You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/259743975

Emerging molecular design strategies of unsymmetrical phthalocyanines for


dye-sensitized solar cell applications

Article  in  RSC Advances · November 2013


DOI: 10.1039/c3ra45170d

CITATIONS READS

58 363

3 authors:

Varun Kumar Singh Ravi Kumar Kanaparthi


Ulsan National Institute of Science and Technology Central University of Kerala
21 PUBLICATIONS   295 CITATIONS    22 PUBLICATIONS   223 CITATIONS   

SEE PROFILE SEE PROFILE

Giribabu Lingamallu
Indian Institute of Chemical Technology
214 PUBLICATIONS   4,100 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dssc application s View project

Corrole Chemistry View project

All content following this page was uploaded by Ravi Kumar Kanaparthi on 17 January 2014.

The user has requested enhancement of the downloaded file.


RSC Advances
REVIEW

Emerging molecular design strategies of


unsymmetrical phthalocyanines for dye-sensitized
Cite this: RSC Adv., 2014, 4, 6970
solar cell applications
Varun Kumar Singh, Ravi Kumar Kanaparthi and Lingamallu Giribabu*

In recent years, dye-sensitized solar cells (DSSCs) have emerged as one of the possible solutions for the
global energy crisis. Among the various components of DSSCs, the sensitizer, which harvests solar
energy and injects electrons in to the semiconductor layer, plays a crucial role in achieving high
efficiency and durability of the cell. To date ruthenium(II) sensitizers exhibit high efficiencies (>11.5%), but
they are not so suitable for roof-top/commercial applications mainly because of their limited harvesting
capability, expensive ruthenium metal and low durability. In this context, various sensitizers which
include porphyrins, phthalocyanines and metal-free organic dye sensitizers have been developed and
some of them were found to exhibit enhanced efficiencies compared to classical ruthenium(II)
sensitizers. Man-made tetrapyrrolic systems, phthalocyanines (Pcs) have also been studied significantly
because of their unique thermal and electronic properties in the red and near-IR regions. Over the years,
Received 17th September 2013
Accepted 14th November 2013
the efficiency of Pc-based DSSC has improved to 6.1% by synthesizing various Pc derivatives and
optimizing fabrication parameters. However, in the present review article, only the recent developments
DOI: 10.1039/c3ra45170d
in Pc-based DSSCs have been documented in detail. This review also emphasizes different molecular
www.rsc.org/advances engineering approaches that the researchers developed for achieving higher efficiency.

1. Introduction
Meeting the world's growing energy demand is one of the most
signicant challenges in modern human society. Today, a major
Inorganic & Physical Chemistry Division, CSIR-Indian Institute of Chemical share of energy produced for mankind comes from fossil fuels.
Technology and CSIR-Network Institutes for Solar Energy (CSIR-NISE), Tarnaka, While the fossil fuels are declining with a greater rate,
Hyderabad 500007, India. E-mail: giribabu@iict.res.in; Fax: +91-40-27160921

Varun Kumar Singh was born in Ravi Kumar Kanaparthi was


Delhi in 1984. He completed his born in Yelavarru, a village near
undergraduate studies (B. Sc., Tenali, in 1982. Aer his
Chemistry) at Delhi University undergraduate studies (B. Sc.) at
in 2005. Then he obtained his S. V. R. M. College, he joined the
Masters degree (materials School of Chemistry, University
chemistry) from Jamia Millia of Hyderabad and received
Islamia University in 2007. At M. Sc. (Chemistry) and Ph.D.
present he is pursuing a Ph.D. at degrees in 2005 and 2011,
CSIR-IICT under the guidance of respectively. Subsequently, he
Dr L. Giribabu of Inorganic and moved to CSIR-Indian Institute
Physical Chemistry Division. His of Chemical Technology, Hyder-
research interests are tetrapyr- abad, India as a Post-Doctoral
roles such as porphyrins, phthalocyanines, corroles and related Fellow working with Dr L. Giribabu. Currently he is an assistant
macromolecules for dye-sensitized solar cell applications and professor at Kerala Central University. His research interests are
photophysical properties of electron donor–acceptor systems. broad absorbing sensitizers for dye-sensitized solar cells, photo-
voltaic and optoelectronic devices, and electron and energy
transfer reactions.

6970 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

greenhouse gases (SO2, CO2, NO2 etc.) are increasing as a result the ability of photogenerated minor carriers (for example elec-
of fossil fuels and causing severe global warming. So, there is a trons in a p-type material) to escape from one side of the device
great demand and urgency for new technologies that are before recombination with the major carriers. Thus, the minor
economically feasible and environmental friendly. Among the carriers' properties such as lifetime and diffusion length are
many renewable energy sources, solar energy is the best for essential parameters of the device. Even though the interfaces
electrical power generation as it is abundant, clean, and viable. are important in these p–n junction devices, the charge carrier
Therefore, if we could accomplish harvesting merely a tiny processes of photogeneration, separation and recombination
fraction of the solar energy that reaches the planet earth, we occur primarily in bulk material which determines the ultimate
would solve many of our problems not only in energy, but also power conversion efficiency. Therefore, bulk semiconductor
global, environmental and political issues. properties such as crystallinity and chemical purity oen
A solar cell is a device that directly converts light energy into control the efficiency of conventional solar cells and unfortu-
electrical energy and it works on the photovoltaic effect. The nately optimizing these properties is more expensive and
rst practical photovoltaic cell was developed in 1954 at Bell arduous. On the other hand, DSSCs belong to major carrier
Laboratories based on elemental silicon with 6% efficiency.1 devices in which electrons are found exclusively in one phase
Since then, a great variety of photovoltaic devices have been and holes in another phase. Charge carriers are generated at the
developed and some of them are already proven to show good interface between the electron-conducting and hole-conducting
power conversion efficiency. As a whole, solar cells are broadly phases via exciton dissociation. In other words, all important
categorized into three generations.2–4 The rst two generations charge carrier processes like photogeneration, separation and
are mainly based on crystalline and multi-crystalline silicon recombination occur primarily or exclusively at the interface.
(c-Si), amorphous silicon (a-Si), Cd–Te, Ga–As and CIGS etc. The Henceforth, the interfaces properties are of paramount impor-
materials that are associated with the rst two generations of tant and bulk properties are less critical. As a result, the
solar cells are either hazardous or not cost effective. In order to investigated materials do not need to be highly pure which
overcome the drawbacks of rst and second generation solar apparently reduces the cost of overall device fabrication. These
cells, scientists have developed excitonic solar cells, tandem differences between DSSCs and p–n junction solar cells provide
solar cells etc., which are categorized as third-generation solar an opportunity to work on various essential requirements of the
cells. The excitonic solar cells are broadly divided into nano- DSSCs for better understanding and improving the present
crystalline, dye-sensitized solar cells (DSSCs) and organic/poly- state of art and perhaps for these reasons it is believed that
mer solar cells. among many research areas, the study of DSSCs is a fruitful
Dye-sensitized solar cells are the most promising alternative topic.
to conventional solar cells emerging dramatically in recent O'Regan and Gratzel were the rst to report the DSSC
years. These DSSCs work under entirely different principles concept in 1991 with a remarkable efficiency of 7.1% based on
compared to the conventional solar cells and their mechanism natural photosynthesis.5 Over the last twenty years, the DSSCs
is close to natural photosynthesis. Unlike the conventional solar efficiency improved further and recently it has reached 12.3%.6
cells (silicon p–n junction cells), in DSSCs interfacial processes Among many other components of DSSC devices, the sensitizer
play a vital role. Conventional solar cells are also known as is known to play a crucial role in achieving high efficiency and
minor carrier devices in which their efficiency is determined by durability of the device. Extensively studied sensitizers are Ru(II)
polypyridyl complexes with the general structure ML2(X)2 where
L stands for 2,2-bipyridyl-4,40 -dicarboxylic acid, M is Ru, and X
Lingamallu Giribabu was born in represents thiocyanate (–NCS). In fact, the ruthenium complex
1969. He received his Ph.D. cis-RuL2(NCS)2, N3 dye has become a paradigm of heteroge-
degree in chemistry from Univer- neous charge-transfer sensitizers for mesoporous DSSCs.7 The
sity of Hyderabad under the high-energy conversion efficiency of this dye is generally
supervision of the late Professor attributed to the ultra-fast electron injection from dye to the
Bhaskar G. Maiya in 2000. TiO2 conduction band and a much slower reverse process. A
Subsequently, he moved to number of modications have been made in order to further
Central Queens Land University, improve the efficiency and durability of the device.8–12 Even
Australia and University of though ruthenium(II) complexes are found to be good in terms
Houston, USA for postdoctoral of power conversion efficiency, they have certain drawbacks:
assignments. He has worked as a ruthenium sensitizers are very expensive due to the rarity of the
senior scientist in Indian Insti- metal, several difficult synthetic protocols, lack of absorption
tute of Chemical Technology and lower molar extinction coefficients in the red region of the
since 2003. He is national co-ordinator of the dye-sensitized solar absorption spectrum oen limit their broad utility in the
cells area of CSIR-TAPSUN programme. His research interests manufacturing of prototype DSSCs panels and thereby
include development of new sensitizers for dye-sensitized solar cells, commercialization of the technology. In order to overcome
donor–acceptor systems, non-linear optical properties based on these problems, a great variety of alternative sensitizers has
tetrapyrrolic systems and photoelectrochemistry of been studied for the last two decades based on non-ruthenium
macromolecules. metal complexes, metal-free organic sensitizers, tetrapyrrolic

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6971
RSC Advances Review

compounds which include porphyrins, phthalocyanines and


corroles etc.13–28 The present review’s emphasis is mainly on
recent developments and emerging trends of Pc-based sensi-
tizers for DSSC applications. Although few review articles have
appeared on Pcs and their application in solar cells, for last two
years, several research papers have been published in the
literature which shows their importance in the DSSCs research.
Therefore, we reect to document these recent emerging trends
in such a way that this article enables the scientic community
to understand the current state of art and how to improve the
power conversion efficiency of the phthalocyanine sensitizers
further.

2. Working principle
The basic working principle and essential requirements of
DCCS are well documented in the literature.29–31 The DSSC is
composed of a dye/sensitizer anchored to nanocrystalline Fig. 2 Molecular structures of symmetrical phthalocyanine dyes.
mesoporous TiO2 lm on conducting glass (working electrode),
electrolyte, spacer in between two electrodes, platinum (Pt)
sputtered conducting glass (counter electrode) and a cell
the sensitizer, Fermi level of TiO2 (located near the conduction
sealant. Fig. 1 presents a schematic energy diagram of a DSSC
band level) and the redox potential of the redox couple (I/I3)
and working principle. The dye photosensitizer anchored to the
in the electrolyte solution. The photocurrent generated from the
TiO2 absorbs incident photon ux and is excited from the
DSSC is determined by the energy difference between the
ground state owing to its MLCT transition. Then the excited dye
HOMO and LUMO of the photosensitizer corresponding to
injects electrons into the conduction band of TiO2 and results in
the band gap of the semiconductor, TiO2. The smaller the
oxidation of the sensitizer. The injected electrons in the
HOMO–LUMO gap, the larger the photocurrent generation
conduction band of the TiO2 lm are transported through the
because of utilization of the long-wavelength region of the solar
surface or the interconnected TiO2 nanoparticles by diffusion
radiation. The LUMO of photosensitizer must be sufficiently
towards the back contact (conducting FTO glass) and conse-
negative with respect to the conduction band of TiO2 in order to
quently they reach the counter electrode through external load.
inject electrons effectively. In addition, substantial electronic
The oxidized photosensitizer is regenerated by accepting
coupling between the LUMO and the conduction band of TiO2 is
electrons from the redox couple (e.g. I/I3 in acetonitrile). I
also essential for effective electron injection thus to improve the
ion regenerates the dye photosensitizer to its ground state by
overall DSSC efficiency.
giving one electron and is apparently itself oxidized to I3 ion.
Then the oxidized I3 ion diffuses to the counter electrode
(Pt-coated FTO glass plate) and is reduced back to I ion. The
DSSC performance is predominantly judged based on four
3. Symmetrical phthalocyanine
energy levels: ground state (HOMO) and excited state (LUMO) of sensitizers
Phthalocyanines are tetrapyrrolic cyclic aromatic (18-p electron)
organic molecules, demonstrated as potential candidates for

Fig. 3 Molecular structures of symmetrical phthalocyanine dyes 5 and


Fig. 1 Energy levels and basic working principle of a typical DSSC. 6.

6972 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

Fig. 4 Molecular engineering strategies of phthalocyanines.

Fig. 6 Molecular structures of axial substituted phthalocyanine


Fig. 5 Molecular structures of axial substituted phthalocyanines dyes. sensitizers.

several practical applications that include optoelectronic, pho- them as “photovoltaic windows”: a red/near-IR absorbing
tocatalysis, chemical sensors, photodynamic therapy and pho- photovoltaic cell. In place of a window, one can place this
todegradation of organic pollutants etc. owing to their unique photovoltaic window which allows visible light to enter a
optical and electronic properties.32–37 In general, phthalocya- building while harvesting the red/near IR part of solar energy.
nines are well characterized by an intense Soret-band Moreover, with this kind of photovoltaic window, solar heating
(350 nm) and Q-band (650–700 nm) with molar absorptivity of buildings can be reduced or avoided completely. The thermal
(3) values that exceed 105 L mol1 cm1, which will be inter- and electronic properties of Pcs are suitable for the sensitization
esting features for solar-cell applications. Pcs exhibit good of wide band-gap semiconductor oxides such as TiO2, SnO2 etc.
absorption in the red and/or near-IR region of the solar spec- The optical, thermal and physical properties of Pcs can be tuned
trum and they can be tuned to be transparent over a large region by introducing various substituents at peripheral and non-
of the visible spectrum, thereby enabling the possibility of using peripheral positions as well as by introducing various metals at

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6973
RSC Advances Review

Fig. 7 Molecular structures of unsymmetrical phthalocyanines.

Fig. 9 Molecular structures of unsymmetrical tert-butyl phthalocya-


nines with different anchoring groups.

Fig. 8 Molecular structures of unsymmetrical tert-butyl phthalocya-


nines with different anchoring groups.

Fig. 10 Molecular structures of unsymmetrical phthalocyanines


its central cavity. However, the Pcs oen exhibit low power bearing different anchoring groups.
conversion efficiencies due to poor solubility in common
organic solvents, aggregation due to the planar Pc macrocycle
and lack of electron directionality in its excited state. In 1999,
for the rst time, Nazeeruddin et al. reported zinc(II) (Dye 1 and even though these Pcs sensitizers do not have any anchoring
3) and aluminium(III) (Dye 2 and 4) symmetrical Pcs bearing group they have shown reasonable IPCE values, 4–9%.
carboxylic acid and sulfonic acid anchoring groups to facilitate A few hyper-branched zinc(II) phthalocyanines have also
proper adsorption on to the TiO2 surface (Fig. 2).38 Among these been synthesized and nearly 67% IPCE was observed, but their
dyes the zinc(II) tetracarboxy phthalocyanine showed maximum overall power conversion efficiencies are still restricted to 1%.40
overall power conversion efficiency of 1% and the IPCE values The main reason for the low performance of these Pc-based
are between 10% and 45%.The low efficiencies of these Pcs were DSSC was found to be due to aggregation which promotes non-
attributed to deactivation of the excited Pc because of their high radiative deactivation of the dye excited states in aggregates.
aggregation originated from the planar conguration. Since Later He et al.41 reported a zinc(II) phthalocyanine bearing
then many efforts appeared in the literature which deal with tyrosine substituents (Dye 5) in order to increase solubility and
reducing the aggregation and thereby increasing the efficiency. to reduce aggregation at the semiconductor surface (Fig. 3).
Aranyos et al. have reported a series of Pcs with aryl groups However, this dye showed merely 0.54% efficiency with an IPCE
connected at the peripheral positions of the Pcs.39 Interestingly, of 24%, attributed to the fast recombination of injected

6974 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

Table 1 Photovoltaic data of various phthalocyaninesa

Sensitizer IPCEb (%) Jscb (mA cm2) Vocb (V) FFb hb (%)

Dye 1 43 5.40 0.416 — 1.00


Dye 2 13 1.10 0.466 — 0.42
Dye 3 30 0.48 0.453 — 0.77
Dye 4 10 0.60 0.382 — 0.14
Dye 5 24 2.25 0.360 0.670 0.54
Dye 6 —c 6.94 0.940 0.630 4.10e
Dye 7 60 —d —d —d —d
Dye 8 21 3.15 0.360 0.540 0.61
Dye 9 45 2.61 0.340 0.650 0.58
Dye 10 —c — — — 0.20
Fig. 11 Molecular structures of unsymmetrical phthalocyanines Dye 11 25 6.56 0.360 0.38 0.90
bearing organic chromophores. Dye 12 80 19 0.460 0.51 4.50
Dye 13 75 6.5 0.635 0.743 3.05
Dye 14 25 2.81 0.525 0.764 1.13
Dye15 80 7.60 0.617 0.75 3.52
Dye 16 47 5.63 0.557 0.75 2.35
Dye 17 9 0.90 0.55 0.72 0.40
Dye 18 56 4.80 0.61 0.74 2.20
Dye 19 65 6.80 0.613 0.74 3.10
Dye 20 16 1.44 0.611 0.75 0.67
Dye 21 65 9.37 0.605 0.725 4.10
Dye 22 75 7.37 0.609 0.74 3.28
Dye 23 50 5.88 0.587 0.75 2.55
Dye 24 55 6.80 0.576 0.69 2.64
Dye 25 67 9.15 0.600 0.72 3.96
Dye 26 40 6.86 0.584 0.72 2.87
Dye 27 70 7.01 0.610 0.73 3.26
Dye 28 21 3.70 0.520 0.73 1.40
Dye 29 36 7.07 0.563 0.75 2.97
Dye 30 85 7.67 0.559 0.76 3.24
Dye 31 40 5.25 0.541 0.73 2.07
Dye 32 40 4.96 0.543 0.74 1.98
Dye 33 50 4.8 0.58 0.77 2.10
Dye 34 86 10.4 0.63 0.70 4.60
Dye 35 85 1.31 0.585 0.76 6.13
Dye 36 50 8.77 0.584 0.71 3.63
Dye 37 65 12.8 0.61 0.68 5.30
Dye 38 82 13.7 0.613 0.70 5.90
Dye 39 —c 3.26 0.604 0.67 1.07
Dye 40 —c 2.33 0.504 0.75 0.89
Fig. 12 Molecular structures of sterically crowded unsymmetrical
phthalocyanine dyes.
a
Redox electrolyte is I/I3 in volatile organic solvent. b IPCE: Incident
photon to current conversion efficiency; Jsc: current density; Voc: open
circuit voltage; FF : Fill factor; h : Efficiency. c IPCE not reported.
d
Photovoltaic data not reported. e Counter electrode is PEDOT:PSS,
otherwise Pt counter electrode.
electrons and insufficient steric hindrance of the tyrosine
moiety. In another study, myristic acid was employed as a
co-adsorbent of aluminium(III) tetraphenoxy phthalocyanine the HNO3-treated photo-electrode retards back electron transfer
hydroxide for the sensitization of nanocrystalline TiO2.42,43 It at the interface with electrolyte and increases the amount of
was found that the co-adsorbent showed a dramatic decrease in dye. In a recent work, Jin and co-workers have reported zinc(II)
the aggregation of Pc on the nanocrystalline surface and as a phthalocyanine with eight octacarboxylic acid anchoring sites.
result the test cell efficiency was improved. Balaraju et al.44 have Although the aggregation is reduced, the DSSC device was not
used iron tetrasulfonic acid phthalocyanine (Dye 6) sensitizer found be efficient.45
and the corresponding device has shown an overall conversion In order to further improve the efficiency one has to mini-
efficiency of 4.1% by employing a PEDOT:PSS-coated FTO mize the aggregation of a Pc macrocycle through molecular
counter electrode instead of the usual Pt counter electrode and engineering. There are two methods to minimize the aggrega-
dilute HNO3 treatment of TiO2. The increase in power conver- tion of phthalocyanine macrocycles. One is the substitution of
sion efficiency of the DSSC based on nitric acid treatment for anchoring or aromatic groups at axial position(s) of the resident
the photo-electrode is mainly attributed to the increase in metal ion of the phthalocyanine. A second method is intro-
photocurrent. A comparative photovoltaic investigation of duction of bulky substituents as well as anchoring groups at
DSSCs using HCl-treated TiO2 photo-electrode indicates that peripheral positions i.e., ‘push–pull’ concept (Fig. 4).

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6975
RSC Advances Review

Fig. 16 Molecular structures of PTCDI, DCM and Rhodamine B.

Fig. 13 Molecular structures of sterically crowded unsymmetrical


phthalocyanine dyes.
Fig. 17 Molecular structures of BL315 and BL302.

however its overall efficiency was not impressive due to


desorption of sensitizer molecules on the TiO2 surface because
of weak the coordinating ability of the Ru(II) metal with the
pyridine ligand. Yanagisawa et al. have compared the sensiti-
zation properties of an axially substituted anchoring phthalo-
cyanine with that of peripherally substituted phthalocyanines
under identical test cell conditions.47 The axial substituted
sensitizer, Dye 8 has shown better efficiency (0.61%) than the
Fig. 14 Molecular structures of the co-sensitizers employed with the
peripherally substituted counterpart sensitizer Dye 9 (0.58%)
phthalocyanine sensitizers. and the decreased efficiency was attributed to reduced stability
of adsorbed Pcs on the TiO2 surface. As the axially substituted
Pcs are showing somewhat better efficiencies (owing to their
reduced aggregation and near-IR region absorption) Durrant,
Morandiera et al. have studied interfacial electron transfer
dynamics. They demonstrated that ultra-fast electron transfer
rates are not required for efficient electron injection yield.
Aerward Morandeira et al. studied the effect of co-adsorbents
on the device performance of axially substituted ruthenium(II)
phthalocyanies in order to understand the DSSC device physical
aspects. The effect of lithium cation (Li+) and co-adsorbent
chenodeoxycholic acid (CHENO) was studied systematically.48
From the previous studies it is known that the addition of Li+ is
expected to lower the conduction band edge of TiO2 and
decrease the photovoltage as well. However, an optimum
Fig. 15 Schematic representation of the DSSC with energy relay dye. concentration of Li+ increases the electron injection yield and
photocurrent of the DSSC device. Interestingly, it was observed
that Ru-phthalocyanine sensitizers showed a remarkable
decrease in the aggregation compared to symmetrical phtha-
4. Axially substituted phthalocyanines locyanines sensitizers reported till that time. On the other hand,
the co-adsorbent CHENO has also shown a signicant effect on
Nazeeruddin and co-workers tried to overcome the aggregation the device performance. The CHENO hinders the formation of
problem by synthesizing axially substituted Pcs and reported aggregates and thereby enhances the electron injection yield
ruthenium(II) phthalocyanine sensitizer, Dye 746 bearing bis(3,4- and ultimately increases the photocurrent response. Moreover,
dicarboxy)pyridine as anchoring groups (Fig. 5). The sensitizer 7 it is also found that recombination (dye cations with the elec-
had shown a promising 60% IPCE in the near IR region, trons in TiO2) becomes slower and regeneration of the dye

6976 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

cations by the redox electrolyte is faster which apparently substituent at one of the isoindoline units for better binding
enhanced the photocurrent. The reduced recombination can with the TiO2 surface. For their potential applications as
also due to increased hole–electron (phthalocyanine ring–TiO2 photosensitizers, the other three isoindoline units can be
surface) distance because of the unique geometry of CHENO functionalized with different substituents to make them more
which helps it to stand straight on the TiO2 surface thereby soluble, facilitate ease of purication and hinder aggregation
slowing down the dye cation–electron recombination. which has an undesirable effect in its DSSC application. Then
Torres and co-workers have utilized the axial position of the obtained precursors are reacted together and the resulting
titanium(IV) phthalocyanines, (Dye 10) for DSSC applications products can be carefully puried by column chromatography.
(Fig. 6).49 Unlike Gratzel and co-workers, they have connected In 2004, Yanagisawa and co-workers reported two unsym-
anchoring carboxy catechol at the axial position. By using metrical phthalocyanines with different metal ions at the
covalent bonds, desorption from the surface of nanocrystalline phthalocyanine's central cavity.51 The zinc-Pc-based sensitizer
TiO2 can be minimized. The solution absorption spectrum of showed an overall conversion efficiency of 0.03% whereas Ru(II)
Dye 10 exhibits an intense Q-band (lmax ¼ 702 nm, 3 ¼ 135 000 phthalocyanine sensitizer showed 0.4% efficiency. The low
dm3 mol1 cm1) and the dye is strongly adsorbed on to the efficiency of these phthalocyanines was attributed to the
TiO2 surface. However, its photovoltaic properties (IPCE ¼ 19% rupture in the conjugation between the macrocycle and the
and h ¼ 0.2%) are not impressive due to poor electron injection anchoring group which resulted in poor electron injection from
from the LUMO of the phthalocyanine to the TiO2 conduction the excited state of the phthalocyanine to the TiO2 conduction
band. Axially substituted Pcs such as Si(IV) phthalocyanines band. Later, in order to further improve the efficiency of DSSC
have also been reported and tested for their applicability in devices, unsymmetrical phthalocyanines have been designed
DSSCs. In this context, it is worthwhile to discuss a recent paper and developed based on the “push–pull” molecular architecture.
by Sellinger and Gratzel in which Dye 11 and Dye 12 have been The rst breakthrough appeared in 2007 by Giribabu and co-
shown to be very good candidates for DSSCs.50 The silicon workers in which they reported a very interesting unsymmet-
naphthalo–phthalocyanine hybrid sensitizer, Dye 12 showed rical phthalocyanine sensitizer (Dye 13) which gave the highest
excellent light harvesting efficiency (IPCE ¼ 80%) and high overall conversion efficiency of 3.05% using the iodide–triio-
photocurrent density of 19 mA cm2. Bulky tert-butyl substitu- dide electrolyte (Fig. 7).52–54 The Dye 13 sensitizer was designed
ents were added to the core periphery in order to avoid aggre- in such a way that it has three bulky tert-butyl groups to increase
gation and trihexylsiloxy groups were introduced as solubility and to decrease aggregation. The tert-butyl group also
coordinating ligands to the silicon center. The sensitizer Dye 12 acts as an electron-releasing (push) group and the succinic acid
has shown a Jsc of 19 mA cm2, open circuit voltage, Voc ¼ group serves as an electron-withdrawing group (pull) to anchor
0.46 V, ll factor (FF) ¼ 0.51 and overall efficiency of 4.5% under with the TiO2 surface. The Dye 13 sensitizer showed an excellent
standard irradiation conditions. But, the sensitizer Dye 11 has IPCE of 75% and its high efficiency was attributed to the excited
shown a lower efficiency of 0.9% which was attributed to state directionality thereby providing efficient electron injec-
insufficient driving forces for electron injection into the tion. By using a durable redox electrolyte (high boiling point
conduction band of TiO2 and regeneration of the oxidized dye solvents such as g-butyrolactone) the Dye 13 sensitizer showed
by the redox electrolyte. Although there are a few other Si(IV) 2% conversion efficiency and the device was found to be stable
phthalocyanines in the literature, their photovoltaic behaviour for at least 1000 hours.54 The Dye 13 sensitizer gives an IPCE of
is not efficient. 43% and overall conversion efficiency of 0.87% by using solid-
state organic hole transporter 2,20 ,7,70 -tetrakis(N,N-di-p-
methoxyphenylamine)-9,90 -spirobiuorene (spiro-MeOTAD)
5. Unsymmetrical phthalocyanine instead of liquid electrolyte. Subsequently, Giribabu and co-
sensitizers for DSSCs workers53 have designed and synthesized a stronger “push–pull”
derivative Dye 14 and its photovoltaic performance was
Based on many reports, symmetrical Pcs have indeed been compared with Dye 13. The new sensitizer Dye 14 possesses six
shown to be potential sensitizers for DSSC applications. donor butoxy groups at the a-position due to which the
Although their synthesis and purication are not a matter of absorption spectrum shied by 10 nm compared to Dye 13.
concern, the majority of them are inefficient in achieving high But, unfortunately, Dye 14 sensitizer gives only 1.13% overall
power conversion efficiencies mainly due to poor solubility in efficiency and 25% IPCE. The prime reason for the low efficiency
common organic solvents, aggregation and lack of direction- was poor electron injection quantum yield which also resulted
ality in the excited state. Hence, it was realized that in order to in the decreased Jsc value.
improve efficiency, one has to modify the molecular structure of In another study by Torres and co-workers, Dye 13 was
the existing symmetrical Pcs. By placing appropriate substitu- redesigned having only one carboxylic acid group instead of two
ents on the Pcs, one can tune desired properties such as solu- acid anchoring groups. The DSSC device fabricated using Dye
bility and aggregation. Moreover, it is indeed possible to break 15 sensitizer showed improved overall efficiency of 3.5% under
the symmetry of Pcs by placing two different functional groups standard irradiation conditions and the IPCE value at the
on the molecule. This can be achieved by mixed condensation Q-band absorption maxima was found to be 80%.55 The reason
of different substituted phthalonitriles and desired the Pc can for enhanced efficiency of Dye 15 is that the anchoring group
be synthesized in such a way that it should have an acid directly attached to the phthalocyanine macrocycle, which did

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6977
RSC Advances Review

not break the aromaticity as was the case in Dye 13. Later, efficiency (h ¼ 2.2%) because of its high electron injection yield
Giribabu and co-workers reported Dye 1656 sensitizer by rede- owing to improved directionality in the excited state.
signing the parent sensitizer Dye 13 based on the “push–pull” In another study, they demonstrated the effect of the
architecture and extended the p-conjugation concept. Malonic anchoring group on the stability of DSSC and the device
acid is the anchoring group in Dye 16. Although, the sensitizer performance.58,59 Pc Dye 21 was designed to possess two
Dye 16 exhibited better light harvesting ability (broader carboxylic acid anchoring groups in order to improve direc-
absorption) compared to the parent sensitizer, the overall tionality in the excited state and its efficiency in the presence of
conversion efficiency (2.35%) was not improved as expected; volatile and non-volatile electrolytes was compared with Dye 15
under similar test cell conditions Dye 13 has shown a conver- (Fig. 9). Dye 21 has shown improved conversion efficiencies of h
sion efficiency of 2.80%. The reason is that the IPCE of Dye 16 ¼ 3.33% and h ¼ 4.10% in the presence of volatile and non-
was only 47%, whereas Dye 13 has shown 71%. volatile electrolytes, respectively. Broad absorption and
Despite the fact that several researchers are working on the enhanced photocurrent response of Dye 21 are the governing
development of phthalocyanine-based sensitizers, the efficiency factors for its higher efficiency. This major breakthrough a
has reached little higher than 3% under standard conditions. couple of years aer the Dye 15 report had further sparked many
These sensitizers are highly attractive candidates because of researchers to design and synthesize Dye 22, Dye 23, Dye 24, Dye
their high absorption coefficients but still their efficiencies are 25 and Dye 26 sensitizers (by modifying anchoring groups) in
not enough for commercialization. Therefore, a great deal of order to explore the factors governing injection efficiency and
research still needs to be done for understanding the device the device performance has been compared with the parent Dye
physics and performance of the unsymmetrical sensitizers. 15. When a double bond was incorporated between the donor
macrocycle and the anchoring carboxylic acid (Dye 22), a small
decrease in IPCE was observed and the overall conversion effi-
ciency dropped down to h ¼ 3.28%. Then in Dye 23 and Dye 24,
6. Molecular design optimization: the cyanoacrylic acid group was used as the anchoring group.
study based on unsymmetrical tert- These sensitizers showed inferior device performance
butyl phthalocyanines compared to the parent dye. In Dye 23, a lower Jsc of 5.88 mA
cm2, Voc of 587 mV and h ¼ 2.55% was observed. The extended
Although in many studies co-adsorbents have been employed as conjugation between phthalocyanine and cyanoacrylic acid (Dye
aggregation controllers in DSSC devices none of them have 24) was also not helpful in achieving better efficiency ( Jsc ¼
proven helpful in improving the 3% efficiency. So, in later 6.80 mA cm2, Voc ¼ 576 mV and h ¼ 2.64%). However,
studies of electron injection, recombination lifetime of sensi- improved photovoltaic behaviour was observed in case of Dye 25
tizers has become a major concern in achieving high efficiency. sensitizers in which cyanoacrylic acid anchoring group was
In this context, a few studies by Torres and coworkers57 on replaced with a malonic acid anchoring group. Enhanced
unsymmetrical Pcs sensitizers for DSSC applications are note- photocurrent response ( Jsc of 9.15 mA cm2) and a slight
worthy to mention at this point of discussion. For example, they decrease in the photovoltage were observed. The decrease in
have designed and synthesized unsymmetrical zinc-phthalocy- photovoltage response was attributed to low electron lifetime
anines based on the “push–pull” molecular arrangement in and an increase in the number of carboxylic groups. The DSSC
which bulky tert-butyl donor groups and carboxylic acid device has shown overall conversion efficiency of 3.96% which
anchoring group are linked to the core phthalocyanine ring is higher than the Dye 15 sensitizer. However, when the
through different spacers. The DSSC devices fabricated with conjugation was extended between the malonic acid anchoring
these “push–pull” phthalocyanines reveal that the nature of the group and the macrocycle in Dye 26, a sudden breakdown in the
spacer unit and distance between the phthalocyanine core and Jsc ¼ 6.86 mA cm2, Voc ¼ 584 mV and thus overall conversion
the anchoring group play a signicant role in determining the efficiency 2.87% was observed. It was concluded that the
efficiency of the device. It was found that Dye 15 (without a sensitizers bearing malonic acid anchoring group could
spacer unit) has the highest overall power conversion efficiency possibly be the choice of anchoring group for future designs.
(h ¼ 3.52%) compared to Dye 17 (h ¼ 0.4%) possessing a exible The efficiencies exhibited by all the sensitizers revealed the
and unconjugated pentoxy group spacer (Fig. 8). Dye 20 pos- signicant involvement of high order directionality in the
sessing a phenoxy spacer has also shown merely 0.67% excited state of the dye which is a prerequisite for good elec-
conversion efficiency. Based on the results, it was proposed that tronic coupling between the excited state of the dye and the
the loss of directionality in the excited state led to poor electron titanium 3d-orbitals.
coupling of the donor with titanium 3d-orbitals and thus shows Nazeeruddin, Torres et al. examined the effect of an ethynyl
remarkably low efficiencies. Further it was found that due to the spacer between the carboxylic acid anchoring group and the
exible and non-directional nature of the oxygen linker (pentoxy phthalocyanine. It is expected that due to rigid planar spacer,
and phenoxy), there is a possibility that the sensitizer might not the phthalocyanine core will be perpendicular and closer to the
be able to stand straight on the TiO2 surface. However, on other TiO2 surface which enhances the electronic coupling between
hand Dye 18 and Dye 19 have shown comparable efficiencies the dye and Ti-3d orbital. In order to understand the ethynyl
with Dye 15. Even though Dye 18 shows faster recombination effect, Dye 27 and Dye 28 have been designed and synthesized
dynamics (compared to the other sensitizers), it showed better (Fig. 10).60 While Dye 27 gives a 3.26% overall conversion

6978 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

efficiency (IPCE ¼ 68%), Dye 28 exhibits only 1% efficiency presence of co-adsorbents like CDCA which demonstrates that
(IPCE ¼ 22%). The higher efficiency with one carboxyethynyl there is adequate room for reducing the aggregation tendency
spacer was attributed to the longer electron lifetime and higher of these molecules. For the rst time Imahori and coworkers
short circuit current density. However, the drop in the efficiency reported a substituted phthalocyanine (b,b0 -position) synthe-
on introduction of an additional carboxyethynyl spacer was due sized for DSSCs applications.63 They synthesized free base and
to the reduced Jsc owing to low driving force for charge injection zinc phthalocyanines with 4-tert-butylphenyl substituents
from excited dye into the TiO2 conduction band. covalently attached at six b,b0 -positions of the phthalocyanines
Gratzel and co-workers61 have synthesized two new phtha- and benzoic acid anchoring groups at the other two b,b0 -posi-
locyanine sensitizers Dye 29 and Dye 30 by employing a phos- tions. But the zinc phthalocyanine sensitized cell displayed
phinic acid anchoring group and their photovoltaic properties poor power conversion efficiency of 0.57% (Jsc ¼ 1.47 mA cm2;
were compared with Dye 15. It was observed that Dye 29 has Voc ¼ 0.54 V; FF ¼ 0.71) and 4.9% IPCE. The free base coun-
shown a lower IPCE of 36%, attributed to the poor adsorption of terpart of the same substituted phthalocyanine sensitizer has
Dye 29 on mesoporous semiconductor lm. It was found that shown even lower power conversion efficiency. Moreover, it was
the phosphinic acid anchoring groups enhance long term observed that the these sensitizer did not show any change in
photostability of the dye compared to the carboxylic acid device performance even in the presence of the CDCA co-
anchoring group. However, their overall efficiencies were found adsorbent which indeed proved the ability of such a molecular
to be slightly lower because of the decrease in the dye adsorp- design (bulky alkyl substitutions at b,b0 -positions) to reduce
tion on the semiconductor surface. However, the sensitizers, aggregation. Poor photovoltaic performance of these sensitizers
Dye 29 and Dye 30 have shown superior conversion efficiency to was attributed to the small driving force for electron injection
Dye 15 in the presence of co-adsorbent, CDCA. from the LUMO of free base/zinc phthalocyanines to TiO2 and
The phthalocyanines have high absorption coefficients in lack of proper electronic coupling between the dye excited state
the near-IR (NIR) region. However, they lack absorption in the and the TiO2 conduction band.
400–600 nm region. So, by synthesizing panchromatic Later Mori et al. designed a promising “push–pull” phthalo-
(absorption in the UV-Vis region) Pc sensitizers one can indeed cyanine sensitizer demonstrating the importance of three
improve their efficiencies even further.62 For this reason, Torres dimensional (3D) enlargement of the phthalocyanine molecular
and Nazeeruddin have developed conjugated panchromatic structure to suppress the aggregation completely. Three new
sensitizers by covalently attaching an organic chromophore at sensitizers Dye 18, Dye 33 and Dye 34 were synthesized and their
three positions which has absorption in the complimentary aggregation reducing order is Dye 34 > Dye 33 > Dye 18
region with the Pc core and the carboxylic acid group used for (Fig. 12).67 The Dye 34 sensitizer showed a maximum IPCE of
anchoring on to the TiO2 nanoparticles.63 In Dye 31, the 78%, high Jsc of 10.4 mA cm2 and record overall conversion
peripheral positions are substituted with triarylamine-termi- efficiency of 4.6%. Completely suppressed aggregation,
nated bisthiophene units and in Dye 32, hexylbisthiophene improved electron injection yield and increased directional
units are attached (Fig. 11). Unfortunately, both the sensitizers electron-transfer behaviour were found to be the primary
have shown only 40% IPCE and 3% overall conversion effi- reasons for the record efficiency of Dye 34. In an another study,
ciency by using a co-adsorbent. The reduced efficiencies of Bisquert et al.68 have studied the electron transfer dynamics
these dyes have been assigned to the aggregation and poor based on two sterically hindered phthalocyanines (Zn(II) and
electron injection efficiencies. free base) and compared their results with the benchmark N719
dye. These phthalocyanines possess electron-donating tert-
octylphenoxy groups and the electron-withdrawing acid anhy-
7. Sterically crowded unsymmetrical dride group served as an anchoring group. Both the sensitizers
phthalocyanine sensitizers have shown lower efficiencies when compared to N719 dye.
However, relatively zinc phthalocyanine gave slightly lower
The peripheral substitution of phthalocyanines is one of the efficiency than that of free base phthalocyanine, attributed to
promising ways to reduce planarity and thus aggregation and it the higher number of protons in free base phthalocyanine
is possible to manipulate the physicochemical properties as interacting with the semiconductor surface which lowers the
well. Several researchers have reported phthalocyanines with conduction band edge. It was concluded that increasing the
long alkyl chains or bulky substituents to suppress the inter- injection yield and liing the energy levels of the phthalocya-
molecular interaction and thereby aggregation.64–66 The most nine could help in improving the DSSC efficiency. The photo-
viable way of suppressing aggregation is to have bulky substit- voltaic data of all phthalocyanines are summarized in Table 1.
uents at adjacent positions like b,b0 -positions of the phthalo- In another study by following a similar structural design to
cyanine periphery. In fact a similar strategy has also been used Dye 34, Ragoussi and co-workers60 replaced the benzoic acid
before in designing various phthalocyanine sensitizers. For anchoring group with a carboxyethynyl anchoring group and
example, the good photovoltaic behaviour of Dye 13 and Dye 15 reported Dye 35 and Dye 36. The two dyes differed only in the
was due to a signicant decrease in aggregation and enhanced number of carboxyethynyl groups. Dye 35 with one ethynyl
directionality in the excited state owing to the presence of bulky spacer group has shown h ¼ 6.1% conversion efficiency, the
tert-butyl groups at the periphery of the phthalocyanine. It was highest efficiency reported so far using phthalocyanines and the
found that these dyes have performed even better in the IPCE also reached more than 85% at 700 nm. On the other

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6979
RSC Advances Review

hand, Dye 36 has shown h ¼ 3.54% and 55% IPCE. The high Voc comparatively hindered structure had indeed shown even lower
of Dye 35 was attributed to the enhanced electron lifetime and efficiency (0.89%). On the basis of theoretical calculations, it
decreased aggregation. On the other hand, the decreased effi- was stated that the charge delocalization of the LUMO extends
ciency of Dye 36 (compared to Dye 35) was due to its low to only one of the carboxylic anchoring groups, which may
injection efficiency, evident from the LUMO level (calculated decrease charge regeneration. Very recently, Giribabu and co-
using electrochemical techniques) of the Dye 36. These results workers have designed an unsymmetrical zinc(II) phthalocya-
suggests that the ethynyl group is the most optimal bridge nine with tert-butylthiophenyl groups at the non-peripheral
between the Pc-core and the carboxylic anchoring group which positions.72 Due to substitution at the non-peripheral positions
provides good electronic coupling between the Pc and TiO2. the phthalocyanine has strongly red shied absorption
However, the presence of two ethynyl spacer anchoring groups centered at 750 nm; the absorption in the near-infrared region
was found to decrease the efficiency mainly due to the can help to harvest a greater number of photons and thereby
decreased driving force for electron injection from the excited increase the cell efficiency. However, it was observed that this
dye into the TiO2 conduction band. sensitizer could harvest a very limited amount of solar energy
Mori, Kimura and co-workers have further explored the and the cell showed 0.40% efficiency.
molecular design strategy based on 3D-enlargement by
substituting methoxy groups at the peripheral positions and
reported Dye 3769 and Dye 38 (Fig. 13).70 Dye 37 has better 8. Co-sensitization effects in
solubility and electron donation ability to the phthalocyanine unsymmetrical phthalocyanine based
core. The sensitizer Dye 37 differed only by a methoxy group on DSSCs
the peripheral diphenylphenol moiety compared to Dye 34.
While the aromatic substituent in these two dyes were kept Over the years, photovoltaic performance of phthalocyanines
constant to optimize the electronic linker structure, in Dye 38 a has improved quite signicantly and in fact they have been
different adsorption site (compared to Dye 37) had been intro- shown to be one of the alternative sensitizers for commercial-
duced with a vision to improve the electronic coupling with izing DSSC technology. This came true only aer nding proper
TiO2. The sensitizer Dye 37 has shown high photocurrent solutions for reducing aggregation and enhancing the direc-
response, IPCE of 72% and 5.3% conversion efficiency. It was tionality in the dye excited state. However, there has been
found that the better electron donating ability of methoxy limited work on the improvement of phthalocyanine absorption
groups in 2,6-diphenyl-4-methoxyphenoxy of Dye 37 helped to in the 400–550 nm region, where the molar absorption coeffi-
increase injection yield and thus enhanced overall conversion cient of the phthalocyanine is minimal. Co-sensitization is a
efficiency. The structural optimization was carried out by convincing way to enhance the device performance through a
replacing the bulky 2,6-diphenyl-4-methoxyphenoxy groups combination of two or more dyes with complimentary absorp-
with 2,6-diisopropylphenoxy groups in Dye 38 to facilitate close tions characteristics on the semiconductor lms. For this
packing of dye molecules on the TiO2 surface which is reected purpose, the dyes have to be chosen in such a way that they
in its photovoltaic properties. The Dye 38 showed an improved should broaden the absorption, which essentially enhances the
power conversion efficiency of 5.9% (Jsc ¼ 13.7 mA cm2, Voc ¼ harvesting ability thus charge transfer efficiency and to reduce
0.613 V, FF ¼ 0.70) when compared to Dye 37. Based on charge the back transfer of electrons from the interconnected semi-
separation and injection efficiency, it was suggested a direct conductor oxide nanoparticle material to the sensitizing dye.
link from the carboxylic acid group to the phthalocyanine core The co-sensitization strategy has been reported before
was the appropriate choice in designing new sensitizers. employing ruthenium complex,73,74 metal-free organic dyes75,76
Molecular modeling studies conrmed that the undesirable and porphyrin systems77,78 to show enhanced photovoltaic
orientation of Dye 37 on TiO2 was the main reason for low performance compared to their single dye systems.
adsorption density and a decreased efficiency. This study The rst phthalocyanine–organic dye co-sensitization system
further revealed that the speculation “greater the steric for dye solar cell application was reported in 2007 by Torres and
hindrance, greater will be the efficiency” (keeping the anchoring co-workers.55 They demonstrated the enhanced photovoltaic
group same) may not always be correct. The efficiencies greatly response of Dye 15 co-sensitized with a reported organic dye
depend on the adsorption density of dye on the TiO2 surface as JK279,80 and observed enhanced device performance, Jsc ¼ 16.20
well. mA cm2, Voc ¼ 0.66 V, FF ¼ 0.72 and h ¼ 7.74% compared to
Recently, Giribabu, Filippo, Nazeeruddin and co-workers the individual phthalocyanine dye (3.5%) and JK2 (7.08%)
have reported new sterically hindered phthalocyanines based (Fig. 14). The photoresponse of the cocktail system has reached
on methoxy substituents having different patterns.71 The up to 700 nm with incident photon to current conversion effi-
sensitizer Dye 39 possess 3,4-dimethoxyphenyl substituents at ciency of 72% at 690 nm. The increased performance of the
the b-positions while Dye 40 has 2,6-dimethoxyphenoxy cocktail sensitizer was related to the high molar extinction
substituents as electron donors. However, with Dye 39 only coefficient of the dyes which would provide sufficient space on
1.07% power conversion efficiency has been achieved due to the semiconductor surface to allow absorption of another dye
lack of directionality in the excited state. Unfortunately, the having complimentary absorption.
sensitizer Dye 40 which was supposed to show better efficiency In 2012, Mori and co-workers69 co-sensitized Dye 37 with two
than that of Dye 39 because of decreased aggregation due to organic dyes, D10281,82 and D13183 resulting in an enhanced

6980 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

photovoltaic performance of the cocktail sensitizer. The DSSC tetra-tert-butyl zinc(II) phthalocyanine has been dissolved in the
device fabricated from cocktail sensitizer Dye 37/D102 exhibited electrolyte of a ruthenium polypyridine complex acceptor,
17% enhancement in the Jsc value while Dye 37/D131 cocktail unbound with no direct chemical bond or covalent linker,
sensitizer showed 28% enhancement in the Jsc compared to Dye achieving four-fold increase in the quantum yield for red
37 DSSC devices. An impressive IPCE of 80% even aer co- photons in dye-sensitized nanowire array solar cells.94 The close
sensitization showed the capability of good electron injection of packed nanoarray architecture was proven to be essential to
the cocktail sensitizers. There was no decrease noticed in the ensure special Forster Resonance Energy Transfer (FRET)
IPCE value in the near-IR region which also points towards zero between the donor and the acceptors. In an effort to increase
interaction between the two dyes. The Jsc of 15.3 mA cm2, Voc of panchromatic response, which is essential to enhance the light
0.63 V, FF of 0.64 corresponding to an overall efficiency of 6.2% harvesting capability thereby solar conversion efficiency,
was achieved with the Dye 37/D131 system while the Dye recently multiple energy relay dyes (with complementary
37/D102 cocktail system has shown Jsc of 13.9 mA cm2, Voc of absorption characteristics) in the electrolyte has been demon-
0.625 V, FF of 0.64 giving conversion efficiency h ¼ 5.6%. The strated. Two dyes DCM and Rhodamine B with complementary
authors proposed that the bulky substituents not only reduced absorption have been used with Dye 15 and high excitation
the aggregation but were also very effective in inhibiting the transfer efficiency with 35% increase in the photovoltaic
interaction between the two complimentary dyes because of an performance has been achieved. Fluorescence resonance energy
increase in the distance between their molecular frameworks. transfer from both relay dyes to the sensitizing dye was found to
be the dominant mechanism for increased light harvesting
system; however, the need to design energy relay dyes which are
9. Energy relay dyes: a strategy to more inert to quenching by the electrolyte was also
increase the light harvesting ability of highlighted.95
phthalocyanine dyes Very recently, two energy relay dyes BL302 and BL315 have
been used together with Dye 15 with a 65% increase in the
To have an efficient cocktail sensitizer based on the co-sensiti- efficiency of the DSSC device (Fig. 17).96 The excitation transfer
zation phenomena, it is necessary that the constituent dyes efficiency was limited to 70% only, which is low compared to the
should adsorb strongly on the TiO2 surface, transfer charge previously reported energy relay dyes. The majority of the losses
efficiently into the TiO2, and be effective at regeneration84 and in excitation transfer efficiency were found to be due to energy
charge injection.85–88 However, so far there have been limited transfer to the deabsorbed sensitizing dyes and static quench-
dyes which exhibit all these characteristics. Moreover, the co- ing of the energy relay dye by complex formation. These studies
sensitization methodology has a constraint regarding the further suggested that the energy relay dyes have fundamentally
number of anchoring sites on the TiO2 surface that are available different function mechanism and design rules than the
for the dye attachment. In this context, recently, long range sensitizing dyes; this architecture greatly expands the range of
energy transfer has been employed to increase the light har- dyes that can be used in DSSCs. Still better understanding of the
vesting ability in DSSCs (Fig. 15).89–91 energy relay dye design rules is anticipated to use them for
Recently, Cid et al. have reported an energy relay dye using a complimentary light harvesting systems in record devices.
highly uorescent chromophore, PTCDI, dissolved in electrolyte
and tested using Dye 15 (Fig. 16).55 A 28% increase in the
photocurrent response was observed due to the increase in 10. Conclusions
external quantum efficiency from the 400–600 nm region giving
a total 26% increase in the power conversion efficiency. Considering their optical and electrochemical properties,
However, no change in open circuit potential and ll factor of phthalocyanines have been successfully tested in various prac-
the relay device was found. The device efficiency was found to be tical DSSC applications. In the recent years, the potential of
2.55% and 3.21% in the absence and presence of PTCDI, phthalocyanines in dye-sensitized solar cells has been widely
respectively.92 In another work by Hardin and others,93 the explored. Various design strategies yielded stable, near-infrared
energy relay dye 4-(dicyanomethylene)-2-methyl-6-(4-dimethy- absorbing, less aggregating phthalocyanines with high internal
laminostyryl)-4H-pyran (DCM) was used with Dye 15 and the quantum efficiencies and moreover the directionality of the
overall power conversion efficiency increased from 3.5% to electron ow in the excited state has also been improved.
4.5%. In this study extremely high average excitation transfer However, new molecular design strategies for the construction
efficiency (>96%) with transparent TiO2 lms was demon- of highly efficient devices are still essential and a profound
strated. However, the external quantum efficiency of the relay understanding of the morphology of phthalocyanine sensitizers
dye was limited to 40% in the optimized device due to low is indeed necessary in order to control their photophysical
absorption of the relay dye. They have further showed that the properties. On the other hand, the use of energy relay dyes has
energy relay dyes should have a high molar absorption coeffi- emerged as a new concept for increasing the light harvesting
cient and very good solubility in the electrolyte to achieve an using these sensitizers through FRET in dye-sensitized solar
optimum DSSC architecture. cells. Though the energy relay dyes have paved the way to new
In a demonstration of the enhancement in the photovoltaic Pc-based cocktail systems for DSSCs, still greater understanding
performance of a DSSC device using energy relay dyes, of the energy loss mechanisms in such systems is required.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6981
RSC Advances Review

21 H. Imahori, T. Umeyama and S. Ito, Acc. Chem. Res., 2009, 42,


Acknowledgements
1809–1818.
VKS thanks to CSIR for a senior research fellowship. LG thanks 22 I. Radivojevic, A. Varotto, C. Farley and C. M. Drain, Energy
CSIR-NISE and Department of Science & Technology (DST), Environ. Sci., 2010, 3, 1897–1909.
Government of India for nancial support to carry out this work 23 M. V. Martinez Diaz, G. de la Torre and T. Torres, Chem.
under two major projects DST-EU (‘ESCORT’) and DST-UK Commun., 2010, 46, 7090–7108.
(‘APEX’) and also CSIR-TAPSUN programme. 24 M. G. Walter, A. B. Rudine and C. C. Wamser, J. Porphyrins
Phthalocyanines, 2010, 14, 759–792.
Notes and references 25 M. J. Griffith, K. Sunahara, P. Wagner, K. Wagner,
G. G. Wallace, D. L. Officer, A. Furube, R. Katoh, S. Mori
1 D. M. Chapin, C. S. Fuller and G. L. Pearson, J. Appl. Phys., and A. J. Mozer, Chem. Commun., 2012, 48, 4145–4162.
1954, 25, 676–677. 26 H. Imahori, T. Umeyama, K. Kurotobi and Y. Takano, Chem.
2 K. L. Chopra, P. D. Paulson and V. Dutta, Progr. Photovolt.: Commun., 2012, 48, 4032–4045.
Res. Appl., 2004, 12, 69–92. 27 M. K. Panda, K. Ladomenou and A. G. Coutsolelos, Coord.
3 M. Gratzel, J. Photochem. Photobiol., C, 2003, 4, 145–153. Chem. Rev., 2012, 256, 2601–2627.
4 O. Shevaleevskiy, Pure Appl. Chem., 2008, 80, 2079–2089. 28 M. V. Martinez Diaz, M. Ince and T. Torres, Monatsh. Chem.,
5 B. O'Regan and M. Grätzel, Nature, 1991, 353, 737–740. 2011, 142, 699–707.
6 A. Yella, H. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, 29 B. O'Regan and M. Gratzel, Nature, 1991, 353, 737–740.
M. K. Nazeeruddin, E. W. Diau, C. Y. Yeh, 30 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,
S. M. Zakeeruddin and M. Gratzel, Science, 2011, 334, 629– E. Muller, P. Liska, N. Vlachopoulos and M. Gratzel, J. Am.
634. Chem. Soc., 1993, 115, 6382–6390.
7 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker, 31 A. Hagfeldt, G. Boschloo, L. C. Sun, L. Kloo and
E. Muller, P. Liska, N. Vlachopoulos and M. Gratzel, J. Am. H. Pettersson, Chem. Rev., 2010, 110, 6595–6663.
Chem. Soc., 1993, 115, 6382–6390. 32 L. Valli, Adv. Colloid Interface Sci., 2005, 116, 13–44.
8 P. Wang, S. M. Zakeeruddin, J. E. Moser, M. K. Nazeeruddin, 33 A. Pailleret, F. Bedioui, in N4-Macrocyclic Metal Complexes,
T. Sekiguchi and M. Grätzel, Nat. Mater., 2003, 2, 402–407. ed. J. Zagal, F. Bedioui and J. P. Dodelet, Springer, New
9 T. Bessho, E. Yoneda, J.-H. Hum, M. Guglielmi, I. Tavernelli, York, 2006, pp. 363–423, ch. 8.
H. Imai, U. Rohlisberger, M. K. Nazeeruddin and M. Gratzel, 34 C. G. Classens, W. J. Blau, M. Cook, M. Hanack,
J. Am. Chem. Soc., 2009, 131, 5930–5934. R. J. M. Nolte, T. Torres and D. Wöhrle, Monatsh. Chem.,
10 L. Giribabu, C. V. Kumar, Ch. S. Rao, V. G. Reddy, 2001, 132, 3–11.
P. Y. Reddy, M. Chadrasekharam and Y. Soujanya, Energy 35 A. J. Duro, G. de la Torre, J. Barber, J. L. Serrano and
Environ. Sci., 2009, 2, 770–773. T. Torres, Chem. Mater., 1996, 8, 1061–1066.
11 L. Giribabu, T. Bessho, M. Srinivasu, C. V. Kumar, 36 P. S. Lai, P. J. Lou, C. L. Peng, C. L. Pai, W. N. Yen,
Y. Soujanya, V. G. Reddy, P. Y. Reddy, J.-H. Yum, M. Y. Huang, T. H. Young and M. J. Shieh, J. Controlled
M. Gratzel and M. K. Nazeeruddin, Dalton Trans., 2011, 40, Release, 2007, 122, 39–46.
4497–4504. 37 S. V. Rao, N. Venkatram, L. Giribabu and D. N. Rao, J. Appl.
12 T. Suresh, G. Rajkumar, S. P. Singh, P. Y. Reddy, A. Islam, Phys., 2009, 105, 053109.
L. Y. Han and M. Chandrasekharam, Org. Electron., 2013, 38 M. K. Nazeeruddin, R. H. Baker, M. Grätzel, D. Wöhrle,
14, 2243–2248. G. Schnurpfeil, G. Schneider, A. Hirth and N. Trombach,
13 S. Ito, H. Miura, S. Uchida, M. Takata, K. Sumioka, P. Liska, J. Porphyrins Phthalocyanines, 1999, 3, 230–237.
P. Comte, P. Pechy and M. Gratzel, Chem. Commun., 2008, 39 V. Aranyos, J. Hjelm, A. Hagfeldt and H. Grennberg,
5194–5196. J. Porphyrins Phthalocyanines, 2001, 5, 609–616.
14 K. Hara, Z.-S. Wang, T. Sato, A. Furube, R. Katoh, 40 L. Yong, L. Peifen, Y. Xingzhong, J. Lu and P. Zhonghua, RSC
H. Sugihara, Y. Dan-oh, C. Kasada, A. Shinpo and S. Suga, Adv., 2013, 3, 545–558.
J. Phys. Chem. B, 2005, 109, 15476–15482. 41 J. He, G. Benkö, F. Korodi, T. Polı́vka, R. Lomoth,
15 R. Chen, X. Yang, H. Tian and L. Sun, J. Photochem. B. Akermark, L. Sun, A. Hagfeldt and V. Sundström, J. Am.
Photobiol., A, 2007, 189, 295–300. Chem. Soc., 2002, 124, 4922–4932.
16 G. Zhang, H. Bala, Y. Cheng, D. Shi, X. Lv, Q. Yu and 42 Y. Amao and T. Komori, Langmuir, 2003, 19, 8872–
P. Wang, Chem. Commun., 2009, 2198–2200. 8875.
17 R. K. Kanaparthi, J. Kandhadi and L. Giribabu, Tetrahedron, 43 T. Komori and Y. Amao, J. Porphyrins Phthalocyanines, 2003,
2012, 68, 8383–8393. 7, 131–136.
18 L. Giribabu, R. K. Kanaparthi and V. Velkannan, Chem. Rec., 44 P. Balaraju, M. Kumar, M. S. Roy and G. D. Sharma, Synth.
2012, 12, 306–328. Met., 2009, 159, 1325–1331.
19 L. Giribabu and R. K. Kanaparthi, Curr. Sci., 2013, 104, 847– 45 L. Jin, Z. L. Ding and D. J. Chen, J. Mater. Sci., 2013, 48, 4883–
855. 4891.
20 L. Giribabu, K. Sudhakar and V. Velkannan, Curr. Sci., 2012, 46 M. K. Nazeeruddin, R. Humphry-Baker, M. Grätzel and
102, 991–1000. B. A. Murrer, Chem. Commun., 1998, 719.

6982 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014
Review RSC Advances

47 M. Yanagisawa, F. Korodi, J. He, L. Su, V. Sundstrom and 66 S. Eu, T. Katoh, T. Umeyama, Y. Matano and H. Imahori,
B. Akermark, J. Porphyrins Phthalocyanines, 2002, 6, 217–224. Dalton Trans., 2008, 5476–5483.
48 A. Morandeira, I. Lopez-Duarte, B. O’Regan, M. V. Martinez- 67 S. Mori, M. Nagata, Y. Nakahata, K. Yasuta, R. Goto,
Diaz, A. Forneli, E. Palomares, T. Torres and J. R. Durrant, M. Kimura and M. Taya, J. Am. Chem. Soc., 2010, 132,
J. Mater. Chem., 2009, 19, 5016–5026. 4054–4055.
49 E. Palomares, M. V. Martinez Diaz, S. A. Haque, T. Torres and 68 E. M. Barea, J. Ortiz, F. J. Paya, F. F. Lazaro, F. Fabregat-
J. R. Durrant, Chem. Commun., 2004, 2112–2113. Santiago, A. Sastre-Santos and J. Bisquert, Energy Environ.
50 B. Lim, G. Y. Margulis, J. Yum, E. L. Unger, B. E. Hardin, Sci., 2010, 3, 1985–1994.
M. Gratzel, M. D. McGehee and A. Sellinger, Org. Lett., 69 M. Kimura, H. Nomoto, N. Masaki and S. Mori, Angew. Chem.,
2013, 15, 784–787. 2012, 124, 4447; Angew. Chem., Int. Ed., 2012, 51, 4371.
51 M. Yanagisawa, F. Korodi, J. Bergquist, A. Holmberg, 70 M. Kimura, H. Nomoto, H. Suzuki, T. Ikeuchi, H. Matsuzaki,
A. Hagfeldt, B. Akermark and L. Sun, J. Porphyrins T. N. Murakami, A. Furube, N. Masaki, M. J. Griffith and
Phthalocyanines, 2004, 8, 1228–1235. S. Mori, Chem.–Eur. J., 2013, 19, 7496–7502.
52 P. Y. Reddy, L. Giribabu, C. Lyness, H. J. Snaith, 71 L. Giribabu, V. K. Singh, T. Jella, Y. Soujanya, A. Amat, F. De
Ch. Vijaykumar, M. Chandrasekharam, M. Lakshmikantam, Angelis, A. Yella, P. Gao and M. K. Nazeeruddin, Dyes Pigm.,
J. H. Yum, K. Kalyanasundaram, M. Grätzel and 2013, 98, 518–529.
M. K. Nazeeruddin, Angew. Chem., Int. Ed., 2007, 46, 373–376. 72 V. K. Singh, P. Salvatori, A. Amat, S. Agrawal, F. De Angelis,
53 L. Giribabu, Ch. Vijay Kumar, V. Gopal Reddy, P. Yella M. K. Nazeeruddin, N. V. Krishna and L. Giribabu, Inorg.
Reddy, Ch. Srinivasa Rao, S.-R. Jang, J.-H. Yum, Chim. Acta, 2013, 407, 289–296.
Md. K. Nazeeruddin and M. Grätzel, Sol. Energy Mater. Sol. 73 L. Y. Han, A. Islam, H. Chen, C. Malapaka, B. Chiranjeevi,
Cells, 2007, 91, 1611–1617. S. F. Zhang, X. D. Yang and M. Yanagida, Energy Environ.
54 L. Giribabu, C. V. Kumar, M. Raghavender, K. Somaiah, Sci., 2012, 5, 6057–6060.
P. Y. Reddy and P. Y. Rao, J. Nano Res., 2008, 2, 39–48. 74 R. Y. Ogura, S. Nakane, M. Morooka, M. Orihashi, Y. Suzuki
55 J. J. Cid, J. H. Yum, S. R. Jang, M. K. Nazeeruddin, and K. Noda, Appl. Phys. Lett., 2009, 94, 073308.
E. M. Ferrero, E. Palomares, J. Ko, M. Grätzel and 75 Y. Chen, Z. Zeng, C. Li, W. Wang, X. Wang and B. Zhang, New
T. Torres, Angew. Chem., Int. Ed., 2007, 46, 8358–8362. J. Chem., 2005, 29, 773–776.
56 L. Giribabu, C. Vijaykumar, P. Y. Reddy, J. H. Yum, 76 D. Kuang, P. Walter, F. Nuesch, S. Kim, J. Ko, P. Comte,
M. Grätzel and M. K. Nazeeruddin, J. Chem. Sci., 2009, 121, S. M. Zakeeruddin, M. K. Nazeeruddin and M. Gratzel,
75–82. Langmuir, 2007, 23, 10906–10909.
57 J. J. Cid, M. G. Iglesias, J. H. Yum, A. Forneli, J. Albero, 77 T. Bessho, S. M. Zakeeruddin, C. Y. Yeh, E. W. G. Diau and
E. M. Ferrero, P. Vázquez, M. Grätzel, M. K. Nazeeruddin, M. Gratzel, Angew. Chem., Int. Ed., 2010, 49, 6646–6649.
E. Palomares and T. Torres, Chem.–Eur. J., 2009, 15, 5130– 78 M. S. Kang, S. H. Kang, S. G. Kim, I. T. Choi, J. H. Ryu,
5137. M. J. Ju, D. Cho, J. Y. Lee and H. K. Kim, Chem. Commun.,
58 M. G. Iglesias, J. J. Cid, J. H. Yum, A. Forneli, P. Vázquez, 2012, 48, 9349–9351.
M. K. Nazeeruddin, E. Palomares, M. Grätzel and 79 J. H. Yum, S. R. Jang, P. Walter, T. Geiger, F. Nuesch, S. Kim,
T. Torres, Energy Environ. Sci., 2011, 4, 189–194. J. Ko, M. Gratzel and M. K. Nazeeruddin, Chem. Commun.,
59 M. Garcia-Iglesias, J. Yum, R. H. Baker, S. M. Zakeeruddin, 2007, 4680–4682.
P. Pechy, P. Vazquez, E. Palomares, M. Gratzel, 80 S. Kim, J. K. Lee, S. O. Kang, J. Ko, J. H. Yum, S. Fantacci,
M. K. Nazeeruddin and T. Torres, Chem. Sci., 2011, 2, F. De Angelis, D. Di Censo, M. K. Nazeeruddin and
1145–1150. M. Gratzel, J. Am. Chem. Soc., 2006, 128, 16701–16707.
60 M. Ragoussi, J. J. Cid, J. Yum, G. Torre, D. D. Censo, 81 L. S. Mende, U. Bach, R. H. Baker, T. Horiuchi, H. Miura,
M. Gratzel, M. K. Nazeeruddin and T. Torres, Angew. S. Ito and M. Gratzel, Adv. Mater., 2005, 17, 813–815.
Chem., Int. Ed., 2012, 51, 4375–4378. 82 W. H. Howie, F. Claeyssens, H. Miuraand and L. Perter,
61 I. Lopez-Duarte, M. Wang, R. H. Baker, M. Ince, M. Victoria J. Am. Chem. Soc., 2008, 130, 1367–1375.
Martinez-Diaz, M. K. Nazeeruddin, T. Torres and M. Gratzel, 83 T. Dentani, Y. Kubota, K. Furnabiki, J. Jin, T. Yoshida,
Angew. Chem., Int. Ed., 2012, 51, 1895–1898. H. Minoura, H. Miura and M. Matsui, New J. Chem., 2009,
62 L. Giribabu, V. K. Singh, C. Vijaykumar, Y. Soujanya, 33, 93–101.
P. Y. Reddy and M. L. Kantam, Sol. Energy, 2011, 85, 1204– 84 J. N. Clifford, E. Palomares, M. K. Nazeeruddin, M. Gratzel
1212. and J. R. Durrant, J. Phys. Chem. C, 2007, 111, 6561–6567.
63 M. Ince, F. Cardinali, J. Yum, M. V. Martinez-Diaz, 85 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,
M. K. Nazeeruddin, M. Gratzel and T. Torres, Chem.–Eur. E. Muller, P. Liska, N. Vlachopoulos and M. Gratzel, J. Am.
J., 2012, 18, 6343–6348. Chem. Soc., 1993, 115, 6382–6390.
64 M. Durmus and T. Nyokong, Polyhedron, 2007, 26, 2767– 86 Y. Tachibana, J.-E. Moser, M. Gratzel, D. R. Klug and
2776. J. R. Durrant, J. Phys. Chem., 1996, 100, 20056–20062.
65 N. B. McKeown, S. Makheed, K. J. Msayib, L. L. Ooi, 87 S. A. Haque, Y. Tachibana, R. L. Willis, J.-E. Moser,
M. Hellowell and J. E. Wareen, Angew. Chem., Int. Ed., M. Gratzel, D. R. Klug and J. R. Durrant, J. Phys. Chem. B,
2005, 44, 7546. 2000, 104, 538–547.

This journal is © The Royal Society of Chemistry 2014 RSC Adv., 2014, 4, 6970–6984 | 6983
RSC Advances Review

88 Y. Tachibana, M. K. Nazeeruddin, M. Gratzel, D. R. Klug and M. Gratzel and M. D. McGehee, Nat. Photonics, 2009, 3,
J. R. Durrant, Chem. Phys., 2002, 285, 127–132. 406–411.
89 B. E. Hardin, E. T. Hoke, C. Peters, T. Geiger, F. Nuesch, 93 B. E. Hardin, J.-H. Yum, E. T. Hoke, Y. C. Jun, P. Pechy,
M. Gratzel, M. K. Nazeeruddin and M. D. McGehee, in T. Torres, M. L. Brongersma, M. K. Nazeeruddin,
Long Range Forster Energy Transfer to Increase Light M. M. Gratzel and D. McGehee, Nano Lett., 2010, 10, 3077–3083.
Absorption in Dye-Sensitized Solar Cells Materials Research 94 K. Shankar, X. Feng and C. A. Grimes, ACS Nano, 2009, 3,
Society Fall Meeting, Materials Research Society, Boston, 788–794.
MA, 2008. 95 J.-H. Yum, B. E. Hardin, E. T. Hoke, E. Baranoff,
90 J. H. Yum, B. E. Hardin, S. J. Moon, E. Baranoff, F. Nuesch, S. M. Zaeeruddin, M. K. Nazeeruddin, T. Torres,
M. D. McGehee, M. Gratzel and M. K. Nazeeruddin, Angew. M. D. McGehee and M. Gratzel, ChemPhysChem, 2011, 12,
Chem., Int. Ed., 2009, 48, 9277–9280. 657–661.
91 E. T. Hoke, B. E. Hardin and M. D. McGehee, Opt. Express, 96 G. Y. Margulis, B. Lim, B. E. Hardin, E. L. Unger, J. H. Yum,
2010, 18(4), 3893–3904. J. M. Feckl, D. F. Rohlng, T. Bein, M. Gratzel, A. Sellinger
92 B. E. Hardin, E. T. Hoke, P. B. Armstrong, J. H. Yum, and M. D. McGehee, Phys. Chem. Chem. Phys., 2013, 15,
P. Comte, T. Torres, J. M. Frechet, M. K. Nazeeruddin, 11306–11312.

6984 | RSC Adv., 2014, 4, 6970–6984 This journal is © The Royal Society of Chemistry 2014

View publication stats

You might also like