nonaqueous Ag/Ag+ (0.01 M) as the
reference electrode, and platinum foil as the counter electrode) with TBAClO4 (0.1 M) at a scan rate of 50 mV s−1. (b) Energy-level diagram for XSQ
dyes with DSSC device components.

Photophysical Studies. The optical properties of XSQ1−4 leads to dimer formation. The formation of a dimer was
in CH2Cl2 and on a transparent thin film of TiO2 (6 μm) are confirmed by carrying out the UV−vis studies of chenodeox-
shown in Figure 2 and summarized in Table 1. Dyes XSQ1−4 ycholic acid (CDCA)-co-adsorbed dye-sensitized TiO2 films,
displayed intense absorption (ε ≥ 105 M−1 cm−1) toward the far- where the intensity of peaks observed at 600−640 nm was
red region in solution due to efficient ICT. decreased because of the reduction in dimer formation
The absorption spectra mainly spanned from 545 to 698 nm compared to that for the dye-sensitized TiO2 films without
for XSQ1−2, whereas XSQ3−4 had the major absorption CDCA (Figure S2). The ratio of the optical density of the
between 553 and 711 nm. The absorption maxima (λmax) of monomer to the dimer (Adimer/Amonomer) is given in Table 1, and
XSQ1−2 were found to be at 665 and 666 nm, whereas XSQ3− it helped in concluding the extent of aggregation for each dye on
4 showed a red shift of ca. 15 nm with λmax at 679 and 681 nm. the surface, which revealed that XSQ2 is the least aggregated in
Compared with its parent squaraine dye SQ1 (λmax = 640 nm),57 comparison to other dyes. Apart from the peaks at the low-
XSQ1−2 and XSQ3−4 had 25 and 40 nm red-shifted energy region, there is a band of high-energy absorption between
absorption, respectively, because of the extended conjugation 400 and 550 nm with relatively low absorbance. The light-
by fused fluorenylindolenine and benz[e]indole moieties. The harvesting efficiency (LHE = 1 − 10−A) represents the capacity
shoulder peaks between 600 and 640 nm were observed due to of sensitizers to absorb a wide range of photons on the TiO2 thin
the vibronic progression. To understand the aggregation of dyes film (Figure 2d). The LHE profile of all of the dyes indicated
in solution, the concentration-dependent UV−vis study has over 90% light harvesting in the wide spectral range of 560−730
been carried out using quartz cuvette of 0.1 cm path length. It nm, and XSQ3−4 showed a 20 nm red-shifted onset compared
was observed that dyes exist as monomers in moderate with XSQ1−2. The variable LHE response was observed in the
concentrations (6.25−50 μM); however, at the higher high-energy region of 400−550 nm where XSQ4 showed the
concentration (100 μM), the deviation in the Beer−Lambert maximum intensity.
plot suggests the formation of aggregates in solution (Figure S1). Electrochemical Properties. To evaluate the feasibility of
The absorption spectra of XSQ dyes on chemisorbed TiO2 films charge injection and dye regeneration process for XSQ1−4
were obtained by dipping a transparent, 6 μm thick, TiO2 film in dyes, cyclic voltammetry was carried out in a CH2Cl2 solution
0.1 mM dye solution in EtOH for 15 min (Figure 2c). The with tetrabutylammonium perchlorate (TBAClO4) as the
broadening of the spectra with coverage from 550 to 715 nm was electrolyte and ferrocene as an external reference. The cyclic
observed for XSQ1−2, whereas spectra covered from 550 to 730 voltammograms of XSQ dyes exhibit two-electron oxidation
nm for XSQ3−4. The λmax on the TiO2 film is slightly red-shifted peaks, which is a characteristic of squaraine dyes (Figure 3a). For
to about 5 nm compared to that for the solution, and there was efficient electron transfer from the excited dye to TiO2, the
emergence of an additional peak also at a shorter wavelength excited-state potential (LUMO) of the dye should be more
(620−640 nm) because of H-type aggregation on TiO2, which negative than the conduction band (CB) edge (−0.5 V vs
26339 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Figure 4. Optimized geometry structure of XSQ2 and XSQ4 and CDCA (a) lateral view and (b) top view (distance between the sp3-C atom of
fluorenyl and the oxygen atom of carboxylic acid and distance between the sp3-C atom of indoline and the oxygen atom of carboxylic acid were
provided).

Figure 5. Isodensity plots of XSQ2 and XSQ4 (alkyl chains are removed for clarity).

normal hydrogen electrode (NHE)) of TiO2 (at least 0.2 eV66), group in XSQ2 (Figure 4). In the XSQ4 dye, the branched
whereas the ground-state potential (HOMO) of the dye should chains were at distances of 16 and 21 Å from the terminal atom
be more positive than the redox potential (0.4 V vs NHE) of of the anchoring group because of the increase in the length
iodide/triiodide (at least 0.15 eV67) to provide sufficient driving caused by the extended benz[e]indole unit. The top view of
force for efficient regeneration of the oxidized dye. The HOMO these molecules showed that the alkyl chains are stretched up to
level of XSQ dyes was calculated from the first oxidation peak 18−20 Å in the out-of-plane direction, which can help in
and found to be at 0.76 V (XSQ1), 0.75 V (XSQ2), 0.67 V controlling aggregation of dyes on the TiO2 surface efficiently.
(XSQ3), and 0.66 V (XSQ4), which are lower than the Because the dimension of the CDCA molecule (14.2 Å) is on
electrochemical potential of redox mediator (I−/I3−) to provide the order of the distance between the lower branched chain and
enough driving force for dye regeneration (Figure 3b). anchoring groups of XSQ1 and XSQ2 (14.7 Å), the adsorption
The HOMO levels of XSQ3−4 were 90 mV above those of of CDCA may affect the dye loading. However, in the case of
XSQ1−2, which could be attributed to the stronger electron- XSQ3 and XSQ4, the distance between the lower branched
releasing effect of the carboxybenz[e]indole unit present on the chain and the anchoring group unit is larger than the dimensions
anchoring end of XSQ3−4. The LUMO energy levels of the of the CDCA molecule and the CDCA molecules can occupy
dyes were calculated by subtracting E0−0 from the HOMO levels the space under the dye molecules when co-adsorbed on TiO2,
of the dyes and found to be at −1.09 V (XSQ1), −1.10 V which may help in passivating the TiO2 surface more effectively.
(XSQ2), −1.15 V (XSQ3), and −1.15 V (XSQ4), which lie well Furthermore, the additional N-alkyl group also increases the
above the conduction band level of TiO2, hence suitable for density of the alkyl group around the lower branching unit of
efficient electron injection into TiO2. dyes XSQ2 and XSQ4, which decreases the amount of dye
Computational Investigations. Optimized molecular loaded on the surface compared to that in XSQ1 and XSQ3. The
geometries and electronic distribution of frontier molecular electron density distribution in the frontier molecular orbitals
orbitals in XSQ dyes were computed by Gaussian 09.68 The two (HOMO and LUMO) of the dyes under investigation is shown
opposite branched alkyl chains on the fluorenylindolenine unit by isodensity plots in Figure 5. The electron density in HOMO
were 14.7 and 20 Å away from the anchoring carboxylic acid of both indole (XSQ1 and XSQ2) and benzindole (XSQ3 and
26340 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Figure 6. (a) J−V curve and (b) IPCE spectrum of XSQ1−4-sensitized cells with 3 equiv of CDCA (TiO2 electrode thickness = 8 + 4 μm (transparent
+ scattering layer), area = 0.36 cm2, [dye] = 0.1 mM in EtOH, [CDCA] = 0.3 mM, dipping time = 12 h at room temperature (rt), electrolyte: iodolyte
Z-50, mask aperture 0.64 cm2 from Solaronix).

Figure 7. (a) J−V curve and (b) IPCE spectrum of XSQ2-sensitized dyes with varying CDCA concentrations in the dye solution in EtOH (TiO2
electrode thickness = 8 + 4 μm (transparent + scattering layer), area = 0.36 cm2, [dye] = 0.1 mM in EtOH, dipping time = 12 h at rt, electrolyte:
iodolyte Z-50, mask aperture 0.64 cm2 from Solaronix).

Table 2. Photovoltaic Performance of XSQ Dyes with and without CDCA under 1 sun Illumination and the Results of Five
Devices for Each Set with Deviation Are Summarized
dye VOC (V) JSC (mA cm−2) INT
JSC (mA cm−2)a ff (%) η (%) dye loading (mol cm−2)
XSQ1 0.621 ± 0.003 7.2 ± 0.11 6.55 70.1 ± 0.13 3.21 ± 0.13 8.68 × 10−8
XSQ1/CDCA (1:3) 0.652 ± 0.003 12.6 ± 0.13 11.85 70.0 ± 0.4 5.75 ± 0.15 8.06 × 10−8
XSQ2 0.634 ± 0.002 9.21 ± 0.2 8.82 71.1 ± 0.2 4.12 ± 0.12 7.81 × 10−8
XSQ2/CDCA (1:3) 0.657 ± 0.002 13.8 ±0.19 13.57 71.2 ± 0.3 6.42 ± 0.15 7.11 × 10−8
XSQ3 0.616 ± 0.002 5.3 ± 0.21 5.23 69.1 ± 0.13 2.25 ± 0.11 1.36 × 10−7
XSQ3/CDCA (1:3) 0.643 ± 0.002 8.98 ± 0.2 8.5 70.2 ± 0.2 4.05 ± 0.14 1.10 × 10−7
XSQ4 0.591 ± 0.002 8.3 ± 0.21 8.11 68.9 ± 0.4 3.38 ± 0.16 6.27 × 10−8
XSQ4/CDCA (1:3) 0.633 ± 0.002 12.1 ± 0.18 11.77 70.1 ± 0.4 5.37 ± 0.13 6.22 × 10−8
a
Calculated by integrating IPCE over the corresponding wavelengths.

XSQ4) dyes is located mostly in the central squaraine unit with a Photovoltaic Performance. Current−voltage (J−V) char-
slight spread in the fluorenylindolenine unit. In the LUMO acteristics and the incident photon-to-current conversion
energy level, there is a clear shift of the electron density from the efficiency (IPCE) profiles are displayed in Figures 6 and 7,
fluorenylindolenine unit toward the anchoring group for both and photovoltaic data are summarized in Table 2. All of the dyes
the dyes, which is significant for the facile charge transfer. It also gave moderate DSSC performance in the absence of any
ensures the good electronic coupling between the excited-state coadsorbent with the XSQ2 dye, exhibiting the maximum power
dye and TiO2 because of the presence of sufficient electron conversion efficiency (PCE) of 4.12% with JSC of 9.22 mA cm−2
density on the anchoring group, which is important for the and VOC of 0.636 V, whereas XSQ3 showed the minimum
efficient injection of electrons from the excited dye molecule efficiency of 2.36% with JSC of 5.51 mA cm−2, VOC of 0.618 V,
into the conduction band of TiO2. However, the bent structure and fill factor (ff) of 69.23%. XSQ1 and XSQ4 showed the
of XSQ3−4 due to the benz[e]indole group may lead to efficiency of 3.34 and 3.54%, respectively. Although the VOC of
nondirectionality in the molecular dipole, which can be observed these cells varies slightly, the significant variation in the
by comparing the calculated dipole moments of XSQ1 (10.0 D) efficiencies of the dyes comes from the difference in their JSC.
and XSQ2 (10.6 D) with XSQ2 (7.4 D) and XSQ4 (7.3 D) The gradual increase in the amount of coadsorbent (CDCA)
(Figure S7). The resultant dipole moment exerted by the dye improved the performance of the cells, and all of the XSQ dyes
molecules on the TiO2 surface plays a vital role in modulating gave their best photovoltaic properties at 3 equiv of CDCA.
the conduction band energy level and hence the VOC.69,70 XSQ2 showed the highest efficiency among all of the dyes with
26341 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Figure 8. Impedance analysis of XSQ1−4 DSSC devices. (a) Nyquist plot at an applied bias of −0.5 V (inset: equivalent circuit), (b) log Rct vs applied
potential, (c) log Cμ vs applied potential, and (d) τ vs applied potential.

PCE of 6.57% mostly due to 50% increment in JSC (13.99 mA and 7a,b). These observations can be explained if we take into
cm−2) and 5% incremental improvement in VOC (0.657 V) after account the possibility that instead of replacing the dye CDCA is
addition of 3 equiv of the CDCA. The addition of CDCA helps taking up the void spaces between the dyes. It is also supported
avoiding the dye−dye interaction (dye aggregation) and by the density functional theory (DFT) calculations which show
passivates the TiO2 surface so that the oxidized electrolyte that the branched alkyl chains are stretching up to 20 Å in the
may not diffuse through the hydrophobic layer to reach the out-of-plane direction, which may create enough gap for the
surface of TiO2 that avoids the charge recombination, which CDCA molecule to anchor between dyes, and also the distance
resulted in an increase in JSC and VOC. The trends in JSC can be between the lower branch chain and the terminal oxygen atom of
correlated to IPCE spectra of the cells (Figure 6b). carboxylic acid (14.7 and 16 Å for XSQ2 and XSQ4,
The IPCE spectra of the XSQ dyes showed a good response in respectively) is greater than the length of CDCA (14.2 Å).
the region of 500−750 nm with XSQ2 having the best profile Hence, it is conceivable that the CDCA molecule might anchor
with 76% IPCE at 620 and ca. 40% IPCE between 400 and 500 just between the dyes and the dye amount essentially remains
nm. XSQ1 and XSQ4 displayed similar IPCE trace, result of the same. Such preferential self-assembled monolayers of dyes
which could be observed in their similar JSC values of 12.73 and act as a hydrophobic barrier layer because of the presence of out-
12.28 mA cm−2, respectively. XSQ3 has the poorest response of-plane alkyl groups at sp3-C atoms and alkyl chains at the N-
managing up to only 40% IPCE between 600 and 720 nm and atom, which avoids the charge recombination of electrons from
around 22% between 400 and 550 nm. It was observed that by TiO2 to the oxidized electrolyte. It is also noted that dyes with N-
increasing the concentration of CDCA up to 3 equiv, JSC values hexyl alkyl chains showed better efficiency than N-methylated
were increased significantly without affecting the distribution of dyes as XSQ1 showed the η of 5.90%, which is lower in
IPCE response from the aggregated state. This indicated that the comparison to that of XSQ2. A similar trend was observed for
formation of larger aggregates was decreased and the smaller the benz[e]indole-based dye where XSQ3 showed η of 4.19%,
aggregates could effectively participate in the charge injection which is lower than the efficiency of the XSQ4 (η = 5.5%)-based
process. Apart from substantial increase in JSC, the modest cell. This observation suggests that the N-hexyl chain could help
increase in VOC was also obtained upon addition of CDCA the dyes assemble favorably on to the TiO2 surfaces besides
because of improved charge injection and reduced charge passivating the surface. Additionally, the VOC values of indole-
recombination. A moderate decrease in VOC and JSC upon based XSQ1−2 are slightly higher than those of XSQ3−4
further increasing the CDCA concentration indicated depletion (ΔVOC ≈ 10 mV), which can be attributed to the upshift in the
of the amount of dye on the TiO2 surface. It is interesting to note conduction band (CB) edge due to greater dipole moments of
that the decrease in the amount of adsorbed dyes in the presence XSQ1−2 in the direction normal to TiO2.69,70
of 3 equiv of CDCA is very low, ca. 8−19%, whereas the increase Electrochemical Impedance Spectroscopy (EIS). To
in short-circuit current density is significantly high, 44−74%. For understand the charge transfer and recombination processes at
example, in the case of XSQ2, the decrease in dye loading is only the various interfaces of a DSSC device, EIS study was carried
8.9% after addition of 3 equiv. of CDCA, whereas the increase in out in the dark at various applied potentials, viz., 0.38, 0.41, 0.47,
JSC is ca. 36%. Similarly, XSQ1 observed 74% increase in JSC in and 0.50 V.
the presence of 3 equiv of CDCA, which shows that CDCA is As the device architecture and electrolytes, including
able to control aggregation without replacing the dye (Figures 6 fabrication conditions, are similar for all of the cells, any
26342 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

characteristic feature obtained in EIS is mainly due to the aperture 0.64 cm2) and without the mask (active area 0.22 cm2) have
interaction of dyes at the TiO2−dye/electrolyte interface. The been provided in the Supporting Information.
Nyquist plots at an applied bias of 0.50 V are shown in Figure 8a. Synthetic Procedure. Di-tert-butyl 1-(9,9-dioctyl-9H-fluoren-2-
yl)hydrazine-1,2-dicarboxylate (2). n-BuLi (1.6 M, 4.8 mL, 7.7 mmol)
The larger semicircle in the intermediate frequency region gives
was added slowly to a stirred solution of 9,9-di-n-octyl-2-bromo-
the information regarding charge transfer dynamics at the dye- fluorene, 164 (3.0 g, 6.4 mmol), in anhydrous tetrahydrofuran (THF, 35
sensitized TiO2−electrolyte interface, and the radius of the mL) at −78 °C, under a nitrogen atmosphere. After stirring for 30 min,
semicircle gives the recombination resistance (Rct). di-tert-butylazodicarboxylate (3.2 g, 14.1 mmol) dissolved in anhydrous
The Rct values of XSQ-sensitized cells do not vary THF (35 mL) was added to the reaction mixture. After stirring at −78
significantly, and they have the approximately same Rct value °C for 2 h, the mixture was allowed to warm to rt and stirred for 18 h.
at −0.50 V applied bias, which could be expected as all of the Glacial acetic acid (1.5 mL) was added slowly, and the mixture was
dyes have similar crowded branched chains, resulting in equal stirred for a further 4 h at rt. Et2O (50 mL) and water (50 mL) were
surface passivation to TiO2. The trend in chemical capacitance added. The aqueous layer was separated and extracted with ether (2 ×
(Cμ) is slightly different, and XSQ1−2 have a higher Cμ 60 mL). The combined organic extracts were then washed with brine
(60 mL), dried over Na2SO4, filtered, and concentrated. The crude
compared to that of XSQ3−4, which ultimately results in the
material was purified by column chromatography on silica (SiO2,
longer lifetime (τ = Rct × Cμ) for XSQ1−2 dyes than for XSQ2− CH2Cl2/petroleum ether as eluents) to give 2 as sticky gum (2.72 g,
4. Generally, the electron lifetime on TiO2 affects the VOC of 68%). 1H NMR (200 MHz, CDCl3) δ: 7.72−7.53 (m, 2H), 7.42−7.25
DSSC and the τ values of XSQ-dye-based cells (without CDCA) (m, 5H), 1.93 (t, J = 8.0 Hz, 4H), 1.50 (s, 18H), 1.08 (d, J = 19.7 Hz,
follow the same trend as VOC, i.e., XSQ3 (9.8 ms) < XSQ4 (10.0 20H), 0.81 (t, J = 6.7 Hz, H), 0.73−0.47 (m, 4H); 13C NMR (100 MHz,
ms) < XSQ1 (13.9 ms) < XSQ2 (14.6 ms) (Table 3). Dyes CDCl3) δ: 155.7, 154.0, 153.9, 151.0, 141.6, 140.7, 138.9, 126.9, 126.9,
122.9, 119.7, 119.6, 82.2, 81.6, 55.2, 40.5, 31.9, 30.2, 29.4, 28.3, 23.9,
Table 3. EIS Parameters of XSQ Dye Cells at −0.5 V 22.7, 14.2; high-resolution mass spectrometry (HRMS) (ESI) m/z: [M
+ H]+ calcd for C39H61O4N2: 621.4626; found: 621.4628.
XSQ dyes Rs (Ω) Rct (Ω) Cμ (mF) τ (ms) 2-(9,9-Dioctyl-9H-fluoren-2-yl)hydrazin-1-ium Chloride (3). A
XSQ1 13.6 6.3 2.2 13.9 solution of concentrated hydrochloric acid (4.1 mL) in 1,4-dioxane
XSQ2 13.9 6.1 2.4 14.6 (12 mL) was added to a solution of 2 (2.8 g, 4.5 mmol) in 2-propanol
XSQ3 12.4 6.1 1.6 9.8
(25 mL). The reaction mixture was heated at 85 °C for 1 h. After
cooling to room temperature, the solvent was completely removed to
XSQ4 13.1 5.9 1.7 10.0
give 3 as pink sticky gum (1.85 g, 90%), which was used directly in the
next step without purification.
3-Decyl-2-methyl-3,9,9-trioctyl-3,9-dihydroindeno[1,2-f ]indole
XSQ1 and XSQ2 have higher dipole moments, which may bring (5). Compound 3 (1.5 g, 3.3 mmol) and 3-octyltridecan-2-one, 465 (2.0
the conduction band slightly upward, allowing a higher g, 6.6 mmol), were taken with 30 mL of acetic acid in a round-bottomed
concentration of electrons on TiO2 in comparison with XSQ2 flask. The resultant mixture was heated for 3 days at 90 °C. After the
and XSQ4 dyes, which leads to a greater electron lifetime in completion of the reaction, the solvents were removed under reduced
XSQ1−2-based DSSCs.69,70 pressure. The residue was dissolved in CH2Cl2 and washed with 5%


KOH solution. The organic layer was collected, dried over Na2SO4, and
CONCLUSIONS purified by column chromatography (SiO2, CH2Cl2/petroleum ether)
to afford 5 as dark yellow oil (0.570 g, 25%). 1H NMR (200 MHz,
The electronic and steric factors of the dyes play a significant CDCl3) δ: 7.68 (d, J = 7.6 Hz, 1H), 7.47 (s, 1H), 7.45 (s, 1H), 7.38−
role in determining the device performance of DSSCs. A series 7.27 (m, 3H), 2.23 (s, 3H), 2.03−1.92 (m, 4H), 1.75 (d, J = 35.2 Hz,
of XSQ dyes were synthesized with a fused fluorenylindolenine- 4H), 1.13 (dd, J = 30.5 Hz, 17.4 Hz, 46H), 0.80 (t, J = 6.6 Hz, 12H),
based donor that not only helps in increasing conjugation but 0.61 (s, 6H); 13C NMR (100 MHz, CDCl3) δ: 186.6, 155.2, 151.3,
also allows appending out-of-plane alkyl chains through sp3-C 151.0, 141.6, 141.4, 138.4, 126.7, 126.6, 122.9, 119.3, 114.3, 112.9, 62.5,
atoms to avoid aggregation. All of the dyes showed good 55., 40.7, 37.4, 32.0, 31.9, 30.2, 30.0, 29.9, 29.7, 29.6, 29.4, 29.4, 29.3,
absorption in the NIR spectral region, with the absorption of 23.9, 23.8, 22.8, 22.7, 16.4, 14.2; matrix-assisted laser desorption
benz[e]indole-based dyes XSQ3−4 more red-shifted than that ionization time-of-flight m/z: [M + H]+ calcd for C50H82N: 696.6447;
of indole-based XSQ1−2. Owing to extensive coverage of found: 696.4977.
3-Decyl-1,2-dimethyl-3,9,9-trioctyl-3,9-dihydroindeno[1,2-f ]-
branched out-of-plane alkyl chains on both sides of the indol-1-ium Iodide (6a). A mixture of 5 (0.15 g, 0.22 mmol) and MeI
molecular plane, aggregation was effectively controlled in all of (0.125 g, 0.86 mmol) in MeCN (5 mL) was heated under a N2
the dyes. DFT-based calculations predicted efficient electron atmosphere at 85 °C for 24 h in a sealed pressure tube. The reaction
transfer toward TiO2 and show higher dipole moments for mixture was cooled to rt before the solvent and excess MeI were
XSQ1−2 in comparison to those for XSQ3−4. XSQ2 showed completely removed under vacuum to give compound 6a (70%, 0.13 g),
the best photovoltaic performance with the PCE of 6.57% in the which was used further without purification.
presence of 3 equivalents of CDCA due to the higher short- 3-Decyl-1-hexyl-2-methyl-3,9,9-trioctyl-3,9-dihydroindeno[1,2-
circuit current density, and EIS analysis indicated a higher f ]indol-1-ium Iodide (6b). A mixture of 5 (0.15 mg, 0.22 mmol) and 1-
iodohexane (0.183 g, 0.86 mmol) in MeCN (5 mL) was heated under
electron lifetime on TiO2 for XSQ2-sensitized solar cell, which nitrogen at 85 °C for 24 h in a sealed pressure tube. The reaction
explains its slightly greater VOC.


mixture was cooled to room temperature before the solvent and excess
1-iodohexane were completely removed in vacuo to give compound 6b
EXPERIMENTAL SECTION (65%, 0.130 g), which was used further without purification.
Precursors 1,64 8,61 and 1045 were synthesized by following a literature (E)-2-((2-Hydroxy-3,4-dioxocyclobut-1-en-1-yl)methylene)-1,3,3-
procedure. Details on the reagents, equipment utilized to characterize trimethylindoline-5-carboxylic Acid (8). In a 50 mL round-bottomed
the synthesized precursors, final dye molecules, further evaluation of the flask, 7 (0.37 g, 1 mmol) was dissolved in acetone (15 mL). HCl (2 N, 2
photophysical, electrochemical, and photovoltaic parameters and mL) was added to the mixture and refluxed for 8 h. The solvent was
electrochemical impedance spectroscopy were provided in the removed under reduced pressure to afford product 8, which was used
Supporting Information. The procedures for the DSSC device without any further purification (0.28 g, 89%). 1H NMR (500 MHz,
fabrication with the mask (active area of the cell 0.36 cm2 and mask dimethyl sulfoxide (DMSO)-d6) δ: 7.81−7.93 (m, 2H), 7.16 (d, J =

26343 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

8.77 Hz, 1H), 5.53 (s, 1H), 3.36 (s, 3H), 1.57 (s, 6H); 13C NMR (100 (ESI) m/z: [M + H]+ calcd for C68H97N2O4: 1005.7448; found:
MHz, DMSO-d6) δ: 193.2, 192.6, 174.0, 167.3, 165.4, 147.3, 140.1, 1005.7446.
130.3, 123.6, 122.8, 107.9, 83.5, 46.3, 29.8, 26.7; HRMS (ESI) m/z: [M (E)-2-(((E)-5-Carboxy-1,3,3-trimethylindolin-2-ylidene)methyl)-4-
+ H]+ calcd for C17H16NO5: 314.1023; found: 314.1012. ((3-decyl-1-hexyl-3,9,9-trioctyl-3,9-dihydroindeno[1,2-f ]indol-1-
7-Carboxy-1,1,2,3-tetramethyl-1H-benzo[e]indol-3-ium Iodide ium-2-yl)methylene)-3-oxocyclobut-1-en-1-olate (XSQ2). Started
(10). Compound 945 (1 g, 3.95 mmol), MeI (2.8 g, 19.75 mmol), with 6b (0.2 g, 0.22 mmol) and 8 (0.104 g, 0.330 mmol). Yield: 0.15
and MeCN (8 mL) were taken in a sealed tube and heated at 150 °C for g, 60%. 1H NMR (500 MHz, CDCl3) δ: 8.13 (d, J = 8.4 Hz, 1H), 8.07
48 h. After the completion of the reaction, the contents were poured (s, 1H), 7.73 (d, J = 7.4 Hz, 1H), 7.59 (s, 1H), 7.34 (td, J = 14.5, 7.1 Hz,
into 50 mL of Et2O and the precipitate obtained was filtered. The 3H), 6.96 (d, J = 7.7 Hz, 2H), 6.22 (s, 1H), 6.00 (s, 1H), 4.19 (s, 2H),
residue was washed with diethyl ether (3 × 20 mL) and dried under 3.52 (s, 3H), 3.11 (s, 2H), 2.04 (dd, J = 17.7 Hz, 9.7 Hz, 4H), 1.99−
vacuum to give compound 10 (0.78 g, 50%). 1H NMR (200 MHz, 1.91 (m, 2H), 1.86 (s, 6H), 1.47−1.38 (m, 2H), 1.35−1.24 (m, 6H),
DMSO-d6) δ: 8.87 (s, 1H), 8.52 (d, J = 6.3 Hz, 1H), 8.47 (d, J = 6.1 Hz, 1.20−1.15 (m, 6H), 1.12−1.01 (m, 37H), 0.86 (t, J = 7.0 Hz, 3H),
1H), 8.26−8.13 (m, 2H), 4.11 (s, 3H), 2.90 (s, 3H), 1.76 (s, 6H); 13C 0.83−0.74 (m, 15H), 0.70−0.45 (m, 8H); 13C NMR (125 MHz,
NMR (100 MHz, DMSO-d6) δ: 197.3, 167.1, 141.3, 136.5, 132.4, CDCl3) δ: 181.8, 175.3,171.3, 170.7, 168.10, 151.6, 150.76, 147.78,
143.8, 147.7, 143.8, 141.9, 140.6, 139.4, 138.7, 131.3, 127.1, 124.1,
132.3, 132.2, 129.1, 129.0, 127.5, 124.0, 114.9, 55.5, 35.3, 21.3, 14.2;
124.1 122.9, 119.67, 114.2, 113.87, 108.0, 104.7, 88.2, 59.32, 59.2, 55.3,
HRMS (ESI) m/z: [M]+ calcd for C17H18NO2+: 268.1332; found:
48.2, 40.7, 40.4, 32.0, 31.93, 31.90, 31.78, 31.7, 30.5, 30.1, 29.7, 29.6,
268.1325.
29.4, 29.3, 29.2, 27.8, 27.5, 27.1, 24.2, 23.9, 22.7, 22.67, 22.65, 14.2,
(E)-2-((2-Butoxy-3,4-dioxocyclobut-1-en-1-yl)methylene)-1,1,3-
14.1; HRMS (ESI) m/z: [M + H]+ calcd for C73H107N2O4: 1075.8231;
trimethyl-2,3-dihydro-1H-benzo[e]indole-7-carboxylic Acid (11). To
found: 1075.8220.
a mixture of 10 (0.5 g, 1.26 mmol) in 20 mL of n-BuOH, 3,4-
(E)-2-((E)-(7-Carboxy-1,1,3-trimethyl-1,3-dihydro-2H-benzo[e]-
dibutoxycyclobut-3-ene-1,2-dione (0.72 g, 3.16 mmol) was added. To indol-2-ylidene)methyl)-4-((3-decyl-1-methyl-3,9,9-trioctyl-3,9-
the stirring mixture, NEt3 (0.38 g, 3.78 mmol) was added dropwise. The dihydroindeno[1,2-f ]indol-1-ium-2-yl)methylene)-3-oxocyclobut-
resultant mixture was stirred at room temperature for 12 h followed by 1-en-1-olate (XSQ3). Started with 6a (0.180 g, 0.22 mmol) and 12
heating at 70 °C for 3 h. Solvents were evaporated after the completion (0.156 g, 0.43 mmol). Yield: 0.11 g, 48%. 1H NMR (500 MHz, CDCl3)
of reaction, and the crude product was purified by column δ: 8.74 (s, 1H), 8.36−8.12 (m, 2H), 7.99 (s, 1H), 7.72 (d, J = 7.3 Hz,
chromatography (SiO2, MeOH/CH2Cl2) to afford 11 (0.37 g, 70%). 1H), 7.57 (s, 1H), 7.39−7.27 (m, 4H), 6.95 (s, 1H), 6.19 (s, 1H), 6.05
1
H NMR (200 MHz, DMSO-d6) δ: 8.63 (s, 1H), 8.23 (d, J = 9.1 Hz, (s, 1H), 3.64 (s, 5H), 3.11 (s, 1H), 2.22−2.06 (m, 6H), 2.06−1.89 (m,
1H), 8.17 (d, J = 9.1 Hz, 1H), 8.00 (dd, J = 8.9 Hz, 1.6 Hz, 1H), 7.69 (d, 6H), 1.24−0.99 (m, 46H), 0.87−0.71 (m, 14H), 0.67−0.51 (m, 6H);
J = 8.9 Hz, 1H), 5.39 (s, 1H), 4.82 (t, J = 6.6 Hz, 2H), 3.52 (s, 3H),
13
C NMR (125 MHz, CDCl3) δ: 179.8, 176.8, 170.8, 170.3, 151.7,
1.95−1.70 (m, 8H), 1.46 (dq, J = 14.4, 7.2 Hz, 2H), 0.96 (t, J = 7.3 Hz, 150.6, 144.2, 142.6, 140.8, 138.2, 134.0, 133.7, 131.7, 131.1, 130.1,
3H); 13C NMR (125 MHz, DMSO-d6) δ: 192.3, 188.1, 186.9, 172.1, 127.1, 126.9, 125.2, 122.9, 122.7, 119.5, 113.7, 110.8, 103.9, 88.6, 87.5,
170.0, 167.4, 142.6, 132.5, 131.5, 131.3, 129.8, 129.4, 126.6, 125.3, 58.9, 55.3, 50.7, 40.7, 40.4, 32.0, 31.9, 30.1, 29.8, 29.74, 29.70, 29.6,
122.3, 111.7, 81.4, 73.6, 48.9, 31.6, 30.3, 26.2, 18.2, 13.6; HRMS (ESI) 29.42, 29.39, 29.35, 29.3, 29.2, 27.1, 24.3, 23.8, 22.7, 22.69, 14.2;
m/z: [M + H]+ calcd for C25H26O5N: 420.1805; found: 420.1794. HRMS (ESI) m/z: [M + H]+ calcd for C72H99N2O4: 1055.7605; found:
(E)-2-((2-Hydroxy-3,4-dioxocyclobut-1-en-1-yl)methylene)-1,1,3- 1055.7596.
trimethyl-2,3-dihydro-1H-benzo[e]indole-7-carboxylic Acid (12). To (E)-2-((E)-(7-Carboxy-1,1,3-trimethyl-1,3-dihydro-2H-benzo[e]-
a solution of 11 (0.35 g, 0.834 mmol) in 10 mL of acetone, 1 mL of 2 N indol-2-ylidene)methyl)-4-((3-decyl-1-hexyl-3,9,9-trioctyl-3,9-
HCl was added. The resultant mixture was refluxed for 8 h, and solvents dihydroindeno[1,2-f ]indol-1-ium-2-yl)methylene)-3-oxocyclobut-
1-en-1-olate (XSQ4). Started with 6b (0.18 g, 0.22 mmol) and 12
were removed under reduced pressure after the completion of reaction.
(0.156 g, 0.43 mmol). Yield: 0.100 g, 54%. 1H NMR (500 MHz,
The crude compound 12 (0.286 g, 94%) was used further without
CDCl3) δ: 8.77 (s, 1H), 8.35−8.15 (m, 2H), 8.01 (d, J = 6.1 Hz, 1H),
purification. 1H NMR (200 MHz, DMSO-d6) δ: 8.60 (s, 1H), 8.21 (d, J 7.75 (d, J = 7.4 Hz, 1H), 7.60 (s, 1H), 7.40−7.29 (m, 4H), 6.96 (s, 1H),
= 8.8 Hz, 1H), 8.12 (d, J = 8.9 Hz, 1H), 7.97 (d, J = 8.5 Hz, 1H), 7.62 (d, 6.24 (s, 1H), 6.07 (s, 1H), 4.17 (s, 2H), 3.66 (s, 3H), 3.16 (s, 2H), 2.14
J = 8.8 Hz, 1H), 5.56 (s, 1H), 3.47 (s, 3H), 1.83 (s, 6H); 13C NMR (100 (s, 6H), 2.10−2.01 (m, 4H), 2.01−1.94 (m, 2H), 1.92−1.85 (m, 2H),
MHz, DMSO-d6) δ: 192.7, 192.1, 191.6, 173.6, 167.7, 167.4, 142.9, 1.49−1.43 (m, 2H), 1.36−1.27 (m, 4H), 1.23−1.18 (m, 6H), 1.16−
132.5, 131.3, 130.7, 129.9, 129.0, 126.4, 124.9, 122.1, 111.4, 82.3, 48.4, 1.05 (m, 37H), 0.88 (t, J = 7.0 Hz, 3H), 0.85−0.75 (m, 15H), 0.72−
30.0, 26.3; HRMS (ESI) m/z: [M + H]+ calcd for C21H18O5N: 0.48 (m, 8H); 13C NMR (125 MHz, CDCl3) δ: 179.1, 175.4, 170.56,
364.1179; found: 364.1170. 170.13, 170.0, 151.5, 150.6, 144.0, 142.6, 140.8, 139.0, 139.0, 138.2,
General Procedure for the Synthesis of Unsymmetrical 135.4, 133.9, 131.6, 131.1, 130.1, 127.1, 126.9, 125.5, 123.5, 122.8,
Squaraine Dyes, XSQ1−4. 6a or 6b (1 equiv) and semisquaric acid 122.6, 119.5, 114.2, 113.7, 110.7, 104.4, 88.5, 87.3, 58.9, 55.2, 50.6,
derivatives 8 or 12 (2 equiv) were taken in a 100 mL round-bottomed 40.7, 32.0, 31.96, 31.9, 31.8, 30.1, 29.7, 29.6, 29.5, 29.4, 29.3, 29.2, 27.7,
flask fitted with a Dean−Stark apparatus. n-BuOH/PhMe (20 mL, 1:1) 27.2, 24.3, 23.9, 22.7, 22.7, 22.4, 14.2, 14.1; HRMS (ESI) m/z: [M +
was added, and the contents were refluxed for 24 h. After completion of H]+ calcd for C77H109N2O4: 1125.8387; found: 1125.8378.


reaction, solvents were removed under reduced pressure and the
residue was purified by column chromatography (MeOH/CH2Cl2) to
ASSOCIATED CONTENT
give pure compound XSQ dyes. For the synthesis of XSQ3−4, 2 mL of
DMSO was added along with 20 mL of n-BuOH/PhMe (1:1). *
S Supporting Information
(E)-2-(((E)-5-Carboxy-1,3,3-trimethylindolin-2-ylidene)methyl)-4- The Supporting Information is available free of charge on the
((3-decyl-1-methyl-3,9,9-trioctyl-3,9-dihydroindeno[1,2-f ]indol-1- ACS Publications website at DOI: 10.1021/acsami.8b09866.
ium-2-yl)methylene)-3-oxocyclobut-1-en-1-olate (XSQ1). Started
with 6a (0.2 g, 0.24 mmol) and 8 (0.156 g, 0.48 mmol). Yield: 0.165 J−V curve, IPCE trace, Nyquist plots, photovoltaic
g, 68%. 1H NMR (500 MHz, CDCl3) δ: 8.12 (d, J = 8.3 Hz, 1H), 8.06 parameters, and EIS data of XSQ1−4-based dye cells,
(s, 1H), 7.73 (d, J = 7.4 Hz, 1H), 7.58 (s, 1H), 7.38−7.28 (m, 3H),
7.03−6.92 (m, 2H), 6.18 (s, 1H), 5.99 (s, 1H), 3.71 (s, 3H), 3.52 (s, dye desorption curves, materials and methods, synthetic
3H), 2.09−2.00 (m, 4H), 1.98−1.92 (m, 2H), 1.85 (s, 6H), 1.21−1.16 procedure, 1H and 13C NMR spectra (PDF)


(m, 6H), 1.14−1.01 (m, 40H), 0.83−0.74 (m, 14H), 0.63−0.51 (m,
6H); 13C NMR (125 MHz, CDCl3) δ: 182.1, 170.6, 151.9, 150.7, 147.7, AUTHOR INFORMATION
144.1, 142.0, 140.6, 139.1, 138.8, 131.3, 127.1, 124.1, 123.0, 119.7,
113.8, 108.0, 88.3, 59.3, 55.4, 48.2, 40.7, 32.0, 31.9, 30.6, 30.1, 29.7, Corresponding Author
29.6, 29.39, 29.35, 29.3, 29.2, 27.5, 24.2, 23.8, 22.7, 22.7, 14.2; HRMS *E-mail: j.nithyanandhan@ncl.res.in.
26344 DOI: 10.1021/acsami.8b09866
ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

ORCID ing Light-Harvesting and Retarding Charge Recombination. J. Mater.


Rajesh Bisht: 0000-0003-0150-9696 Chem. A 2014, 2, 14649−14657.
Jayaraj Nithyanandhan: 0000-0002-3429-4989 (17) Zhu, H.; Wu, Y.; Liu, J.; Zhang, W.; Wu, W.; Zhu, W.-H. D−
A−π−A Featured Sensitizers Containing an Auxiliary Acceptor of
Notes Benzoxadiazole: Molecular Engineering and Co-Sensitization. J. Mater.
The authors declare no competing financial interest.


Chem. A 2015, 3, 10603−10609.
(18) Pei, K.; Wu, Y.; Wu, W.; Zhang, Q.; Chen, B.; Tian, H.; Zhu, W.
ACKNOWLEDGMENTS Constructing Organic D−A−π−A−Featured Sensitizers with a
Funding from NWP0054 (CSIR-TAPSUN) and SERB-EMR/ Quinoxaline Unit for High−Efficiency Solar Cells: The Effect of an
2016/007114, India, is sincerely acknowledged. R.B. and V.S. Auxiliary Acceptor on the Absorption and the Energy Level Alignment.
thank CSIR, New Delhi, India, and N.K. thanks UGC, New Chem. − Eur. J. 2012, 18, 8190−8200.
(19) Zhu, H.; Liu, B.; Liu, J.; Zhang, W.; Zhu, W.-H. D−A−π−A
Delhi, India, for research fellowships. J.N. thanks Dr.
Featured Sensitizers by Modification of Auxiliary Acceptor for
Kothandam Krishnamoorthy, PSE Division, CSIR-NCL, Pune,
Preventing “Trade-off” Effect. J. Mater. Chem. C 2015, 3, 6882−6890.
for his help with device fabrication.


(20) Yum, J.-H.; Holcombe, T. W.; Kim, Y.; Rakstys, K.; Moehl, T.;
Teuscher, J.; Delcamp, J. H.; Nazeeruddin, M. K.; Grätzel, M. Blue-
REFERENCES Coloured Highly Efficient Dye-Sensitized Solar Cells by Implementing
(1) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Dye- the Diketopyrrolopyrrole Chromophore. Sci. Rep. 2013, 3, No. 2446.
Sensitized Solar Cells. Chem. Rev. 2010, 110, 6595−6663. (21) Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.; Hanaya,
(2) Grätzel, M. Solar Energy Conversion by Dye-Sensitized M. Highly-Efficient Dye-Sensitized Solar Cells with Collaborative
Photovoltaic Cells. Inorg. Chem. 2005, 44, 6841−6851. Sensitization by Silyl-Anchor and Carboxy-Anchor Dyes. Chem.
(3) Hardin, B. E.; Snaith, H. J.; McGehee, M. D. The Renaissance of Commun. 2015, 51, 15894−15897.
Dye-Sensitized Solar Cells. Nat. Photonics 2012, 6, 162−169. (22) Zhang, S.; Yang, X.; Numata, Y.; Han, L. Highly Efficient Dye-
(4) Chen, C.-Y.; Wang, M.; Li, J.-Y.; Pootrakulchote, N.; Alibabaei, L.; Sensitized Solar Cells: Progress and Future Challenges. Energy Environ.
Ngoc-le, C.; Decoppet, J.-D.; Tsai, J.-H.; Grätzel, C.; Wu, C.-G.; Sci. 2013, 6, 1443−1464.
Zakeeruddin, S. M.; Grätzel, M. Highly Efficient Light-Harvesting (23) Brogdon, P.; Cheema, H.; Delcamp, J. H. Near-Infrared-
Ruthenium Sensitizer for Thin-Film Dye-Sensitized Solar Cells. ACS Absorbing Metal-Free Organic, Porphyrin, and Phthalocyanine
Nano 2009, 3, 3103−3109.
Sensitizers for Panchromatic Dye-Sensitized Solar Cells. ChemSusChem
(5) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B. F.
2018, 11, 86−103.
E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, M.
(24) Martín-Gomis, L.; Fernández-Lázaro, F.; Sastre-Santos, Á
K.; Grätzel, M. Dye-Sensitized Solar Cells with 13% Efficiency
Achieved through the Molecular Engineering of Porphyrin Sensitizers. Advances in Phthalocyanine-Sensitized Solar Cells (PcSSCs). J.
Nat. Chem. 2014, 6, 242−247. Mater. Chem. A 2014, 2, 15672−15682.
(6) Mishra, A.; Fischer, M. K. R.; Bäuerle, P. Metal-Free Organic Dyes (25) Li, L.-L.; Diau, E. W.-G. Porphyrin-Sensitized Solar Cells. Chem.
for Dye-Sensitized Solar Cells: From Structure: Property Relationships Soc. Rev. 2013, 42, 291−304.
to Design Rules. Angew. Chem., Int. Ed. 2009, 48, 2474−2499. (26) Yella, A.; Lee, H.-W.; Tsao, H. N.; Yi, C.; Chandiran, A. K.;
(7) Liang, M.; Chen, J. Arylamine Organic Dyes for Dye-Sensitized Nazeeruddin, M. K.; Diau, E. W.-G.; Yeh, C.-Y.; Zakeeruddin, S. M.;
Solar Cells. Chem. Soc. Rev. 2013, 42, 3453−3488. Gratzel, M. Porphyrin-Sensitized Solar Cells with Cobalt (II/III)-Based
(8) Mahmood, A. Triphenylamine Based Dyes for Dye Sensitized Redox Electrolyte Exceed 12 Percent Efficiency. Science 2011, 334,
Solar Cells: A Review. Sol. Energy 2016, 123, 127−144. 629−634.
(9) Wu, Y.; Zhu, W. Organic Sensitizers from D−π−A to D−A−π−A: (27) Bessho, T.; Zakeeruddin, S. M.; Yeh, C.-Y.; Diau, E. W.-G.;
Effect of the Internal Electron-Withdrawing Units on Molecular Grätzel, M. Highly Efficient Mesoscopic Dye-Sensitized Solar Cells
Absorption, Energy Levels and Photovoltaic Performances. Chem. Soc. Based on Donor-Acceptor-Substituted Porphyrins. Angew. Chem., Int.
Rev. 2013, 42, 2039−2058. Ed. 2010, 49, 6646−6649.
(10) Velusamy, M.; Thomas, K. R. J.; Lin, J. T.; Hsu, Y.-C.; Ho, K.-C. (28) Urbani, M.; Grätzel, M.; Nazeeruddin, M. K.; Torres, T. Meso-
Organic Dyes Incorporating Low-Band-Gap Chromophores for Dye- Substituted Porphyrins for Dye-Sensitized Solar Cells. Chem. Rev. 2014,
Sensitized Solar Cells. Org. Lett. 2005, 7, 1899−1902. 114, 12330−12396.
(11) Wu, Y.; Zhang, X.; Li, W.; Wang, Z.-S.; Tian, H.; Zhu, W. (29) Kang, S. H.; Jeong, M. J.; Eom, Y. K.; Choi, I. T.; Kwon, S. M.;
Hexylthiophene-Featured D−A−π−A Structural Indoline Chromo- Yoo, Y.; Kim, J.; Kwon, J.; Park, J. H.; Kim, H. K. Porphyrin Sensitizers
phores for Coadsorbent-Free and Panchromatic Dye-Sensitized Solar with Donor Structural Engineering for Superior Performance Dye-
Cells. Adv. Energy Mater. 2012, 2, 149−156. Sensitized Solar Cells and Tandem Solar Cells for Water Splitting
(12) Qu, S.; Wu, W.; Hua, J.; Kong, C.; Long, Y.; Tian, H. New Applications. Adv. Energy Mater. 2017, 7, No. 1602117.
Diketopyrrolopyrrole (DPP) Dyes for Efficient Dye-Sensitized Solar (30) Saccone, D.; Galliano, S.; Barbero, N.; Quagliotto, P.; Viscardi,
Cells. J. Phys. Chem. C 2010, 114, 1343−1349.
G.; Barolo, C. Polymethine Dyes in Hybrid Photovoltaics: Structure−
(13) Qu, S.; Qin, C.; Islam, A.; Wu, Y.; Zhu, W.; Hua, J.; Tian, H.;
Properties Relationships. Eur. J. Org. Chem. 2016, 2016, 2244−2259.
Han, L. A Novel D−A−π−A Organic Sensitizer Containing a
(31) Chen, G.; Sasabe, H.; Igarashi, T.; Hong, Z.; Kido, J. Squaraine
Diketopyrrolopyrrole Unit with a Branched Alkyl Chain for Highly
Efficient and Stable Dye-Sensitized Solar Cells. Chem. Commun. 2012, Dyes for Organic Photovoltaic Cells. J. Mater. Chem. A 2015, 3, 14517−
48, 6972−6974. 14534.
(14) Cui, Y.; Wu, Y.; Lu, X.; Zhang, X.; Zhou, G.; Miapeh, F. B.; Zhu, (32) Qin, C.; Wong, W.-Y.; Han, L. Squaraine Dyes for Dye-Sensitized
W.; Wang, Z.-S. Incorporating Benzotriazole Moiety to Construct D− Solar Cells: Recent Advances and Future Challenges. Chem. − Asian J.
A−π−A Organic Sensitizers for Solar Cells: Significant Enhancement 2013, 8, 1706−1719.
of Open-Circuit Photovoltage with Long Alkyl Group. Chem. Mater. (33) Yum, J.-H.; Walter, P.; Huber, S.; Rentsch, D.; Geiger, T.;
2011, 23, 4394−4401. Nüesch, F.; De Angelis, F.; Grätzel, M.; Nazeeruddin, M. K. Efficient
(15) Ni, J.-S.; Yen, Y.-C.; Lin, J. T. Organic Sensitizers with a Rigid Far Red Sensitization of Nanocrystalline TiO2 Films by an Unsym-
Dithienobenzotriazole-Based Spacer for High-Performance Dye- metrical Squaraine Dye. J. Am. Chem. Soc. 2007, 129, 10320−10321.
Sensitized Solar Cells. J. Mater. Chem. A 2016, 4, 6553−6560. (34) Alex, S.; Santhosh, U.; Das, S. Dye Sensitization of Nanocrystal-
(16) Li, H.; Wu, Y.; Geng, Z.; Liu, J.; Xu, D.; Zhu, W. Co-Sensitization line TiO2: Enhanced Efficiency of Unsymmetrical versus Symmetrical
of Benzoxadiazole Based D−A−π−A Featured Sensitizers: Compensat- Squaraine Dyes. J. Photochem. Photobiol., A 2005, 172, 63−71.

26345 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

(35) Otsuka, A.; Funabiki, K.; Sugiyama, N.; Yoshida, T.; Minoura, Toward Efficient Panchromatic Multi-Chromophoric Sensitizers for
H.; Matsui, M. Dye Sensitization of ZnO by Unsymmetrical Squaraine Dye Sensitized Solar Cells. Chem. Mater. 2015, 27, 6305−6313.
Dyes Suppressing Aggregation. Chem. Lett. 2006, 35, 666−667. (53) Shi, Y.; Hill, R. B. M.; Yum, J.-H.; Dualeh, A.; Barlow, S.; Grätzel,
(36) Kuster, S.; Sauvage, F.; Nazeeruddin, M. K.; Grätzel, M.; Nüesch, M.; Marder, S. R.; Nazeeruddin, M. K. A High-Efficiency Panchromatic
F. A.; Geiger, T. Unsymmetrical Squaraine Dimer with an Extended π- Squaraine Sensitizer for Dye-Sensitized Solar Cells. Angew. Chem., Int.
Electron Framework: An Approach in Harvesting near Infra-Red Ed. 2011, 50, 6619−6621.
Photons for Energy Conversion. Dyes Pigm. 2010, 87, 30−38. (54) Delcamp, J. H.; Shi, Y.; Yum, J.-H.; Sajoto, T.; Dell’Orto, E.;
(37) Li, C.; Wang, W.; Wang, X.; Zhang, B.; Cao, Y. Molecular Design Barlow, S.; Nazeeruddin, M. K.; Marder, S. R.; Grätzel, M. The Role of π
of Squaraine Dyes for Efficient Far-Red and Near-IR Sensitization of Bridges in High-Efficiency DSCs Based on Unsymmetrical Squaraines.
Solar Cells. Chem. Lett. 2005, 34, 554−555. Chem. − Eur. J. 2013, 19, 1819−1827.
(38) Mann, J. R.; Gannon, M. K.; Fitzgibbons, T. C.; Detty, M. R.; (55) Bae, S. H.; Seo, K. D.; Choi, W. S.; Hong, J. Y.; Kim, H. K. Near-
Watson, D. F. Optimizing the Photocurrent Efficiency of Dye- IR Organic Sensitizers Containing Squaraine and Phenothiazine Units
Sensitized Solar Cells through the Controlled Aggregation of for Dye-Sensitized Solar Cells. Dyes Pigm. 2015, 113, 18−26.
Chalcogenoxanthylium Dyes on Nanocrystalline Titania Films. J. (56) Bisht, R.; Munavvar, M. M. K.; Singh, A. K.; Nithyanandhan, J.
Phys. Chem. C 2008, 112, 13057−13061. Panchromatic Sensitizer for Dye-Sensitized Solar Cells: Unsymmetrical
(39) de Miguel, G.; Ziółek, M.; Zitnan, M.; Organero, J. A.; Pandey, S. Squaraine Dyes Incorporating Benzodithiophene π-Spacer with Alkyl
S.; Hayase, S.; Douhal, A. Photophysics of H- and J-Aggregates of Chains to Extend Conjugation, Control the Dye Assembly on TiO2,
Indole-Based Squaraines in Solid State. J. Phys. Chem. C 2012, 116, and Retard Charge Recombination. J. Org. Chem. 2017, 82, 1920−
9379−9389. 1930.
(40) Mulhern, K. R.; Detty, M. R.; Watson, D. F. Aggregation-Induced (57) Alagumalai, A.; Munavvar, M. M. K.; Vellimalai, P.; Sil, M. C.;
Increase of the Quantum Yield of Electron Injection from Nithyanandhan, J. Effect of Out-of-Plane Alkyl Group’s Position in
Chalcogenorhodamine Dyes to TiO2. J. Phys. Chem. C 2011, 115, Dye-Sensitized Solar Cell Efficiency: A Structure−Property Relation-
6010−6018. ship Utilizing Indoline-Based Unsymmetrical Squaraine Dyes. ACS
(41) Mulhern, K. R.; Detty, M. R.; Watson, D. F. Effects of Surface- Appl. Mater. Interfaces 2016, 8, 35353−35367.
Anchoring Mode and Aggregation State on Electron Injection from (58) Tang, Y.; Wang, Y.; Li, X.; Ågren, H.; Zhu, W.-H.; Xie, Y.
Chalcogenorhodamine Dyes to Titanium Dioxide. J. Photochem. Porphyrins Containing a Triphenylamine Donor and up to Eight
Photobiol., A 2013, 264, 18−25. Alkoxy Chains for Dye-Sensitized Solar Cells: A High Efficiency of
(42) Kryman, M. W.; Nasca, J. N.; Watson, D. F.; Detty, M. R. 10.9%. ACS Appl. Mater. Interfaces 2015, 7, 27976−27985.
Selenorhodamine Dye-Sensitized Solar Cells: Influence of Structure (59) Yang, G.; Tang, Y.; Li, X.; Ågren, H.; Xie, Y. Efficient Solar Cells
and Surface-Anchoring Mode on Aggregation, Persistence, and Based on Porphyrin Dyes with Flexible Chains Attached to the
Photoelectrochemical Performance. Langmuir 2016, 32, 1521−1532. Auxiliary Benzothiadiazole Acceptor: Suppression of Dye Aggregation
(43) Rao, G. H.; Venkateswararao, A.; Giribabu, L.; Singh, S. P. Near- and the Effect of Distortion. ACS Appl. Mater. Interfaces 2017, 9,
Infrared Unsymmetrical Blue and Green Squaraine Sensitizers. 36875−36885.
Photochem. Photobiol. Sci. 2016, 15, 287−296. (60) Song, H.; Liu, Q.; Xie, Y. Porphyrin-Sensitized Solar Cells:
(44) Geiger, T.; Kuster, S.; Yum, J.-H.; Moon, S.-J.; Nazeeruddin, M. Systematic Molecular Optimization, Coadsorption and Cosensitiza-
K.; Grätzel, M.; Nüesch, F. Molecular Design of Unsymmetrical tion. Chem. Commun. 2018, 54, 1811−1824.
Squaraine Dyes for High Efficiency Conversion of Low Energy Photons (61) Mori, S.; Nagata, M.; Nakahata, Y.; Yasuta, K.; Goto, R.; Kimura,
into Electrons Using TiO2 Nanocrystalline Films. Adv. Funct. Mater. M.; Taya, M. Enhancement of Incident Photon-to-Current Conversion
2009, 19, 2720−2727. Efficiency for Phthalocyanine-Sensitized Solar Cells by 3D Molecular
(45) Park, J.; Barbero, N.; Yoon, J.; Dell’Orto, E.; Galliano, S.; Borrelli, Structuralization. J. Am. Chem. Soc. 2010, 132, 4054−4055.
R.; Yum, J.-H.; Censo, D. D.; Grätzel, M.; K. Nazeeruddin, M.; Barolo, (62) Ikeuchi, T.; Nomoto, H.; Masaki, N.; Griffith, M. J.; Mori, S.;
C.; Viscardi, G. Panchromatic Symmetrical Squaraines: A Step Forward Kimura, M. Molecular Engineering of Zinc Phthalocyanine Sensitizers
in the Molecular Engineering of Low Cost Blue-Greenish Sensitizers for for Efficient Dye-Sensitized Solar Cells. Chem. Commun. 2014, 50,
Dye-Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2014, 16, 24173− 1941−1943.
24177. (63) Matsuzaki, H.; Murakami, T. N.; Masaki, N.; Furube, A.; Kimura,
(46) Maeda, T.; Nitta, S.; Sano, Y.; Tanaka, S.; Yagi, S.; Nakazumi, H. M.; Mori, S. Dye Aggregation Effect on Interfacial Electron-Transfer
Near-Infrared Squaraine Sensitizers Bearing Benzo[c,d]Indolenine as Dynamics in Zinc Phthalocyanine-Sensitized Solar Cells. J. Phys. Chem.
an Acceptor Moiety. Dyes Pigm. 2015, 122, 160−167. C 2014, 118, 17205−17212.
(47) Maeda, T.; Shima, N.; Tsukamoto, T.; Yagi, S.; Nakazumi, H. (64) Liu, S.-J.; Zhao, Q.; Deng, Y.; Xia, Y.-J.; Lin, J.; Fan, Q.-L.; Wang,
Unsymmetrical Squarylium Dyes with π-Extended Heterocyclic L.-H.; Huang, W. π-Conjugated Chelating Polymers with a Charged
Components and Their Application to Organic Dye-Sensitized Solar Iridium Complex in the Backbones: Toward Saturated-Red Phosphor-
Cells. Synth. Met. 2011, 161, 2481−2487. escent Polymer Light-Emitting Diodes. J. Phys. Chem. C 2007, 111,
(48) Pandey, S. S.; Watanabe, R.; Fujikawa, N.; Shivashimpi, G. M.; 1166−1175.
Ogomi, Y.; Yamaguchi, Y.; Hayase, S. Effect of Extended π-Conjugation (65) Karjule, N.; MK, M. F.; Nithyanandhan, J. Heterotriangulene-
on Photovoltaic Performance of Dye Sensitized Solar Cells Based on Based Unsymmetrical Squaraine Dyes: Synergistic Effects of Donor
Unsymmetrical Squaraine Dyes. Tetrahedron 2013, 69, 2633−2639. Moieties and out-of-Plane Branched Alkyl Chains on Dye Cell
(49) Yan, Z.; Guang, S.; Su, X.; Xu, H. Near-Infrared Absorbing Performance. J. Mater. Chem. A 2016, 4, 18910−18921.
Squaraine Dyes for Solar Cells: Relationship between Architecture and (66) Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Ohga, Y.; Shinpo, A.;
Performance. J. Phys. Chem. C 2012, 116, 8894−8900. Suga, S.; Sayama, K.; Sugihara, H.; Arakawa, H. Molecular Design of
(50) Li, J.-Y.; Chen, C.-Y.; Ho, W.-C.; Chen, S.-H.; Wu, C.-G. Coumarin Dyes for Efficient Dye-Sensitized Solar Cells. J. Phys. Chem. B
Unsymmetrical Squaraines Incorporating Quinoline for Near Infrared 2003, 107, 597−606.
Responsive Dye-Sensitized Solar Cells. Org. Lett. 2012, 14, 5420−5423. (67) Wenger, S.; Bouit, P.-A.; Chen, Q.; Teuscher, J.; Censo, D. D.;
(51) Jradi, F. M.; Kang, X.; O’Neil, D.; Pajares, G.; Getmanenko, Y. A.; Humphry-Baker, R.; Moser, J.-E.; Delgado, J. L.; Martín, N.;
Szymanski, P.; Parker, T. C.; El-Sayed, M. A.; Marder, S. R. Near- Zakeeruddin, S. M.; Grätzel, M. Efficient Electron Transfer and
Infrared Asymmetrical Squaraine Sensitizers for Highly Efficient Dye Sensitizer Regeneration in Stable π-Extended Tetrathiafulvalene-
Sensitized Solar Cells: The Effect of π-Bridges and Anchoring Groups Sensitized Solar Cells. J. Am. Chem. Soc. 2010, 132, 5164−5169.
on Solar Cell Performance. Chem. Mater. 2015, 27, 2480−2487. (68) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
(52) Jradi, F. M.; O’Neil, D.; Kang, X.; Wong, J.; Szymanski, P.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.;
Parker, T. C.; Anderson, H. L.; El-Sayed, M. A.; Marder, S. R. A Step Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.;

26346 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347
ACS Applied Materials & Interfaces Research Article

Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.;
Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.;
Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.;
Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi,
J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.;
Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;
Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.;
Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador,
P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .;
Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09;
Gaussian, Inc.: Wallingford, CT, 2009.
(69) Buhbut, S.; Clifford, J. N.; Kosa, M.; Anderson, A. Y.; Shalom, M.;
Major, D. T.; Palomares, E.; Zaban, A. Controlling Dye Aggregation,
Injection Energetics and Catalytic Recombination in Organic Sensitizer
Based Dye Cells Using a Single Electrolyte Additive. Energy Environ. Sci.
2013, 6, 3046−3053.
(70) Liang, Y.; Cheng, F.; Liang, J.; Chen, J. Triphenylamine-Based
Ionic Dyes with Simple Structures: Broad Photoresponse and
Limitations on Open-Circuit Voltage in Dye-Sensitized Solar Cells. J.
Phys. Chem. C 2010, 114, 15842−15848.

26347 DOI: 10.1021/acsami.8b09866


ACS Appl. Mater. Interfaces 2018, 10, 26335−26347

You might also like