Accepted Article: Title: Indenoquinaldine Based Unsymmetrical Squaraine Dyes For
Accepted Article: Title: Indenoquinaldine Based Unsymmetrical Squaraine Dyes For
Accepted Article
This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.
Supported by
Chemistry - A European Journal 10.1002/chem.201803062
FULL PAPER
Abstract: A series of near-infrared (NIR) responsive unsymmetrical been explored. Ruthenium(II)polypyridyl complexes are among
squaraine dyes (ISQ1-3), incorporating a fused indenoquinaldine the most studied sensitizers, and PCE up to 11.5% was
Accepted Manuscript
based donor, were designed and synthesized. The sp3-C center achieved[3c] whereas the DSSC based upon a panchromatic
available on the indene unit of indenoquinaldine was employed to porphyrin sensitizer, containing a benzothiadiazole (BTD) moiety,
incorporate twelve-carbon alkyl chains in an out-of-plane direction to yielded PCE up to 13%.[4b] Metal-free organic dyes have unique
control dye aggregation on titanium dioxide (TiO2) surface. Indole advantages over metal-organic complex based sensitizers in
(ISQ1), benz[e]indole (ISQ2) and quinoline (ISQ3) moieties were terms of low cost, non-toxicity, flexibility in molecular designing
included towards anchoring group to extend the absorption in NIR and tunability in electronic properties. [5a] Several metal-free
region and systematically study the effect of the electronic sensitizers with donor--acceptor (D--A) and donor-acceptor--
modification on DSSC performance. All the dyes exhibited intense acceptor (D-A--A) configuration have been explored.[6] They
absorption ( ≥ 105 M-1cm-1) in the NIR region, and dye-adsorbed show high molar extinction coefficients due to intramolecular
TiO2 films exhibited broad panchromatic absorption. The incident charge transfer (ICT) which occurs within the molecule because
photon-to-current efficiency (IPCE) spectrum of the ISQ3 based of the push-pull effect. The efficiencies over 10% have been
device displayed panchromatic IPCE response up to 880 nm. ISQ3 achieved for metal-free dyes[7] with I-/I3- redox mediator whereas
sensitized device provided the best efficiency of 4.15 % with short indenoperylene based dyes achieved PCE ≥ 12 % with Co(II/III)
circuit current density (JSC) of 10.02 mAcm-2, open-circuit voltage redox shuttle in a single sensitizer device configuration. [8]
(VOC) of 0.58 V and fill factor (ff) of 72% in the presence of 10 Recently, an efficiency (PCE) of >14% was achieved by a metal-
equivalent CDCA. Electrochemical impedance spectroscopy (EIS) free organic dye bearing silyl anchoring group through co-
analysis showed attenuated charge recombination in ISQ3 sensitization.[9]
sensitized DSSC which contributes to its higher VOC compared to In DSSCs, near-infrared absorbing organic dyes are critically
other dyes. needed as they are capable of generating high current density
which can improve the performance of the cell.[10] While there
has been remarkable progress in DSSC performance of metal-
free organic dyes in the last two decades, there is still ample
Introduction scope for improvement in terms of dye design, especially the
absorption in the near-infrared region (750-1000 nm). Squaraine
Dye-sensitized solar cells (DSSCs) have attracted considerable dyes are among the few sensitizers which have potential to
research attention as a promising alternative to inorganic solar absorb in this region other than porphyrin and phthalocyanine.[11]
cells due to their ease of fabrication, economic viability, and Squaraine dyes belong to the family of polymethine dyes and
tunability of the various device components.[1] The performance comprise of resonance-stabilized zwitterionic structures.[12] Due
of DSSCs has been substantially improved due to significant to strong absorption ( ε > 105 M-1) in far red and NIR region,
advances in the key components, such as molecular sensitizers, squaraine dyes have found applications in wide range of areas
n-type semiconductors, redox couples, and cathode materials in such as imaging,[13] sensors,[14] nonlinear optics,[15]
the past two decades.[2] Sensitizing dyes play a crucial role in [16]
photodynamic therapy , and photovoltaics. [17]
There are
directing the various photovoltaic parameters which eventually several successful DSSCs based on squaraine dyes which
affect the power conversion efficiency (PCE or ) of the DSSCs. absorb in the far-red region, however, only a few of them have a
Several sensitizers, which includes ruthenium (II) complexes, [3] spectral response over 800 nm (Figure 1). The extension of
zinc-porphyrin complexes[4] and metal-free organic dyes[5] have indole based core-symmetric squaraine dyes by using a -
spacer in YR6[18] and RSQ2[19], leads to good incident photon to
current efficiency (IPCE) in NIR region (onset at 775-780 nm)
[a] R. Bisht, M.F. M. Kavungathodi, Dr. J. Nithyanandhan with efficiency ( of 6.7 and 6.8% respectively. The core-
Physical and Materials Chemistry Division
dissymmetric squaraine dye JK-217, assembled using
CSIR-National Chemical Laboratory,
Dr. Homi Bhaba Road, Pune-411008, India. thiophenyl, pyrrolyl and indolium groups, exhibits a
E-mail: [email protected] panchromatic light harvesting up to 850 nm to yield efficiency of
[b] R. Bisht, Dr. J. Nithyanandhan 5.5%.[20] (Detailed explanation on squaraine dyes classification
Academy of Scientific and Innovative Research (AcSIR)
is provided in the supporting information, Scheme S1). In order
New Delhi 110025, India
to extend response further in the NIR region, the fused -
Supporting information for this article is given via a link at the end of extended heterocyclic structures have been incorporated in
the document.
FULL PAPER
squaraine dyes as they can induce strong intramolecular charge assembly on the TiO2 is very important and the most popular
transfer (ICT) as well as maintain planarity of the molecule. method involve co-adsorption of dyes with an optically
transparent substance bearing an anchoring group, such as
3α,7α-dihydroxy-5β-cholanic acid (CDCA) which help to break
down the large aggregates to induce the favorable charge
injection.
Incorporation of bulky alkyl chains on dye skeleton is another
method to avoid aggregation in which π-stacking of molecules is
restricted by increasing the bulk surrounding the
chromophore.[29]. Introducing in-plane and out-of-plane alkyl
groups helps controlling self-organization of dye molecules in
organic photovoltaics.[30] In DSSCs, such introduction of out-of-
plane alkyl groups provides better control over aggregation at a
Accepted Manuscript
molecular level compared to CDCA and improves device
performance.[29c,29d,31] Besides, CDCA can reduce the dye
loading beyond optimal amount if used in excess, especially with
highly aggregating dyes, which makes the integration of out-of-
plane alkyl chains on dye preferable.
Controlling the aggregation of dyes on the TiO2 surface
has been a major challenge in NIR absorbing sensitizers. To this
end, we have designed and synthesized a series of
unsymmetrical squaraine dyes (ISQ1-3) for DSSC consisting of
a fused indenoquinaldine based donor which can help in
increasing the absorption in NIR region and also accommodates
long alkyl chains at its sp3-C center to control dye aggregation
on TiO2 surface (Figure 2). Quinaldine is chosen to fuse with
indene as it has the ability to provide excellent absorption in the
NIR region in squaraine dyes.[32] While all the dyes retain
Figure 1. Examples for the best NIR responsive core-dissymmetric squaraine indenoquinalidne as a donor, carboxyindolenine (ISQ1),
dyes with IPCE onset greater than 800 nm. carboxybenz[e]indoleine (ISQ2) and carboxyquinaldine (ISQ3)
were used towards anchoring group to systematically investigate
their effect on electronic properties of dyes and DSSC
Pyrylium, thiopyrylium, benzo[c,d]indolenine, quinolinium and parameters.
indole based heterocyclic components have been used to
synthesize NIR absorbing squaraine dyes which exhibit IPCE
onset at ca. 800-940 nm.[21] The dye WCH-SQ10, consisting of
triphenylamine as a donor and carboxyquinoline unit as acceptor,
extends the absorption further and IPCE up to 1050 nm is
obtained.[22] In another approach, a fused dimeric squaraine dye
BSQ01[23] shows IPCE response up to 800 nm whereas the
dimeric (LSQa)[24] and trimeric (TSQa-b)[25] squaraine dyes
demonstrate improved IPCE onsets at 900 nm and 980 nm
respectively. Despite the strong NIR absorption, the squaraine
dyes based on fused extended heterocyclic structures have
low efficiency (ca. 1-3 %) due to dye aggregation on TiO2 and
lack of directionality for charge injection from the photo-excited
state of the sensitizer.
The aggregation of dyes on TiO2 is a common feature in
squaraine based dyes which occur due to the interaction of
intermolecular dipoles.[26] Head-to-head dipole interaction forms
H-aggregates which appear as the blue-shifted peak, whereas
head-to-tail dipole interaction leads to and J-aggregates which
appear as red-shifted absorption compared to the absorption
peak of the monomeric dye[27]. While the aggregation of dyes on
Figure 2. (a) Design of steric and electronic effects incorporated
TiO2 broadens the absorption profile, it can also reduce the unsymmetrical squaraine dyes, (b) Structure of fused indenoquinaldine based
charge injection into TiO2 through exciton quenching which donor and (c) Unsymmetrical squaraine dyes ISQ1-3
hampers the performance of DSSC.[28] Thus controlling the dye
FULL PAPER
Results and Discussion described in supporting information (Scheme S2). In the final
step, the semi-squaric acid derivative P1 was refluxed with
Synthesis indenoquinaldinium salt 5 in n-BuOH/PhMe mixture under
The synthetic scheme for the dyes ISQ1-3 is illustrated in azeotropic distillation of water employing Dean−Stark apparatus
scheme 1. to give squaraine dyes ISQ1. Similarly, the dyes ISQ2 and ISQ3
were obtained by treating 5 with P2 and P3 respectively. All the
dyes were thoroughly characterized by 1H-NMR, 13C-NMR
spectroscopy, and mass spectrometric techniques.
Optical characterization
Accepted Manuscript
solutions (Figure 3) and data are summarized in (Table 1). ISQ
dyes exhibited intense absorption bands in the far-red region
with high extinction coefficient (≥105 M-1cm-1) due to strong
intramolecular charge transfer (ICT) process. ISQ1 showed the
absorption maximum (max) at 692 nm with onset at 734 nm due
to the strong donating effect of the indenoquinaldine unit. The
red-shifted max at 715 nm with onset at 760 nm was observed
for ISQ2 due to additional conjugation by carboxybenz[e]indole
unit towards anchoring group. In ISQ3, due to the presence of
quinaldine moiety on both ends of central squaraine unit, the
maximum bathochromic shift was observed, and max was
determined at 754 nm with onset at 800 nm (Figure 3a). The
same trend was observed in the emission spectra of the dyes
Scheme 1. Syntheses of a) indenoquinaldinium iodide 5, (b) Indenoquinaldine
donor based unsymmetrical squaraine dyes (ISQ1-3). which displayed emission maxima at 726 nm, 745 nm and 781
nm for ISQ1, ISQ2 and ISQ3 respectively (Figure 3b). The band
gap (E0-0) was obtained from the intersection of absorption and
The key step in the synthetic route of ISQ dyes is the synthesis emission curve and found to be at 1.75 eV, 1.70 eV and 1.62 eV
of indenoquinalinium salt 5, which begins with functionalization for ISQ1, ISQ2, and ISQ3 respectively. Apart from the intense
of 2-nitrofluorene (1) with 12 carbon long chains by treatment absorption in NIR region, a weak absorption ( ≈ 103 M-1cm-1) in
with NaOH in the presence of dodecyliodide. The resultant the visible region was observed between 400 to 500 nm for
compound 2 was reduced to 2-amino-9,9-didodecylfluorene (3) ISQ1-3 which can be beneficial to absorb the high energy
by hydrazine hydrate in the presence of catalyst Pd/C. The photons as well. To understand the self-assembly of dyes on
amine derivative 3 was treated with crotonaldehyde under TiO2 film, the titania coated glass was immersed in a 0.1 mM
Skraup-Doebner-Von Miller reaction condition to afford dye solution for 15 minutes. Exposing the TiO2 film to the dyes
indenoquinaldine derivative 4 which was subsequently reacted for short duration also gives information about the self-
with iodomethane to give N-methyl indenoquinaldinium salt 5. organization of dyes on the TiO2 surface.
The synthesis of semi-squaric acid derivatives of carboxyindole
P1, carboxybenz[e]indole P2, and carboxyquinoline P3 is
Normalized emission (a. u.)
12 ISQ1 ISQ3
ISQ2 0.8
10 ISQ3
8 0.6
6 0.4
4
0.2
2
0
0.0
400 500 600 700 800 900 600 650 700 750 800 850 900
Wavelength (nm) Wavelength (nm)
FULL PAPER
LHE (%)
0.6 60
ISQ1
0.4 40 ISQ2
ISQ3
0.2 20
0.0 0
Accepted Manuscript
400 500 600 700 800 900 400 500 600 700 800 900 1000
Wavelength (nm) Wavelength [nm]
Figure 3. a) The absorption spectra of ISQ1-3 in CHCl3 ([dye] = 6 µM), (b) Normalized emission spectra ([dye] = 6 µM in CHCl3), λexc = 600nm (ISQ1), 700 nm
(ISQ2) and 720 nm (ISQ3) , (c) Normalized absorption spectra of ISQ dyes immobilized on TiO2, and (d) Light harvesting efficiency of the ISQ dyes.
The absorption spectra of dye anchored on TiO 2 film displayed conduction band of the TiO2 and the dye regeneration by the
large spectral broadening compared to the spectra in solution electrolyte (Figure 4).
due to the interaction of anchoring group (-COOH) with TiO2.[33]
and the formation of H-type (blue-shifted) aggregates
predominantly[34] (Figure 3c). The formation of aggregates were
confirmed by comparing the absorption spectra of dyes on TiO 2
in the absence and the presence of CDCA (Figure S2). Addition
of CDCA leads to a decrease in the extent of H-aggregate with a
concomitant increment of monomeric dyes on the surface. All
the dyes indicated the extensive formation of H-aggregates with
blue-shifted peaks of high-intensity observed at 645, 661 and
620 nm for ISQ1, ISQ2 and ISQ3 respectively. The shoulder
peaks were also observed at 693, 714 and 753 nm for ISQ1,
ISQ2, and ISQ3 nm respectively, which belongs to the
absorption by monomers. ISQ3 exhibited maximum peak
broadening with absorption ranging from 550 nm to 900 nm. The
light harvesting efficiency (LHE = 1-10-A) was calculated by
measuring absorption in transmittance mode of dye adsorbed
films which was obtained by immersing the TiO2 film in the dye
solution for a long period of 6 h (Figure 3d). The LHE of dyes
anchored on TiO2 indicates the ability of dyes to absorb the
fraction of photons of different energy across the solar spectrum.
LHE profiles showed that ISQ1-3 dyes are able to harvest the
photons from visible as well as NIR region with LHE ≥ 90% in
400-500 nm and 550-850 nm range, which makes these
sensitizers suitable candidates for panchromatic absorption. The
dyes ISQ1, ISQ2, and ISQ3 showed LHE onset at 850 nm, 880
nm, and 1000 nm respectively which lie well in the NIR region.
Electrochemical characterization
FULL PAPER
max/CH2Cl2
abs emi
max/CH2Cl2 max/TiO2
abs
ε 104 EgDFT EHOMO E0-0opt ELUMO
Dyes
(nm)a (nm)b (nm)c (M-1cm-1)d (eV)e (V vs NHE)f (eV)g (V vs NHE)h
ISQ1 462, 692 726 645, 693 0.62, 12.9 2.09 0.62 1.75 -1.13
ISQ2 480, 715 745 661, 714 0.56, 11.1 1.97 0.55 1.70 -1.15
ISQ3 487, 754 781 620, 682, 753 0.43, 13.5 1.77 0.47 1.62 -1.15
HOMO levels of the dyes ISQ1-3 were obtained from the first Computational studies
oxidation onset and calculated to be at 0.62 V, 0.55V, and 0.47
Accepted Manuscript
V respectively (Vs. NHE). The upward shift of HOMO in case of To deeply understand the structural, electronic, and optical
ISQ2 and ISQ3 can be credited to the stronger donating effect of properties of ISQ dyes, quantum mechanical calculation were
benz[e]indole and quinoline respectively compared to indole. carried out using density functional theory (DFT) and time-
ISQ1 shows high potential offset (-Greg = 220 mV) with respect dependent DFT (TD-DFT) at B3LYP/6-31G (d,p) level in
to redox level of I-/I3- (0.4 V) which provides large driving force Gaussian 09.[35]. Suitable spatial distributions of the frontier
for effective dye regeneration whereas ISQ2 (-Greg = 150 mV) molecular orbitals (FMO) are crucial for facilitating the electron-
and ISQ3 (-Greg = 70 mV) shows moderate driving force (Figure transfer processes. Ideally, the highest occupied molecular
4b). The LUMO levels were calculated by subtracting E0−0 from orbitals (HOMO) should be away from the TiO 2, and lowest
EHOMO and found to be at -1.13 V, -1.15 V and -1.15 V (vs NHE) unoccupied molecular orbitals (LUMO) should be as close as
for ISQ1, ISQ2, and ISQ3 respectively which gives enough possible to the TiO2 for efficient electron injection. Isodensity
energy offset (-Ginject ≈ 700 mV) for charge injection on surface plots of ISQ1-3 dyes show that HOMO are
conduction band (CB) of TiO2 (-0.5 V). While the HOMO levels predominantly confined over central squaric acid units (Figure 5).
varied significantly, the LUMO levels were essentially the same
for all the dyes.
Figure 5. Isodensity surface plots of ISQ1-3 dyes (alkyl chains are removed for clarity).
FULL PAPER
The HOMO and LUMO overlap for the most part of ISQ1-3 dyes,
however, there is slight delocalization of LUMO towards the Photovoltaic characterization
anchoring group in case of ISQ3 LUMO+1 and LUMO+2 are
prominently delocalized towards anchoring group for the ISQ3 The photovoltaic measurements of the DSSCs fabricated using
dye which suggests that it can exhibit better charge injection ISQ1-3 dyes were done under irradiance of 100 mWcm-2
compared to ISQ1-2 and can counterbalance the effect of low simulated AM 1.5G sunlight. The current density-voltage (J-V)
driving force for dye regeneration. Simulated absorption spectra plot and IPCE spectra of the best cells are shown in Figure 7
were plotted using TD-DFT (Figure S1), and spectra were found and summarized in table 1. The DSSC based on ISQ1 gave the
to be consistent with the experimental spectra. By comparing the PCE of 3.24 % whereas ISQ2 and ISQ3 showed PCE of 2.35 %
two spectra it can be predicted that the peaks of high oscillator and 1.85% respectively, without using any co-adsorbent. The
strength at longer wavelength originates exclusively from major difference in their performance comes from the variation in
HOMO-LUMO vertical energy transition (> 99 % contribution) in the short circuit current density (JSC) which is highest for ISQ1
all the dyes whereas the high energy peaks with low oscillator (8.99 mAcm-2), followed by ISQ2 (6.85 mA cm-2) and ISQ3 (5.23
Accepted Manuscript
strength between 350 and 500 nm are associated with the mA cm-2). The low short circuit current density could be ascribed
transition to higher excited states, i.e., LUMO+1, LUMO+2 to the strong aggregation of dye on TiO2 which is maximum in
(Table S1). The optimized geometry in Figure 6 shows that the case of ISQ3. To minimize the negative effect of aggregation on
distance between branched alkyl chains and oxygen atoms of cell performance, the dye was co-adsorbed in the presence of
anchoring unit d is about 20.5 Å, 22.Å and 19.8 Å for ISQ1, CDCA to reduce the self-quenching of excitons and charge
ISQ2, and ISQ3 respectively. recombination. All the ISQ dyes showed a different response
with the addition of CDCA. The efficiency of ISQ1 dyes reduced
to 2.63 % mainly due to a reduction in JSC (6.86 mA cm-2), even
though the slight improvement in VOC was observed (0.558 V). In
contrast, the performance of ISQ2 and ISQ3 sensitized cells
improved when co-adsorbed with CDCA. ISQ2 achieved the
maximum efficiency of 3.68% on addition of 5 equiv CDCA with
VOC of 0.558 V and JSC of 9.62 mA cm-2 whereas ISQ3 reached
the best efficiency of 4.15% with JSC of 10.02 mA cm-2, VOC of
0.576 V, and ff of 72%, in presence 10 equiv of CDCA. Further
addition of CDCA led to a reduction in device efficiency of all the
dyes, primarily due to substantial loss of JSC (Figure S3 and
Table S2). It was suspected that the decrease in efficiency of
ISQ1 on the addition of CDCA could be due to reduced
adsorption of dyes on the TiO2 surface due to competitive
binding of CDCA. To confirm our hypothesis, the dye loading
was calculated by desorption of dye-sensitized photo-anodes in
the presence of 2M ethanolic HCl followed by
spectrophotometric measurement of dye concentration (Table 2).
In the case of ISQ1, the addition of CDCA, as low as 5
equivalents, led to the drastic decrease of 68% in dye content.
The dye loading of ISQ2 with 5 equiv of CDCA (1.22 × 10-7 mol
cm-2) was significantly higher than the dye content of ISQ1
devices with 5 equiv. CDCA (7.56 × 10-8 mol cm-2). ISQ3
sensitized cells, even in the presence of 10 equiv CDCA,
showed higher dye content (1.05 x 10-7 mol cm-2) compared to
ISQ1 (5 equiv CDCA). This shows that dye-TiO2 interaction is
Figure 6. Optimized geometry of ISQ1-3 dyes, a) Lateral view, b) Top view
poor in case of ISQ1 which allows the dyes to be replaced by
CDCA on TiO2 rather easily compared to ISQ2 and ISQ3 (Figure
8b (ii).
Thus, the variable effect of CDCA on performance of DSSC
The dihedral angle between indenoquinalidne and squaraine
unit (θ1) is the same for all the dyes whereas the dihedral angle could be ascribed to competitive binding of ISQ1-3 dyes with
between squaraine units and anchoring unit is different for all the CDCA, where on the one hand CDCA helps in improving JSC
and VOC by reducing exciton quenching induced by aggregation
dyes. The dye ISQ1 is almost planar with θ1= 4.5° and θ2=0°,
on TiO2 surface, on the other hand, it may replace the dye
ISQ2 is non-planar with θ2=11.7°, and planarity is further
reduced in ISQ3 with θ2 = 21.2°. The distance between the molecules to reduce the effective dye content on the photo-
terminal carbons of the two out of plane alkyl chains on the dyes anode leading to poor performance.[19]
is about 27 Å which helps in controlling the dye-dye interactions
on the TiO2 surface (Figure 6b).
FULL PAPER
12 80
(a) ISQ1 (b) ISQ1
10 ISQ2 ISQ2
ISQ3 ISQ3
60
J (mA/cm2)
IPCE (%)
8
6 40
4
20
2
0 0
Accepted Manuscript
0.0 0.2 0.4 0.6 400 500 600 700 800 900
Voltage (V) Wavelength (nm)
12 80
(c) (d) ISQ1/CDCA (1:5)
ISQ2/CDCA (1:5)
10 ISQ3/CDCA (1:10)
60
J (mA/cm2)
8 IPCE [%]
6 40
4
ISQ1/CDCA (1:5) 20
2 ISQ2/CDCA (1:5)
ISQ3/CDCA (1:10)
0 0
0.0 0.2 0.4 0.6 400 500 600 700 800 900
Voltage (V) Wavelength [nm]
Figure 7. J−V and IPCE characterization of ISQ dyes. (a, b) J−V and IPCE curves for ISQ1-3 in the absence of CDCA. (c, d) J−V and IPCE curves for ISQ1-3 in
the presence of CDCA.
Table 2. Photovoltaic parameters for DSSCs employing the I-/I3- electrolyte under the optimized conditionsa
a
See SI for experimental details of measurements and calculations. bCalculated by integrating IPCE over the standard AM1.5G emission spectrum (ASTM G173-
03).
FULL PAPER
The variation in JSC can be further explained through a detailed distinct humps on the curve which are contributed by the H-type
investigation of IPCE spectra as the JSC is directly related to aggregates (39 % at 650 nm), monomers (40 % at 710 nm) and
IPCE and can be expressed as eq 1, where q is the charge of J-type aggregates (34 % at 800 nm). In the case of ISQ3 (10
the electron, and ϕ is the photon flux of solar spectrum (AM 1.5 equiv CDCA), the contribution by J- aggregates was more
G). prominent with IPCE over 50% at 800 nm whereas IPCE of 30-
40 % was obtained in the range of 600-750 nm which was
JSC = q ∫ IPCE (λ) ϕ (λ) dλ (1) contributed mainly by the monomers, dimers and H-type dye
aggregates on the TiO2 surface. The results indicate that for
In the absence of CDCA, ISQ1 showed best IPCE response with ISQ2 and ISQ3, the addition of CDCA facilitated the formation of
a maximum of 46% at 600 nm which is primarily contributed by smaller aggregates in a controlled fashion which helped to inject
the H-type aggregates (Figure 6b). The peak with low IPCE electron more efficiently while maintaining the broad spectrum in
(12%) at 810 nm can be attributed to current generation from J- comparison to the devices without CDCA (Figure 8b (ii)).
aggregates. When CDCA is added, the aggregation was Remarkably, a substantial IPCE response of over 40 % in the
Accepted Manuscript
reduced effectively which can be observed as the visible region (400-550 nm) was also observed for ISQ3. The
disappearance of the peaks at 600 nm and 810 nm (Figure 6d). origin of this response can be traced to the high concentration of
The reduction in aggregation, though improved the VOC from electron density on anchoring group in higher excited states
0.544V to 0.558 V in ISQ1 based cell, the severe decrease in (LUMO, LUMO+1, and LUMO+2) which led to strong electronic
dye amount led to low IPCE which in turn led to reduced JSC. In coupling and good charge injection into the TiO 2 despite low
the case of ISQ2, the H-aggregates (23 % at 640 nm) and J- oscillator strength. Hence, with the help of panchromatic IPCE,
aggregates (25 % at 790 nm) contributed almost equally to IPCE, ISQ3 attained better JSC compared to other dyes and eventually
however, the overall response was much lower compared to higher efficiency.
ISQ1. Similarly, for ISQ3 the IPCE response was quite low with Though ISQ dyes have similar indenoquinaline based donor on
equal contribution from H- (16 % at 660 nm) and J- aggregates the non-anchoring side, the anchoring unit containing different
(16 % at 780 nm). The reason behind the low IPCE for ISQ2 and donor units plays a vital role in determining the stability of the
ISQ3, in the absence of CDCA, could be the formation of closely monolayer thus formed on the TiO2 surface. For example, the
stacked large-sized aggregates which could not inject an nature of IPCE profile of ISQ3 was almost similar with and
electron into the TiO2 (Figure 8a). The extended -surface of without CDCA indicating the contribution from monomers, H-
ISQ2 and ISQ3 due to benz[e]indole and quinoline moieties and J-aggregates (Figure S3(f)). Whereas for ISQ2, the IPCE
respectively may induce such close intermolecular packing, and profile showed the contribution from monomers, H- and J-
the absence of out-of-plane two methyl group on aggregates up to 5 equiv. of CDCA, and further addition of
carboxyquinoline group may cause the ISQ3 to be aggregate CDCA leads to contribution mainly from monomers and dimers
excessively. When CDCA was added, both ISQ2 and ISQ3 (Figure S3(d)). In the case of ISQ1, without CDCA, the obtained
showed improved IPCE response. ISQ2 (5 equiv CDCA) photocurrent is due to both monomer and aggregates (H- and J-
showed IPCE of over 30% in between 650 to 800 nm with three ).
Figure 8. Cartoon representation of the formation of dye aggregate and variation in dye loading in the (a) absence of CDCA and the (b) presence of CDCA.
FULL PAPER
Addition of 5 equiv of CDCA disturbed the self-assembly of ISQ1, VOC= Eredox – EF (2)
and consequently, the photocurrent is derived mainly from EF = ECB + kBT ln (nc/Nc) (3)
monomer and dimer (Figure S3(b)). Hence the nature of the
anchoring unit containing donor units not only modulate the According to eq 3, the variation in EF is linked to the number of
electronic property of ISQ dyes but also controls the self- electrons on the TiO2 (nc) and conduction band (ECB) of TiO2,
assembly of dyes on the TiO2 surface. where k is Boltzmann constant, Nc is the effective density of
state, and T is the absolute temperature (293 K). Thus, VOC is
Electrochemical Impedance Spectroscopy (EIS). intimately susceptible to shift in the TiO2 conduction band edge
(ECB), which can be deduced from the chemical capacitance
To understand the slight difference in VOC values of ISQ (C) and fluctuation of electron density, which is associated to
sensitized cell electrochemical impedance spectroscopy (EIS) the electron lifetime(s) determined by the charge recombination
was carried out under dark conditions. Referring to the energy rate.[37] Typical EIS Nyquist plots for DSSCs based on the ISQ1-
band diagram and the carrier transfer processes in Figure 4, the 3 dyes were measured in the dark under applied bias −0.5, -0.47,
Accepted Manuscript
VOC is calculated from the potential difference between the -0.44 and -0.41 V and are fitted using RS + Rpt/Cpt + Rct/C circuit
quasi-Fermi level of TiO2 (EF) and the chemical potential of model (Figure 9). The larger semicircle at lower frequencies in
redox species (Eredox) in an electrolyte (eq2).[36] Nyquist plots represents the interfacial charge-transfer
resistance (Rct) at the TiO2-dye/electrolyte interface.
40
(b) ISQ1
35 ISQ2/CDCA (1:5)
ISQ3/CDCA (1:10)
30
Rct (ohm)
25
20
15
10
5
0
0.40 0.42 0.44 0.46 0.48 0.50
Applied Potential (V)
ISQ1
20
1.2 (c) (d) ISQ1
ISQ2/CDCA (1:5) ISQ2/CDCA (1:5)
ISQ3/CDCA (1:10) ISQ3/CDCA (1:10)
Cpacitance (mF)
0.9 15
Lifetime (ms)
0.6 10
0.3 5
0.40 0.42 0.44 0.46 0.48 0.50 0.40 0.42 0.44 0.46 0.48 0.50
Applied Potential (V) Applied Potential (V)
Figure 9. (a) The Nyquist plot at an applied bias of 0.5 V, (b) Charge transfer resistance (Rct), (b) Rct vs. applied potential, (c) Cμ vs. applied potential, and d) τ vs.
applied potential.
FULL PAPER
The fitted recombination resistance (Rct) value of ISQ sensitized Details on the reagents, equipments utilized to characterize the
devices (under an applied bias of -0.5 V) increases in the order synthesized precursors, final dye molecules and evaluating the
of ISQ1 (7.25 Ω) < ISQ2 (10.08 Ω) < ISQ3 (13.13 Ω), which is Photophysical, electrochemical, photovoltaics parameters,
consistent with the sequence of VOC values (Figure 8b and Table electrochemical impedance spectroscopy and procedure for the
3). Larger the Rct value, slower is the recombination of electrons DSSC device fabrication is provided in the supporting
from the conduction band of TiO2 to the oxidized I3- species in information.
the electrolyte. The electron lifetime(s) on TiO2 were calculated
from Rct and chemical capacitance Cµ using = Rct × Cµ. The Synthetic procedure
longer electron lifetime in ISQ3 sensitized cell further supports
the higher VOC for ISQ3 (10.24 ms) compared to ISQ1(2.39 ms) 9,9-Didodecyl-2-nitro-9H-fluorene (2). 2-Nitrofluorene (3.0 g, 14.20
and ISQ2 (5.94 ms) based cells (Figure 9d and Table 3). mmol), 1-iodododecane (12.6 g, 42.61 mmol) and tetrabutyl ammonium
bromide (0.3 g, 0.710 mmol) were taken in 70 mL toluene in two-necked
round bottom flask. An aqueous solution of NaOH (23 mL, 50 % in
Table 3. EIS Parameters of ISQ Dye Cells at an Applied Potential of −0.5 V in weight) was added to the mixture rapidly with constant stirring. The
Accepted Manuscript
the Dark.
resultant mixture was stirred at 60 ºC for 8 h under argon atmosphere.
Reaction mixture was cooled to room temperature and 1 M HCl (100 mL)
Rs
ISQ Dyes Rct (ohm) C (mF) (ms) was added to it followed by addition of water (100 mL). The reaction
(ohm)
mixture was extracted with ethyl acetate, washed with brine and dried
over anhydrous sodium sulphate. Solvents were removed under reduced
ISQ1 15.80 7.25 0.33 2.39 pressure and the residue was purified by column chromatography
(EtOAc-pet ether) to afford pure compound 2 as pale yellow oil ( 5.13 g,
ISQ2/CDCA 5 eqv 14.56 10.08 0.59 5.94 66 %). 1H NMR (500 MHz, CDCl3) δ 8.26 (dd, J = 8.3, 2.0 Hz, 1H), 8.20
(d, J = 2.0 Hz, 1H), 7.83 – 7.75 (m, 2H), 7.45 – 7.36 (m, 3H), 2.08 – 1.96
(m, 4H), 1.27 – 1.12 (m, 24H), 1.08 – 1.00 (m, 12H), 0.86 (t, J = 7.0 Hz,
ISQ3/CDCA 10 eqv 14.41 13.13 0.78 10.24 6H), 0.68 – 0.46 (m, 4H). 13C NMR (125 MHz, CDCl3) δ 152.5, 152.1,
147.8, 147.3, 138.9, 129.4, 127.5, 123.4, 123.436, 121.3, 119.9, 118.4,
77.2, 55.8, 40.2, 32.0, 30.0, 29.7, 29.7, 29.6, 29.5, 29.4, 23.9, 22.8, 14.3.
HRMS (ESI) m/z: [M+H]+ Calcd for C37H58NO2 548.4468, found 548.4456.
FULL PAPER
0.85 (t, J = 6.9 Hz, 6H), 0.68 – 0.58 (m, 4H). 13C NMR (100 MHz, CDCl3) – 0.45 (m, 4H). 13C NMR (100 MHz, DMSO-d6 /CDCl3 (1:1)) δ 176.5,
δ 158.3, 153.4, 151.4, 148.2, 140.2, 140.0, 136.5, 128.3, 127.2, 126.3, 170.8, 170.2, 156.1, 151.8, 150.5, 149.1, 143.6, 140.2, 139.2, 138.5,
123.3, 122.2, 121.5, 120.5, 117.0, 55.2, 41.4, 32.0, 30.2, 29.8, 29.7, 29.7, 135.2, 132.0, 131.3, 129.9, 129.5, 128.3, 127.5, 126.7, 125.0, 124.0,
29.4, 29.4, 25.5, 24.1, 22.8, 14.3. HRMS (ESI) m/z: [M+H]+ Calcd for 123.1, 120.5, 119.1, 114.7, 110.3, 98.9, 94.5, 93.7, 79.2, 56.0, 37.5, 36.1,
C41H62N 568.4882, found 568.4877. 31.8, 29.7, 29.5, 29.4, 29.2, 29.1, 23.8, 22.6, 14.3. HRMS (ESI) m/z: [M]+
Calcd for C58H72N2O4 860.5492, found 860.5474.
10,10-Didodecyl-1,2-dimethyl-10H-indeno[1,2-g]quinolin-1-ium
iodide (5). The compound 4 (0.68 g, 1.2 mmol), MeI (0.85g, 6 mmol) and
2 mL acetonitrile were taken in a sealed tube and heated at 100 ºC for 48
Acknowledgments
h. After the completion of the reaction the solvents were removed to give
compound 5 as yellow gum which was used without further purification
(0.767 g, 90%). 1H NMR (200 MHz, CDCl3) δ 8.87 (d, J = 8.5 Hz, 1H), This work is financially supported by the Council of Scientific and
8.39 (s, 1H), 8.22 (s, 1H), 8.01 – 7.87 (m, 2H), 7.52 – 7.40 (m, 3H), 4.77 Industrial Research (CSIR) Network Project NWP0054 (CSIR-
(s, 3H), 3.33 (s, 3H), 2.25 – 2.07 (m, 4H), 1.22 – 1.02 (m, 35H), 0.85 (t, J TAPSUN), and SERB-EMR/2016/007114, India. R.B. thanks
= 6.4 Hz, 6H), 0.63 – 0.44 (m, 4H). 13C NMR (100 MHz, CDCl3) δ 161.7,
Accepted Manuscript
CSIR, New Delhi for research fellowships. J.N. thanks Dr.
159.0, 151.3, 145.4, 143.9, 140.1, 137.4, 130.6, 128.8, 128.1, 124.8, Kothandam Krishnamoorthy, Polymer Science Engineering
123.6, 121.8, 119.4, 112.1, 57.0, 54.5, 41.6, 41.1, 32.0, 30.0, 29.71,
Division, CSIR-National Chemical Laboratory, Pune, India, for
29.65, 29.4, 24.7, 24.2, 22.8, 14.2. HRMS (ESI) m/z: [M]+ Calcd for
C42H64N+ 582.5033, found 582.5031. his support, help with device fabrication and characterization.
Authors also thank Mr. Ananthan Alagumalai for help with
General procedure for synthesizing ISQ dyes: Indenoquinaldinium making a precursor.
iodide, 5 (1 equiv) and respective semi-squaric acid derivatives P1, or P2
or P3 (1.2 equiv) were dissolved in 10 mL of n-BuOH/PhMe (1:1) in a Keywords: unsymmetrical squaraine dyes, Indenoquinaldine,
two-necked round bottom flask and fitted with Dean-Stark apparatus.
H- and J-type aggregation, dye-sensitized solar cells, NIR active
Quinoline (1 mL) was added to the reaction mixture and refluxed for 24 h
under argon atmosphere. The reaction mixture was cooled to room dyes, CDCA
temperature, and the solvents were removed under reduced pressure.
The residue was purified by column chromatography (MeOH-CH2Cl2) to
afford the desired dye as dark green solid. [1] a) A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo, H. Pettersson, Chem.
Rev. 2010, 110, 6595–6663; b) M. Grätzel, J. Photochem. Photobiol.
C Photochem. Rev. 2003, 4, 145–153.
ISQ1: Started with 5 (0.2 g, 0.28 mmol) and P1 (0.105 g, 0.34 mmol). [2] a) B. E. Hardin, H. J. Snaith, M. D. McGehee, Nat. Photonics 2012, 6,
Yield: 0.165 g, 67 %. m.p. 200-202 oC. 1H NMR (400 MHz, CDCl3) δ 9.57 162–169; b) J. Wu, Z. Lan, J. Lin, M. Huang, Y. Huang, L. Fan, G. Luo,
Chem. Rev. 2015, 115, 2136–2173; c) S. Thomas, T. G. Deepak, G. S.
(d, J = 9.3 Hz, 1H), 8.09 (d, J = 8.4 Hz, 1H), 8.03 (s, 1H), 7.98 – 7.87 (m, Anjusree, T. A. Arun, S. V. Nair, A. S. Nair, J. Mater. Chem. A 2014, 2,
2H), 7.85 – 7.72 (m, 1H), 7.54 (s, 1H), 7.44 – 7.34 (m, 3H), 6.87 (d, J = 4474–4490; d) M.-E. Yeoh, K.-Y. Chan, Int. J. Energy Res. 2017, 41,
8.4 Hz, 1H), 6.30 (s, 1H), 5.84 (s, 1H), 4.13 (s, 3H), 3.44 (s, 3H), 2.15 – 2446–2467; e) J. Gong, K. Sumathy, Q. Qiao, Z. Zhou, Renew.
1.96 (m, 4H), 1.83 (s, 6H), 1.24 – 1.02 (m, 36H), 0.84 (t, J = 6.9 Hz, 6H), Sustain. Energy Rev. 2017, 68, 234–246; f) M. Freitag, J. Teuscher, Y.
Saygili, X. Zhang, F. Giordano, P. Liska, J. Hua, S. M. Zakeeruddin,
0.65 – 0.49 (m, 4H). 13C NMR (100 MHz, CDCl3) δ 184.7, 181.8, 180.8,
J.-E. Moser, M. Grätzel, A. Hagfeldt, Nat. Photonics 2017, 11, 372–
170.5, 168.1, 165.4, 156.9, 153.7, 150.6, 148.1, 141.6, 139.8, 138.9, 378; g) P. Pinpithak, A. Kulkarni, H.-W. Chen, M. Ikegami, T.
136.8, 131.3, 128.8, 127.6, 125.7, 125.6, 124.0, 123.3, 123.2, 120.7, Miyasaka, Bull. Chem. Soc. Jpn. 2018, 91, 754–760; h) H.
119.2, 109.7, 107.3, 96.0, 87.8, 56.2, 47.5, 41.1, 37.8, 32.00, 30.4, 30.1, Shimogawa, M. Endo, T. Taniguchi, Y. Nakaike, M. Kawaraya, H.
Segawa, Y. Murata, A. Wakamiya, Bull. Chem. Soc. Jpn. 2017, 90,
29.7, 29.6 29.4, 29.4, 27.6, 24.0, 22.8, 14.2. HRMS (ESI) m/z: [M]+ Calcd
441–450.
for C59H76N2O4 876.5805, found 876.5793. [3] a) B. O’Regan, M. Grätzel, Nature 1991, 353, 737–740; b) A. S. Polo,
M. K. Itokazu, N. Y. Murakami Iha, Coord. Chem. Rev. 2004, 248,
1343–1361; c) C.-Y. Chen, M. Wang, J.-Y. Li, N. Pootrakulchote, L.
ISQ2: Started with 5 (0.2 g, 0.28 mmol) and P2 (0.101 g, 0.34 mmol).
Alibabaei, C. Ngoc-le, J.-D. Decoppet, J.-H. Tsai, C. Grätzel, C.-G.
Yield: 0.121 g, 50 %. m.p.180-182 oC. 1H NMR (500 MHz, CDCl3) δ 9.54 Wu, S. M. Zakeeruddin, M. Grätzel, ACS Nano 2009, 3, 3103–3109.
(d, J = 9.1 Hz, 1H), 8.70 (s, 1H), 8.19 (d, J = 11.0 Hz, 2H), 7.93 (d, J = [4] a) L.-L. Li, E. W.-G. Diau, Chem. Soc. Rev. 2012, 42, 291–304; b) S.
8.0 Hz, 1H), 7.87 (d, J = 12.9 Hz, 2H), 7.77 (d, J = 4.4 Hz, 1H), 7.49 (s, Mathew, A. Yella, P. Gao, R. Humphry-Baker, B. F. E. Curchod, N.
1H), 7.38 (d, J = 6.1 Hz, 3H), 7.27 (s, 1H), 6.25 (s, 1H), 5.89 (s, 1H), 4.08 Ashari-Astani, I. Tavernelli, U. Rothlisberger, M. K. Nazeeruddin, M.
Grätzel, Nat. Chem. 2014, 6, 242–247; c) M. Urbani, M. Grätzel, M. K.
(s, 3H), 3.55 (s, 3H), 2.10 (s, 6H), 2.06 – 1.97 (m, 4H), 1.23 – 1.11 (m, Nazeeruddin, T. Torres, Chem. Rev. 2014, 114, 12330–12396; d) T.
24H), 1.06 – 1.01 (m, 12H), 0.84 (t, J = 7.1 Hz, 6H), 0.64 – 0.55 (m, 4H). Higashino, H. Imahori, Dalton Trans. 2014, 44, 448–463.
13
C NMR (126 MHz, CDCl3) δ 184.9, 181.9, 179.0, 168.8, 168.1, 156.6, [5] a) A. Mishra, M. K. R. Fischer, P. Bäuerle, Angew. Chem. Int. Ed.
153.2, 150.6, 139.9, 139.4, 139.0, 136.1, 133.7, 133.1, 131.4, 131.2, 2009, 48, 2474–2499; b) M. Liang, J. Chen, Chem. Soc. Rev. 2013,
42, 3453–3488; c) Z.-S. Huang, H. Meier, D. Cao, J. Mater. Chem. C
129.7, 128.6, 127.6, 127.0, 125.7, 125.4, 124.9, 123.2, 122.4, 120.6, 2016, 4, 2404–2426.
119.1, 110.5, 109.5, 95.4, 86.8, 56.1, 49.9, 41.1, 37.6, 32.0, 30.5, 30.1, [6] a) M. Liang, J. Chen, Chem. Soc. Rev. 2013, 42, 3453–3488; b) A.
29.7, 29.7, 29.4, 29.4, 27.1, 24.0, 22.8, 14.2. HRMS (ESI) m/z: [M]+ Mahmood, Sol. Energy 2016, 123, 127–144; c) Y. Wu, W.-H. Zhu, S.
Calcd for C63H78N2O4 926.5962, found 926.5956. M. Zakeeruddin, M. Grätzel, ACS Appl. Mater. Interfaces 2015, 7,
9307–9318; d) Y. Wu, W. Zhu, Chem. Soc. Rev. 2013, 42, 2039–2058.
[7] a) W. Zeng, Y. Cao, Y. Bai, Y. Wang, Y. Shi, M. Zhang, F. Wang, C.
ISQ3: Started with 5 (0.2 g, 0.28 mmol) and P3 (0.124 g, 0.34 mmol). Pan, P. Wang, Chem. Mater. 2010, 22, 1915–1925; b) N. Zhou, K.
Yield: 0.16 g, 60%. m.p. 142-144 oC. 1H NMR (400 MHz, [D6]DMSO Prabakaran, B. Lee, S. H. Chang, B. Harutyunyan, P. Guo, M. R.
Butler, A. Timalsina, M. J. Bedzyk, M. A. Ratner, S. Vegiraju, S. Yau,
/CDCl3 (1:1)) δ 9.29 (d, J = 9.3 Hz, 1H), 9.05 (d, J = 9.5 Hz, 1H), 8.01 (s, C.-G. Wu, R. P. H. Chang, A. Facchetti, M.-C. Chen, T. J. Marks, J.
1H), 7.97 (s, 2H), 7.82 (d, J = 9.4 Hz, 1H), 7.75 (d, J = 4.5 Hz, 1H), 7.69 Am. Chem. Soc. 2015, 137, 4414–4423; c) Y. Ezhumalai, B. Lee, M.-
(s, 1H), 7.41 (d, J = 8.9 Hz, 1H), 7.37 – 7.33 (m, 2H), 7.32 – 7.29 (m, 2H), S. Fan, B. Harutyunyan, K. Prabakaran, C.-P. Lee, S. H. Chang, J.-S.
5.84 (s, 1H), 5.57 (s, 1H), 3.97 (s, 3H), 3.64 (s, 3H), 2.08 – 1.96 (m, 4H), Ni, S. Vegiraju, P. Priyanka, Y.-W. Wu, C.-W. Liu, S. Yau, J. T. Lin, C.-
1.17 – 1.05 (m, 24H), 1.00 – 0.95 (m, 12H), 0.79 (t, J = 6.8 Hz, 6H), 0.52
FULL PAPER
G. Wu, M. J. Bedzyk, R. P. H. Chang, M.-C. Chen, K.-C. Ho, T. J. [31] A. Alagumalai, M. F. M. K., P. Vellimalai, M. C. Sil, J. Nithyanandhan,
Marks, J. Mater. Chem. A 2017, 5, 12310–12321. ACS Appl. Mater. Interfaces 2016, 8, 35353–35367.
[8] a) Z. Yao, M. Zhang, H. Wu, L. Yang, R. Li, P. Wang, J. Am. Chem. [32] a) Z. Yan, S. Guang, X. Su, H. Xu, J. Phys. Chem. C 2012, 116,
Soc. 2015, 137, 3799–3802; b) Z. Yao, M. Zhang, R. Li, L. Yang, Y. 8894–8900; b) K. Jyothish, K. T. Arun, D. Ramaiah, Org. Lett. 2004, 6,
Qiao, P. Wang, Angew. Chem. Int. Ed. 2015, 54, 5994–5998; c) Y. 3965–3968.
Ren, D. Sun, Y. Cao, H. N. Tsao, Y. Yuan, S. M. Zakeeruddin, P. [33] K. Hara, Z.-S. Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y.
Wang, M. Grätzel, J. Am. Chem. Soc. 2018, 140, 2405–2408. Dan-oh, C. Kasada, A. Shinpo, S. Suga, J. Phys. Chem. B 2005, 109,
[9] K. Kakiage, Y. Aoyama, T. Yano, K. Oya, J. Fujisawa, M. Hanaya, 15476–15482.
Chem. Commun. 2015, 51, 15894–15897. [34] S.-L. Li, K.-J. Jiang, K.-F. Shao, L.-M. Yang, Chem. Commun. 2006, 0,
[10] S. Zhang, X. Yang, Y. Numata, L. Han, Energy Environ. Sci. 2013, 6, 2792–2794.
1443. [35] M. J. Frisch et al., Gaussian 09;, Gaussian, Inc.: Wallingford, CT, USA,
[11] a) C. Qin, W.-Y. Wong, L. Han, Chem. - Asian J. 2013, 8, 1706–1719; 2009. (Complete Reference Is Provided in Supporting Information) .
b) P. Brogdon, H. Cheema, J. H. Delcamp, ChemSusChem 2018, [36] T. Marinado, K. Nonomura, J. Nissfolk, M. K. Karlsson, D. P. Hagberg,
11, 86–103; c) L. Martín-Gomis, F. Fernández-Lázaro, Á. Sastre- L. Sun, S. Mori, A. Hagfeldt, Langmuir 2010, 26, 2592–2598.
Santos, J. Mater. Chem. A 2014, 2, 15672–15682; d) M.-E. Ragoussi, [37] a) J. Bisquert, D. Cahen, G. Hodes, S. Rühle, A. Zaban, J. Phys.
M. Ince, T. Torres, Eur. J. Org. Chem. 2013, 2013, 6475–6489. Chem. B 2004, 108, 8106–8118; b) E. Ronca, M. Pastore, L. Belpassi,
[12] a) A. Ajayaghosh, Acc. Chem. Res. 2005, 38, 449–459; b) S. Sreejith, F. Tarantelli, F. D. Angelis, Energy Environ. Sci. 2012, 6, 183–193.
P. Carol, P. Chithra, A. Ajayaghosh, J. Mater. Chem. 2008, 18, 264–
Accepted Manuscript
274.
[13] a) K. Y. Law, Chem. Rev. 1993, 93, 449–486; b) S. Luo, E. Zhang, Y.
Su, T. Cheng, C. Shi, Biomaterials 2011, 32, 7127–7138; c) V. J.
Pansare, S. Hejazi, W. J. Faenza, R. K. Prud’homme, Chem. Mater.
2012, 24, 812–827; d) J. O. Escobedo, O. Rusin, S. Lim, R. M.
Strongin, Curr. Opin. Chem. Biol. 2010, 14, 64–70.
[14] a) P. Chithra, R. Varghese, K. P. Divya, A. Ajayaghosh, Chem. –
Asian J. 2008, 3, 1365–1373; b) P. Anees, S. Sreejith, A. Ajayaghosh,
J. Am. Chem. Soc. 2014, 136, 13233–13239; c) P. Anees, K. V.
Sudheesh, P. Jayamurthy, A. R. Chandrika, R. V. Omkumar, A.
Ajayaghosh, Chem. Sci. 2016, 7, 6808–6814; d) Philips Divya S.,
Ghosh Samrat, Sudheesh Karivachery V., Suresh Cherumuttathu H.,
Ajayaghosh Ayyappanpillai, Chem. – Eur. J. 2017, 23, 17973–17980.
[15] a) C. W. Dirk, W. C. Herndon, F. Cervantes-Lee, H. Selnau, S.
Martinez, P. Kalamegham, A. Tan, G. Campos, M. Velez, J. Am.
Chem. Soc. 1995, 117, 2214–2225; b) C. Prabhakar, K.
Bhanuprakash, V. J. Rao, M. Balamuralikrishna, D. N. Rao, J. Phys.
Chem. C 2010, 114, 6077–6089.
[16] a) D. Ramaiah, I. Eckert, K. T. Arun, L. Weidenfeller, B. Epe,
Photochem. Photobiol. 2002, 76, 672–677; b) A. Yuan, J. Wu, X. Tang,
L. Zhao, F. Xu, Y. Hu, J. Pharm. Sci. 2013, 102, 6–28.
[17] a) C. Qin, W.-Y. Wong, L. Han, Chem. – Asian J. 2013, 8, 1706–1719;
b) G. Chen, H. Sasabe, T. Igarashi, Z. Hong, J. Kido, J. Mater.
Chem. A 2015, 3, 14517–14534.
[18] Y. Shi, R. B. M. Hill, J.-H. Yum, A. Dualeh, S. Barlow, M. Grätzel, S. R.
Marder, M. K. Nazeeruddin, Angew. Chem. Int. Ed. 2011, 50, 6619–
6621.
[19] R. Bisht, M. F. M. K., A. K. Singh, J. Nithyanandhan, J. Org. Chem.
2017, 82, 1920–1930.
[20] S. Paek, H. Choi, C. Kim, N. Cho, S. So, K. Song, M. K. Nazeeruddin,
J. Ko, Chem. Commun. 2011, 47, 2874–2876.
[21] a) T. Maeda, S. Nitta, H. Nakao, S. Yagi, H. Nakazumi, J. Phys. Chem.
C 2014, 118, 16618–16625; b) T. Maeda, S. Nitta, Y. Sano, S. Tanaka,
S. Yagi, H. Nakazumi, Dyes Pigments 2015, 122, 160–167; c) T.
Maeda, N. Shima, T. Tsukamoto, S. Yagi, H. Nakazumi, Synth. Met.
2011, 161, 2481–2487; d) K. Funabiki, H. Mase, Y. Saito, A. Otsuka,
A. Hibino, N. Tanaka, H. Miura, Y. Himori, T. Yoshida, Y. Kubota, M.
Matsui, Org. Lett. 2012, 14, 1246–1249.
[22] J.-Y. Li, C.-Y. Chen, W.-C. Ho, S.-H. Chen, C.-G. Wu, Org. Lett. 2012,
14, 5420–5423.
[23] S. Kuster, F. Sauvage, M. K. Nazeeruddin, M. Grätzel, F. A. Nüesch,
T. Geiger, Dyes Pigments 2010, 87, 30–38.
[24] T. Maeda, Y. Hamamura, K. Miyanaga, N. Shima, S. Yagi, H.
Nakazumi, Org. Lett. 2011, 13, 5994–5997.
[25] T. Maeda, S. Arikawa, H. Nakao, S. Yagi, H. Nakazumi, New J. Chem.
2013, 37, 701–708.
[26] L. Zhang, J. M. Cole, J. Mater. Chem. A 2017, 5, 19541–19559.
[27] G. de Miguel, M. Ziółek, M. Zitnan, J. A. Organero, S. S. Pandey, S.
Hayase, A. Douhal, J. Phys. Chem. C 2012, 116, 9379–9389.
[28] a) M. W. Kryman, J. N. Nasca, D. F. Watson, M. R. Detty, Langmuir
2016, 32, 1521–1532; b) M. Ziółek, J. Karolczak, M. Zalas, Y. Hao, H.
Tian, A. Douhal, J. Phys. Chem. C 2014, 118, 194–205.
[29] a) H.-P. Lu, C.-Y. Tsai, W.-N. Yen, C.-P. Hsieh, C.-W. Lee, C.-Y. Yeh,
E. W.-G. Diau, J. Phys. Chem. C 2009, 113, 20990–20997; b) T.
Takeshita, T. Umeda, M. Hara, J. Photochem. Photobiol. Chem. 2017,
333, 87–91; c) J. H. Delcamp, Y. Shi, J.-H. Yum, T. Sajoto, E.
Dell’Orto, S. Barlow, M. K. Nazeeruddin, S. R. Marder, M. Grätzel,
Chem. – Eur. J. 2013, 19, 1819–1827; d) F. M. Jradi, X. Kang, D.
O’Neil, G. Pajares, Y. A. Getmanenko, P. Szymanski, T. C. Parker, M.
A. El-Sayed, S. R. Marder, Chem. Mater. 2015, 27, 2480–2487.
[30] J. Mei, Z. Bao, Chem. Mater. 2014, 26, 604–615.
FULL PAPER
Layout 1:
FULL PAPER
Three indenoquinaldine based Rajesh Bisht, Munavvar Fairoos Mele
Accepted Manuscript
unsymmetrical squaraine dyes were Kavungathodi, and Jayaraj
synthesized for dye-sensitized solar Nithyanandhan*
cells. The dye ISQ3 exhibited strong
absorption in NIR region, Page No. – Page No.
panchromatic IPCE and efficiency up
Indenoquinaldine Based
to 4.15 %.
Unsymmetrical Squaraine Dyes for
Near-Infrared Absorption:
Investigating the Steric and
Electronic Effects in Dye-sensitized
Solar Cells