You are on page 1of 12

View Article Online / Journal Homepage / Table of Contents for this issue

PCCP Dynamic Article Links

Cite this: Phys. Chem. Chem. Phys., 2012, 14, 15963–15974


Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

www.rsc.org/pccp PERSPECTIVE
Adsorption of organic dyes on TiO2 surfaces in dye-sensitized solar cells:
interplay of theory and experiment
Chiara Anselmi, Edoardo Mosconi, Mariachiara Pastore, Enrico Ronca and
Filippo De Angelis*
Received 28th August 2012, Accepted 11th October 2012
DOI: 10.1039/c2cp43006a

First-principles computer simulations can contribute to a deeper understanding of the dye/


semiconductor interface lying at the heart of Dye-sensitized Solar Cells (DSCs). Here, we present
the results of simulation of dye adsorption onto TiO2 surfaces, and of their implications for the
functioning of the corresponding solar cells. We propose an integrated strategy which combines
FT-IR measurements with DFT calculations to individuate the energetically favorable TiO2
adsorption mode of acetic acid, as a meaningful model for realistic organic dyes. Although we
found a sizable variability in the relative stability of the considered adsorption modes with the
model system and the method, a bridged bidentate structure was found to closely match the
FT-IR frequency pattern, also being calculated as the most stable adsorption mode by
calculations in solution. This adsorption mode was found to be the most stable binding also for
realistic organic dyes bearing cyanoacrylic anchoring groups, while for a rhodanine-3-acetic acid
anchoring group, an undissociated monodentate adsorption mode was found to be of comparable
stability. The structural differences induced by the different anchoring groups were related to the
different electron injection/recombination with oxidized dye properties which were experimentally
assessed for the two classes of dyes. A stronger coupling and a possibly faster electron injection
were also calculated for the bridged bidentate mode. We then investigated the adsorption mode
and I2 binding of prototype organic dyes. Car–Parrinello molecular dynamics and geometry
optimizations were performed for two coumarin dyes differing by the length of the p-bridge
separating the donor and acceptor moieties. We related the decreasing distance of the carbonylic
oxygen from the titania to an increased I2 concentration in proximity of the oxide surface, which
might account for the different observed photovoltaic performances. The interplay between
theory/simulation and experiments appears to be the key to further DSCs progress, both
concerning the design of new dye sensitizers and their interaction with the semiconductor and
with the solution environment and/or an electrolyte upon adsorption onto the semiconductor.

1. Introduction
usually TiO2, composed of nanometer-sized particles. The
The answer to the growing demand of environmentally charge hole which is concomitantly created in the dye after
sustainable energy resources at the planetary level might lie excited state charge injection is then transferred to a redox
in the ability to capture and utilize solar energy for a sustainable couple in a liquid electrolyte or to a solid hole conductor
development on a large scale. In this scenario, dye-sensitized (see Scheme 1).
solar cells (DSCs) represent a particularly promising approach Of the three main DSCs constituting materials, i.e. the dye,
to the direct conversion of sunlight into electrical energy at low the semiconductor oxide and the redox shuttle, the chemical
cost and with high efficiency.1–4 nature and structure of the dye is by far the subject which has
In DSCs, a dye sensitizer absorbs the solar radiation and been more vastly investigated, with a general target of increasing
transfers the photoexcited electron to a wide band-gap semi- the dye molar extinction coefficient and shifting the dye absorption
conductor electrode consisting of a mesoporous oxide layer, towards the near-IR region, thus enhancing the overlap between
the solar emission and the dye absorption spectrum and eventually
achieving higher DSCs photocurrents. A further important
Computational Laboratory for Hybrid/Organic Photovoltaics sensitizer’s feature is the matching of ground and excited state
(CLHYO), Istituto CNR di Scienze e Tecnologie Molecolari, Via
Elce di Sotto 8, I-06123, Perugia, Italy. E-mail: filippo@thch.unipg.it; oxidation potentials with the redox shuttle potential and the
Fax: +39 075 585 5606; Tel: +39 075 585 5523 semiconductor conduction band, respectively.5,6 These energetic

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15963
View Article Online

Also the orientation and packing of adsorbed dyes on the


semiconductor surface strictly depends on the binding motif,
possibly affecting the rate and effectiveness of parasitic
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

recombination reactions.35–37 Finally, the sensitizer’s anchoring


group should provide stable grafting of the dye onto the semi-
conductor surface, thus leading to long-term device stability.38–40
In an effort to design new and more stable and efficient dyes,
it is then clear how fundamental is to disclose the anchoring
mode showed by the most commonly employed sensitizers
and by the newly designed ones. This understanding could
contribute to individuate and optimize many aspects that
affect the DSCs performance, such as the solvent used for
the sensitization process,41,42 the use of co-adsorbent species,43,44
Scheme 1 Schematic representation of the charge flow and of the and the electrolyte composition.45,46 From an experimental
energy levels of a dye-sensitized solar cell. standpoint, information on the dye binding modes at the dye-
sensitized semiconductor interface are conventionally accessed by
requirements rule the DSCs kinetics and are needed to allow vibrational spectroscopy [Fourier Transform Infra Red (FT-IR)
for efficient electron injection into the semiconductor manifold spectroscopy and Surface-Enhance Raman Spectroscopy
of unoccupied states and subsequent regeneration of the (SERS)].47,48 These techniques can provide structural details
oxidized dye.7,8 about adsorption modes as well as information on the relative
Ru(II)-polypyridyl complexes have been primarily employed orientation shown by surface-adsorbed molecules. Photoelectron
as dye sensitizers.9–13 The remarkable performance of the N3 Spectroscopy (PES) has also been successfully employed to
dye,9 of its doubly protonated analogue (N719)10 and of the characterize the dye/oxide interface for a series of organic
so-called black dye11,12 had a central role in significantly dyes.49–52 The analysis of the PES spectra yields information
advancing the DSCs technology, with solar to electric power on the molecular and electronic structures at the interface, along
efficiencies exceeding 11%.10,13 In the last few years, a flourishing with basic indications of the dye coverage and of the distance of
family of new generation Ru(II)-dyes have been designed and selected atoms from the surface. More recently, X-ray reflecto-
synthesized to provide higher molar extinction coefficient or metry has also been employed to unravel the orientation of dye
peculiar supramolecular interactions compared to N719, thus molecules adsorbed onto TiO2 surfaces.53
enhancing the DSCs overall stability and/or efficiency.14–18 Fully The anchoring mechanism of the largely employed carboxylic
organic sensitizers have been developed as metal-free dyes because acid group onto the semiconductor surface can be exemplified
of their increased molar extinction coefficient, compared to referring to the coordination modes of the carboxylate fragment
Ru(II)-dyes, spectral tunability and reduced environmental (COO ) to metal ions; there are basically three possible coordina-
impact.19,20 When employing the most common I /I3 electrolyte, tion modes, i.e. monodentate, chelated and bridged bidentate
organic dyes have delivered very high photovoltaic efficiencies, (see Scheme 2).54 An empirical rule, derived by Deacon and
exceeding 10%.21,22 Finally, functionalized donor–acceptor Zn(II)- Philips,55 correlates the difference between the asymmetric and
porphyrins have recently emerged as a new class of sensitizers with symmetric stretching frequencies (Dnas) of COO to the type of
high performance potential, due to the extended absorption coordination. If the measured Dnas for the adsorbed species is
spectrum in the red, up to near IR region.23–25 larger than that measured for the neat salt, a monodentate
From the semiconductor side, TiO2 is essentially established coordination is supposed to take place; otherwise if Dnas is similar
as the most performing metal oxide in DSCs, both in the form or smaller than that of the corresponding salt, a chelated or a
of the commonly employed sintered nanoparticles and nano- bridged bidentate adsorption mode is inferred.
tubes,26,27 although research on alternative oxides such as ZnO Identification of the COO– symmetric and asymmetric
and SnO2 is being actively pursued in various laboratories.28–31 vibrational frequencies is usually possible for simple carboxylic
The I /I3 redox couple in an organic solvent has been acids, but often realistic dyes show a plethora of overlapping
considered for many years the standard high-performance absorptions in the 1400–1600 cm 1 range, which hinder identifi-
redox shuttle in DSCs,10,13 even if the field has very recently been cation and/or assignment of the diagnostic carboxylic modes.
revolutionized by Co(II)/Co(III) redox couples, which allowed for Along with experimental studies, computer simulations
demonstration of photovoltaic efficiency close to 13%.25 have contributed to a deeper understanding of the electronic
As mentioned above, the dye sensitizer plays a crucial role in
efficient DSCs. Apart from the discussed spectral and energetic
requirements, a salient feature of efficient dyes is the presence
of suitable functional groups able to strongly bind to the
semiconducting oxide surface. The anchoring group should
indeed coincide, or be very close, to the dye acceptor moiety,
where unoccupied electronic states are spatially confined. This
promotes electronic coupling between the donor levels of the Scheme 2 Graphical representation of the three possible carboxylate
excited dye and the delocalized acceptor levels of the semiconductor bonding modes: (a) monodentate ester-like, (b) bidentate chelating,
conduction band, and helps the charge injection process.32–34 (c) bidentate bridging.

15964 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012
View Article Online

structure and optical properties of new dyes and to unravel the oxygen via hydrogen-bonding. We considered here two different
intimate dye–semiconductor interactions characterizing the monodentate binding modes (M1 and M2) and a bridged
crucial DSCs heterointerface.56,57 A number of theoretical bidentate (BB) anchoring, see Fig. 1. In M1 the –OH group
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

studies dealing with dye adsorption onto TiO2 have been interacts with a surface oxygen which is not directly bound to
published, see ref. 58–77 for a list of few representative the adsorption-involved Ti atom, while in M2 the interaction is
examples, starting from the pivotal work by Vittadini et al. instead established with the adjoining surface oxygen, see
concerning formic acid adsorption on the TiO2 anatase (101) optimized geometries in Fig. 1.
surface.77 In some cases, calculations showed that for organic For the description of the TiO2 semiconductor we used two
dyes bearing a carboxylic acid as an anchoring group, the different approaches, i.e. a cluster and a periodic model. The
preferred adsorption mode was bidentate bridging, with one cluster model has the advantage of allowing for straight-
proton transferred to a nearby surface oxygen, while the forward use of hybrid exchange–correlation functionals and
monodentate anchoring is usually predicted to be less stable, continuum solvation models. The periodic approach allows
although some dependency of the relative stability of these two for a more rigorous description of the surface properties,
anchoring modes on the employed computational methodology which is free from finite-size effects. We use two stoichiometric
has been outlined.76 As a matter of fact, different stabilities can (TiO2)n clusters, with n = 38 and 82, which have been
be obtained for the e.g. monodentate or bidentate anchoring demonstrated to reproduce to a good degree of accuracy the
modes depending on the employed level of theory.76 DOS of periodic TiO2 surfaces.67 The two cluster models are used
To overcome the individual limitations of experimental and to gauge the stability of the various adsorption modes, allowing
computational techniques, here we propose an integrated for the comparison of hybrid and GGA DFT functionals.
experimental/computational strategy which combines FT-IR We then resort to a (TiO2)32 periodic slab for comparison
measurements with DFT computer simulations to individuate purposes. It was recently shown that TiO2 anatase slabs having
the most energetically favorable adsorption mode for a carboxylic thickness similar to our clusters or periodic models, nicely
acid probe, followed by calculation of the IR spectra and reproduce the electronic structure of thicker films.69 In all cases,
structure validation by comparison with experimental data. In our TiO2 models expose the majority (101) surfaces.
particular, we focus on the adsorption of prototypical dyes,
exploiting the typical diagnostic symmetric and asymmetric
Cluster calculations. Geometry optimizations were carried
stretching frequencies of the carboxylic group. Considering the
out in the gas phase with the ADF program package78
complexity of the most efficient organic dyes, we chose the
employing the PBE exchange–correlation functional79 with a
structural simplicity of acetic acid as a meaningful model for
TZVP(DZVP) basis set for Ti (H, C, N, O, S). We have also
realistic organic dyes. Acetic acid, Scheme 2 with R = Me, offers
taken into account the solvent role as well as the effect of
the advantage of having the most commonly employed carboxylic
calculation level on the relative stability of optimized structures by
acid group of realistic dyes, coupled to a single methyl group,
performing single point calculations in the gas phase and in water
substantially simplifying the IR spectrum.
solution using the B3LYP80 and PBE exchange–correlation func-
In the first part of our paper, we thus investigate, by means
tionals together with a large 6-311G** basis set, of comparable
of combined FT-IR spectroscopy and DFT calculations the
quality to that used for geometry optimizations. Solvation effects
possible anchoring modes onto the TiO2 surface showed by
were taken into account by means of the C-PCM solvation
acetic acid. We consider both monodentate and bidentate
model81 as implemented in Gaussian 03 (G03).82
binding modes, evaluating their relative energies at various
levels of theory, along with simulating the IR spectra of the
surface adsorbed species. The FT-IR spectra of acetic acid Periodic calculations. Periodic DFT calculations have been
vapors as well as of its partly deuterated analogues carried out within the generalized gradient approximation
(CH3COOD) adsorbed on TiO2 have been measured to build (GGA) using the PBE exchange–correlation functional. The
up a correlation between theoretical and experimental data. In Car–Parrinello (CP)83 code was used, as implemented in the
the second part of the manuscript, we compare the results Quantum-Espresso package.84 Electron–ion interactions were
obtained for the acetic acid probe in relation to previous described by ultrasoft pseudopotentials with electrons from O,
results on the adsorption of realistic organic dyes on TiO2 N and C 2s, 2p; H 1s; and Ti 3s, 3p, 3d, 4s shells explicitly
and the implication of the different adsorption modes for included in the calculations. Plane-wave basis set cutoffs for the
DSCs functioning, including injection/recombination kinetics smooth part of the wave functions and the augmented density
and dye–electrolyte interactions. were 25 and 200 Ry, respectively. The dye molecules were

2. Models, methods and experimental


2.1 Computational details
The carboxylic group is initially considered to be protonated,
as required by the dye solution charge neutrality. The acidic
proton can then be either transferred to the oxide surface
(both in a bidentate and a monodentate anchoring mode) or Fig. 1 Optimized geometrical structures of the investigated adsorp-
be retained on the dye and possibly interact with a surface tion modes of acetic acid onto the TiO2 surface.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15965
View Article Online

adsorbed on one side only of the slab and the vacuum between Table 1 Experimental IR frequencies for AcOH and AcOD
the top of the molecules and the adjacent slab was B7 Å.
Vibrational modes (cm 1) AcOH AcOD
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

2.2 l-FTIR measurements nas COO 1558 1556


nsym COO 1424 1428
All the experimental spectra were recorded using a m-FTIR dCH3 1335 1335
instrument consisting of a JASCOs FTIR 4100 spectrometer,
equipped with a liquid nitrogen cooled MCT detector coupled
with an IMV-4000 optical microscope. Measurements were (1420–1412 cm 1), whereas other authors85 assigned this
performed in the transmission mode (through a Cassegrain wavenumber range (1433–1426 cm 1) to the symmetric carboxylate
16X objective) on a micro diamond cell, scanning an energy stretching.
range of 6000–600 cm 1 and with a resolution of 4 cm 1. The To possibly elucidate the involvement of the carboxyl
spectra were recorded using 5000 scans; background correc- proton in binding to TiO2 (e.g. in M1 or M2 structures) we
tion was adopted by means of a spectrum collected on the repeated the experiment with the monodeuterated acetic acid
empty micro diamond cell. TiO2, CH3COOH and CH3COOD CH3COOD (AcOD) as an adsorber. The measured IR spectrum
with purity >99.5% were all purchased from Aldrich and (Fig. 2) essentially overlaps with the one already seen for acetic
used as received. Three samples of TiO2 powder were heated acid. Both the asymmetric and symmetric carboxylate stretchings
for 1 h at 500 1C. They were then put in contact, at room remain localized at the same wavenumbers of AcOH and so
temperature and atmospheric pressure, with vapours of pure does the signal at about 1330 cm 1. Such similarity between
AcOH, d4-AcOH and AcOD respectively. After 20 h a m-FTIR AcOH and AcOD suggests that probably no H (or D)
spectrum of each sample was recorded directly on powders. belonging to acidic OH (or OD) groups is directly interacting
with the surface-adsorbed carboxylic group, as would be the
case for M1 and M2.
3. Results and discussion For the interacting AcOH/TiO2 system, we then calculated
the optimized geometries and the relative energies of the M1,
3.1 Acetic acid on TiO2
M2 and BB adsorption modes, see Fig. 1 and Tables 2 and 3.
Acetic acid (AcOH) vapors were adsorbed at room temperature Different calculation levels were employed and the effect of
and atmospheric pressure (the standard temperature and pressure water solvent on the adsorption stability was also taken into
conditions in which DSCs are normally assembled) onto a pre- account on the geometries optimized in vacuo. Considering the
heated (500 1C) titania (anatase) nanopowder. After 20 h exposure PBE/DZ(TZ)VP level used for geometry optimization on the
a m-FT-IR spectrum was recorded, as shown in Fig. 2. cluster-adsorbed systems, at least two structures should be
In the recorded AcOH spectrum, the main signals recognizable taken into account in terms of relative energy, i.e. BB and M2.
in the wavenumber range of COO– adsorption (apart from that Our data suggest M2 to be the most stable structure, as
of residual water at 1630 cm 1) are due to the asymmetric and previously found for acetic acid adsorption on the (110)
symmetric carboxylate stretchings as well as to methyl group TiO2 surface in ref. 87. On the same (ADF) optimized cluster
modes (see Table 1). geometries, the relative stabilities calculated by the PBE functional
The frequencies experimentally found for AcOH are referred in by ADF, CP and G03 programs in vacuo are totally consistent,
the literature85 to be characteristic of bidentate species, with the despite the different basis sets (Slater, Gaussian and Plane Waves),
bands at 1558 and 1424 cm 1 due to asymmetrical and symmetrical showing a similar description of the dye–TiO2 interaction by all
COO– stretching, respectively. While a large consensus supports such computational approaches. M2 is predicted to be the most
the asymmetrical carboxylate stretching assignment, the signal stable structure, with BB lying very close in energy and M1 much
at 1430 cm 1 is in some cases86 referred to dCH3 vibrations higher. B3LYP data in vacuo are consistent with the corresponding
PBE data, suggesting a marginal effect of the functional.
Moving to the water solution environment (e B 78) leads
BB to be calculated as the most stable structure both with
B3LYP and PBE, with M2 and M1 lying, respectively,
1.3 (1.4) and 6.5 (4.2) kcal mol 1 higher in energy with the
B3LYP (PBE) functional. Our results are in line with previous
calculations, showing a sizable sensitivity of the most stable
adsorption mode for carboxylic acids and organic dyes onto
the (101) TiO2 surface to the presence of explicit solvent (water
or acetonitrile) molecules.76,88,89
Surprisingly, a totally different stability order is calculated if
one considers the CP optimized geometries, employing a PBC
TiO2 description (PBE/PWs entries in Table 2), which shows
the M1 adsorption mode to be the most stable structure, in line
with the results of Vittadini et al. on the prototype formic acid
system.77 As it can be noticed, slightly different geometries are
Fig. 2 m-FTIR spectra of AcOH (red) and AcOD (blue) vapors calculated for cluster and periodic approaches, see Table 3,
adsorbed onto the TiO2 surface. although the differences seem to be quite small (except the OH

15966 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012
View Article Online

Table 2 Relative stabilities (kcal mol 1) of the various investigated adsorption modes of AcH on TiO2. Both results obtained with a cluster (CL)
and a periodic (PBC) TiO2 model are reported. PBE/DZ(TZ)VP data were obtained by the ADF code and PBE/PWs by the CP code; PBE/
6-311G*** and B3LYP/6-311G** have been obtained by the Gaussian code
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

Geometry optimization Energy evaluation BB M1 M2


CL/ADF PBE/DZ(TZ)VP Vac 0.0 +10.1 4.1
PBE/DZ(TZ)VP PBE/PWs Vac 0.0 +8.9 4.1
PBE/6-311G** Vac 0.0 +7.3 1.6
Solv 0.0 +4.2 +1.4
B3LYP/6-311G** Vac 0.0 +10.8 0.9
Solv 0.0 +6.5 +1.3
PBC/CP PBE/PWs Vac 0.0 12.1 10.8
PBE/PWs

Table 3 Main optimized geometrical parameters (Å) of the investi- three intense signals at 1543, 1432 and 1351 cm 1, to be compared
gated systems on the TiO2 cluster by the ADF program (CL-ADF) to experimental signatures at 1558, 1435 and 1340 cm 1, respec-
and the structure optimized by the CP program in Periodic Boundary
Conditions (PBC-CP). See Fig. 1 for labels
tively. The assignment of the 1543 cm 1 feature to the asymmetric
COO– stretching is straightforward, while the two modes
Structure Geometry Ti1–O1 (Ti1/2–O2) C–O1 C–O2 O1–H H–O3/4 calculated at 1432 and 1351 cm 1 result from the admixture
BB CL-ADF 2.06 (2.16) 1.29 1.26 — 0.98 of the symmetric COO– stretching and of the dCH3 mode.
PBC-CP 2.07 (2.10) 1.30 1.27 — 0.97 Thus a clear assignment of these two features cannot be made,
M1 CL-ADF 2.15 1.25 1.31 1.04 1.55 also considering their inverted intensity pattern compared
PBC-CP 2.15 1.24 1.32 1.02 1.63
M2 CL-ADF 2.06 1.29 1.27 1.31 1.13 to the experiment. We note, however, that the separation
PBC-CP 2.18 1.26 1.32 1.04 1.51

distances in M2) to explain the substantial effect on the


stability order. We have no further data to comment on this
marked difference but, based on the similarity of the energetics
by CP, ADF and G03 at fixed geometry, we speculate that this
effect might be due to a different description of proton
interactions with the TiO2 surface by the CP method rather
than by a cluster vs. periodic TiO2 difference.76
This seems to be confirmed by the different M2 structures
calculated by the ADF and the CP code, whereby ADF
predicts the proton to be transferred to the surface while CP
retains it on the dye (notice the different O1–H and H–O3/4
distances in Table 3).
Considering the large variability of computed stabilities
with the employed geometry optimization method and since
based on the relative energies alone it was not possible to
clearly discriminate among the possible AcOH adsorption
modes (the inherent accuracy of most DFT methods is com-
parable to some of the calculated energy differences), for each
of the considered structures, we calculated the IR vibrational
frequencies at the same PBE/DZ(TZ)VP level of theory used
for geometry optimization. A summary of the main IR peaks
and their assignment is reported in Table 4, while in Fig. 3 we
report a comparison between the experimental and calculated
spectra for BB, M1 and M2.
Looking at the calculated frequencies, the BB adsorption
mode well reproduces the three main spectral features, showing

Table 4 Calculated (unscaled) and experimental IR frequencies for


AcOH adsorbed onto TiO2

Frequencies BB M1 M2 Exp.
Fig. 3 Comparison between the experimental (red) and calculated
nas COO 1543 1614 1748 1558 (blue) IR spectra of AcOH on TiO2 for the BB, M1 and M2 adsorption
nsym COO 1432 1425 1377 1424 modes. The calculated spectra have been rescaled so that the high
dCH3 1351 1335 1416 1335 energy features have the same intensity as the experimental band at
D(nas nsym) 111 189 371 134
1558 cm 1.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15967
View Article Online

between the asymmetric and symmetric COO stretching found


experimentally is 138 cm 1, reasonably comparing with the
111 cm 1 difference found between the asymmetric stretching
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

and the band of mixed character calculated at 1432 cm 1.


Based on the comparison between the experimental and
simulated IR spectra, Fig. 3, the presence of the monodentate
species M1 and M2 (either dissociated or undissociated) can be
ruled out, due to the significant discrepancies found in their IR
spectra compared to the experimental data. The experimental
spectra show no trace of peaks of the characteristic carbonyl
stretching belonging to the M2 adsorption mode (B1760 cm 1).
For M1 the feature at B1550 cm 1 is missing and a new feature
at ca. 1250 cm 1 appears, which is not found experimentally. So
even if the carbonyl stretching (1614 cm 1) was overlapping with Fig. 4 Molecular structures of L0, rh-L0, D5L2A1 and D5L2A3.
the residual water peak (1630 cm 1) this mode can be excluded.
It has to be noted that frequency assignment in these two cases is injected electrons into the TiO2 CB. Low electron lifetime
further complicated by the admixture of several modes, including values, due to high recombination rates, therefore determine
OH modes, so that for M1 and M2 the mode assignments low Voc values.
in Table 2 should be taken with some care, although the We investigated the possible anchoring modes onto the TiO2
listed frequencies are those displaying highest intensity in the surface of the two representative L0 and rh-L0 organic
investigated energy range. dyes,50,93 considering both monodentate and bidentate binding.
Summarizing, our combined experimental and computa- The optimized geometries are reported in Fig. 5, while relative
tional modeling of acetic acid adsorption on TiO2 points at energies of BB, M1 and M2 adsorption modes are reported in
the BB structure as the most likely AcOH adsorption mode, Table 5. The data in Table 5 follow the trend outlined above for
consistent with the calculated energetics in solution. acetic acid adsorption, whereby at the PBE/DZ(TZ)VP/vac
level of theory, M2 is slightly more stable, by 1.5 kcal mol 1,
3.2 Organic dyes on TiO2: electron injection and than the bidentate structure, BB, while the M1 adsorption mode
recombination with the oxidized dye seems to be strongly destabilized (ca. +10 kcal mol 1) with
respect to BB. Single point calculations on the optimized
As mentioned above, investigating the possible anchoring structures in water solution, however, always indicate the
modes of dye sensitizers to the TiO2 surface is of crucial bidentate configuration as the preferred one, giving the stability
importance as the bonding type and the extent of electronic ordering BB > M2 > M1. Although all the three investigated
coupling between the dye excited state and the semiconductor structures would be compatible with the PES data for the
unoccupied states can directly influence the cell performances. related D5L2A1 dye,49 which indicates that the molecule lies
An emblematic case is represented by the D5L2A190,91 and with the donor (TPA) group pointing out from the surface and
D5L2A350 dyes, which we recently investigated.37 The two the acceptor CN group close to the TiO2, our calculated
dyes only differ in the anchoring unit, namely a conjugated stabilities are suggestive of a bidentate coordination mode for
cyanoacrylic acid in the former and non-conjugated rhodanine-3-
acetic acid in the latter (see molecular structures in Fig. 4). The
loss of electron conjugation between the donor-linker moiety and
the anchoring –COOH group passing from the cyanoacrylic to
the rhodanine-3-acetic acid was related to the drastic drop in the
photovoltaic efficiency of D5L2A3 (B2%)50 compared to
D5L2A1 (B6%).90,91 Wiberg et al.92 reported that the two
different sensitizers showed unitary electron injection efficiency
but substantially different rates of charge recombination between
the photo-injected electrons and the oxidized dye, with estimated
losses of 80% and 50% of injected electrons for D5L2A3 and
D5L2A1, respectively, within 1.4 ns.92
Recombination between TiO2-injected electrons and the
oxidized dye or the electrolyte directly influences the charge
density into the semiconductor and thus the open-circuit
voltage of the cell. By definition, at open circuit the flux of
the injected electrons equals that of the recombining electrons,
fixing the Fermi level and thus the output voltage of the
device. The Fermi level is determined by the CB edge potential
and by the electron density into the TiO2, n, which is in turn
proportional to the rate of the electron injection and to the Fig. 5 Optimized structures of the L0 (upper panel) and rh-L0 (lower
electron lifetime, t, namely the average time spent by the panel) dyes on TiO2.

15968 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012
View Article Online

Table 5 Relative energies (kcal mol 1) of BB, M1 and M2 adsorption (approximated here by the dye LUMO) and the TiO2 manifold
modes of L0 and rh-L0 on (TiO2)38. From ref. 37. PBE results in water of unoccupied states.61 We thus calculated the partial density
(Solv* entry) solution were obtained with the 6-311G* basis set
of dye unoccupied states (PDOS) for the interacting BB and
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

Energy evaluation BB M1 M2 M1 adsorption modes of the L0 dye. We recall that according


to the Newns–Anderson model, the entity of the dye LUMO
L0 dye
PBE/DZ(TZ)VP Vac 0.0 +10.9 1.5 broadening when interacting with the semiconductor is an
PBE/6-311G** Vac 0.0 +10.8 +2.6 estimate of the intervening electronic coupling.75,97 The results,
Solv* 0.0 +8.4 +7.4 reported in Fig. 6, clearly show a larger broadening (and
B3LYP/6-311G** Vac 0.0 +14.9 +3.6
Solv 0.0 +10.9 +7.7
asymmetry) for the L0 dye LUMO PDOS adsorbed in the BB
rh-L0 dye than in the M1 mode.
PBE/DZ(TZ)VP Vac 0.0 +7.1 1.9 The dye LUMO broadening, estimated as the full width at
PBE/6-311G** Vac 0.0 +6.7 5.6 half maximum of the Lorentzian distributions75 reported at
Solv* 0.0 +0.4 +5.0
B3LYP/6-311G** Vac 0.0 +8.1 +2.9 the bottom of Fig. 6, is ca. 0.05 eV larger for the BB than for the
Solv 0.0 +0.4 +7.7 M1 adsorption mode, reflecting the stronger dye–semiconductor
interaction characteristic of the BB structure.70

L0 and analogous dyes bearing a cyanoacrylic anchoring


3.3 Effect of the dye adsorption mode: recombination with the
group.65
electrolyte
Similar remarks can be made for rh-L0: at the PBE/
DZ(TZ)VP/vac level of optimization, the bidentate structure Recombination between TiO2-injected electrons and oxidized
is the most stable adsorption mode. Although small energy species in the electrolyte is a main loss channel in DSCs,
differences are calculated at some levels of theory, these data leading to reduced Voc values.98 Despite the accepted knowledge
indicate on overall that also for this type of dyes the bidentate that the adsorbed dyes act as an isolating layer, keeping the
coordination is favored, and therefore we conclude that rh-L0 reduced electrolyte far from the direct contact with the oxide
as well as the related D149 and D10236 dyes lie almost flat with surface,99–101 some authors102–106 suggested that particular atoms
an inclination of about 45 degrees with respect to the TiO2 or chemical groups can also provide binding sites for I2 (I3 ),
plane, as also found in previous works.94,95 increasing its concentration close to the TiO2 surface and thus
Differences in the recombination kinetics between TiO2- accelerating the recombination process. Evidence of stable
injected electrons and the dye cations are strongly dependent, Ru-complex–I2 adducts has been reported by Tuikka
upon other factors, from the spatial separation of the charge et al.,106 which isolated N3–I2 crystals. O’Regan et al.102 found
hole localized on the oxidized dye, which can be assimilated to that the replacement of two oxygen ligand atoms with two
the HOMO of the neutral dye, and the TiO2 surface. We recall sulfur atoms in a Ru(II)-dye caused a 2-fold increase in the
indeed that the electronic coupling between the donor (TiO2) recombination rate. This was ascribed to the stronger tendency
and acceptor (oxidized dye) states can be represented in terms of ethylthioether compared to ethylether to bind I2. An
of an exponential function of the spatial separation r between extensive investigation on the effects of dye molecular structure
the donor and the acceptor. Considering the calculated different and electrolyte composition on the recombination dynamics has
orientation of the L0 and rh-L0 dyes with respect to the TiO2 been reported by Miyashita et al.,103 who examined the photo-
surface, determined by the different anchoring unit, and recalling voltaic performances of a series of organic- and Ru(II)-dyes.
that the dye HOMOs are mainly localized on the donor groups, These authors consistently found lower electron lifetimes for the
a different spatial separation between the dye donor groups and metal-free based DSCs compared to those employing Ru dyes.
the TiO2 surfaces can be outlined for L0 and rh-L0. Looking at The lower Voc values were interpreted as arising from an
the more stable BB structures reported in Fig. 5, the dye donor increased I3 concentration in the proximity of the oxide surface
group/TiO2 distance can be roughly estimated to be 8–10 Å for yielding higher recombination rates. Interestingly, among
L0 and 6–8 Å for the rh-L0 dye. By extrapolating these data for the organic sensitizers, those having alkyl chains and larger
the longer D5L2A1 and D5L2A3 dyes examined by Wiberg molecular size showed longer lifetimes, possibly due to the
et al.,92 we calculate a dye donor group/TiO2 separation of formation of a compact aggregate layer able to protect the
14–16 Å and 9–11 Å for the two dyes, respectively. These data TiO2 surface from the I3 approach.
are consistent with the measured differences in the recombina- Various computational investigations on dye–electrolyte
tion rates, whereby a closer proximity of the charge hole interactions were recently reported,107–113 including a time-
localized on the dye to the semiconductor surface would lead domain ab initio study of charge relaxation and recombination
to higher recombination rates. Our model is also in agreement in dye-sensitized TiO2 by Prezhdo and coworkers.114 To
with recent data by Long et al.,96 which showed an increased contribute to the understanding of dye–electrolyte inter-
recombination rate between injected electrons and the oxidized actions, we recently investigated the adsorption mode of various
sensitizer for the D205 dye, bearing the rhodanine-3-acetic acid, prototype organic dyes on TiO2, along with I2 binding to the
compared to the C218 dye, having a cyanoacrylic acid group. various electron-donating sites in the molecule. This was done
It is also interesting to exploit our data for the L0 dye to for a representative group of sensitizers, selected among those
check whether the specific dye adsorption mode, i.e. bridged examined by Miyashita et al.103 Here we discuss results for two
bidentate vs. monodentate (e.g. BB vs. M1) can induce differ- coumarin dyes (NKX2587 and NKX2697), which differ by the
ences in the electronic coupling between the dye excited state number of thiophene rings in the bridge (one and three,

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15969
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05. View Article Online

Fig. 7 Optimized structures of NKX2587 and NKX2697 adducts


with I2.

Having established the affinity of sulfur, oxygen and nitrogen


atoms in the standalone dyes to bind I2, we moved to model the
dye/semiconductor interface, with relation to the anchoring
modes of the dyes and the effective distances of the iodine
from the TiO2 surface. We thus carried out a Car–Parrinello
molecular dynamics simulation on the NKX2587 dye adsorbed
onto the (TiO2)82 cluster in vacuo, starting with the preferred
dissociative bridged bidentate adsorption mode and thermalizing
the system at 350 K. We notice that the presence of explicit
solvation could possibly alter the adsorbed dye dynamics.59 Also
the rather short simulation time (4.5 ps) is not sufficient to
capture possible slow dye rearrangements but we believe this
time slice to be sufficient to locally sample the potential energy
surface around the calculated minimum energy structure. Our
results showed that the BB adsorption mode was maintained
throughout the simulation time, and that the dye remained on
average perpendicular to the TiO2 surface, without significant
geometrical distortions with respect to the optimized adsorption
geometry at 0 K. Hence no significant change in the average
distance of the I2 binding sites from the surface was observed
compared to the optimized geometry. As shown in Fig. 8, the
Fig. 6 Top: partial density of states (PDOS) calculated for the L0 variation in the distance of the carbonylic oxygen from the
dye LUMO anchored to TiO2 in a BB (blue) and M1 (red) surface during simulation time of 4.5 ps is about 1 Å, going
adsorption modes. Bottom: Lorentzian broadening of the dye LUMO from a minimum value of 9.7 to a maximum of 10.8 Å, with an
PDOS. average value of 10.4 Å.
We also optimized the structure of the O–I2 adduct of TiO2-
respectively), see Fig. 7. In a recent work,104 we have shown that adsorbed NKX2587 and, as displayed in Fig. 9, this structure
in the case of two organic dyes containing, respectively, one and shows negligible geometrical rearrangements of the O–I2
two ethylenedioxythiophene (EDOT) moieties, the O–I2 inter- configuration due to the presence of the substrate: the O–I
action is slightly stronger than the S–I2 one and that, in any distance is predicted to be 2.85 Å compared to the value of
case, both sites might effectively bind iodine, thus accounting 2.84 Å reported in Table 1, the C–I–O angle increases from
for the accelerated recombination rate measured for the dye 1201 to ca. 1501 and the distance from the surface is about
bearing two EDOT groups. This was recently confirmed for the 10 Å, as predicted by the Car–Parrinello dynamics.
dyes investigated here, indicating the oxygen atoms as the Therefore, a reasonable approximation of the average distance
preferred binding sites for I2. Moreover carbonylic or car- of the bonded I2 from the surface oxide can be obtained by
boxylic oxygens can also bind positive ions present in the taking the distances from the TiO2 cluster of the various I2
electrolyte solution, such as Li+, thus favoring the electrostatic binding sites in the dye@TiO2 optimized structures at 0 K
interaction between triiodide anions and the dye. The formation (Fig. 9). The distance of the I2 possibly bonded by the CN group
of dye–I2 or dye–I3 complexes in proximity of the TiO2 surface is about 5 Å, while the carbonylic oxygens, which are the
can play a crucial role in determining the short lifetimes predominant binding sites, lie at ca. 10 and 16 Å from the titania
measured for certain organic dyes containing oxygen atoms in in NKX2587@TiO2 and NKX2696@TiO2 respectively. The
the donor and bridge units.104 sulfur atoms in NKX2697 are at about 8, 11 and 14 Å from

15970 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012
View Article Online

interactions characterizing the DSCs functioning, possibly


contributing to the further development of this important
class of photovoltaic devices.
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

The dye sensitizer plays a crucial role in efficient DSCs; a


stringent dye characteristic is the presence of suitable anchoring
groups which should provide effective dye adsorption onto the
semiconducting oxide surface. The dye anchoring group promotes
electronic coupling between the donor levels of the excited dye
and the delocalized acceptor levels of the semiconductor conduc-
tion band, and mediates charge injection. Also the orientation
and packing of adsorbed dyes on the semiconductor surface
strictly depends on the binding motif, possibly affecting the rate
Fig. 8 Distance (in Å) of the carbonylic oxygen from the TiO2 and effectiveness of parasitic recombination reactions. Finally, the
surface against the simulation time for the NKX2587@(TiO2)38 sensitizer’s anchoring group should provide stable grafting of the
system. The average distance (in blue) and the maximum and mini- dye onto the semiconductor surface, thus leading to long-term
mum values (in red) are also reported. device stability. Understanding the dye adsorption mode onto the
semiconductor surfaces could thus contribute to individuate and
optimize many aspects that affect the DSCs performance.
A number of theoretical studies dealing with the adsorption
mode of the most commonly employed carboxylic acid groups
onto TiO2 have been reported to provide such understanding.
In some cases, the preferred adsorption mode was bidentate
bridging, with one proton transferred to a nearby surface
oxygen, while in other cases a monodentate (either dissociative
or undissociated) anchoring was predicted to be favored. Some
dependency of the relative stability of these two anchoring modes
on the employed computational method and the model was
however outlined, which did not allow a conclusive assignment
of the biding motif exhibited by the most performing dyes.
From a different perspective, experimental techniques usually
employed to probe dye surface adsorption (FT-IR, SERS, PES
and X-rays scattering) are often not sufficient to unravel the
atomistic details of the dye–semiconductor binding. As an
example, FT-IR spectroscopy, which is largely employed to
probe the adsorption mode of carboxylic acids on metal oxide
surfaces, usually has to deal with an envelop of overlapping
bands in the diagnostic region of the asymmetric and symmetric
Fig. 9 Optimized molecular structures of the O–I2 adduct of carboxylic stretchings which does not allow a clear distinction
NKX2587 and NKX2697 adsorbed onto the (TiO2)38 cluster. of such bands, thus of the dye adsorption mode.
To overcome the individual limitations of experimental and
the TiO2. The data by Miyashita et al.103 showed that the computational techniques, here we propose an integrated
NKX2587 sensitized cells had shorter lifetime than those based experimental/computational strategy which combines FT-IR
on NKX2697. The increase in the distance of the preferred measurements with DFT first principles simulations to individuate
carbonylic oxygen from the titania going from NKX2587@TiO2 the most energetically favorable adsorption mode for a carboxylic
(E10 Å) to NKX2697 (E16 Å) and hence the increased I2 acid probe, followed by calculation of the IR spectra and structure
concentration in proximity of the oxide surface might account validation by comparison with experimental data. Considering
for the differences in the lifetime values measured in ref. 103 for the complexity of the most efficient organic dyes, we chose the
NKX2587-sensitized solar cells compared to those employing the structural simplicity of acetic acid as a meaningful model for
longer NKX2697 homologous. realistic organic dyes.
We considered a dissociative bridged bidentate and two
monodentate adsorption modes (either dissociative or non-
4. Conclusions
dissociative), and calculated their relative stability with
We have presented a survey of new and recently published various models and at various levels of theory. As anticipated,
results obtained in the first principles computer simulation of we found a sizable variability in the relative stability of the
dye adsorption onto TiO2 surfaces, as models of the hetero- three adsorption modes with the model/method, which did not
interface crucial to the functioning of dye-sensitized solar cells alone allow us to trace clear conclusions concerning the favored
(DSCs). First-principles simulations have contributed to a deeper adsorption mode. We thus measured the FT-IR spectra of AcOH
understanding of the electronic structure and optical properties adsorbed on TiO2 nano-powders and calculated the vibrational
of new dyes and to unravel the intimate dye/semiconductor IR frequencies for the three representative adsorption modes.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15971
View Article Online

By comparing the two data sets we identified the bridged di Tecnologia, Project Seed 2009 ‘‘HELYOS’’, and Consiglio
bidentate adsorption mode as the one which most closely Nazionale delle Ricerche, Project ‘‘EFOR’’ for financial support.
matches the FT-IR frequency pattern, also being calculated The authors thank Paolo Salvatori for helpful discussions.
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

(although by a few kcal mol 1 only) as the most stable


adsorption mode by cluster calculations in solution.
We then compared the results of the present investigation to Notes and references
previous data from our lab, obtained for realistic dye sensiti- 1 B. O’Regan and M. Grätzel, Nature, 1991, 353, 737–740.
zers. Interestingly, the bridged bidentate adsorption mode was 2 M. Grätzel, Nature, 2001, 414, 338–344.
found to be the most stable binding motif also for organic dyes 3 M. Grätzel, Acc. Chem. Res., 2009, 42, 1788–1798.
bearing cyanoacrylic anchoring groups, while in the case of a 4 M. Grätzel, Inorg. Chem., 2005, 44, 6841–6851.
5 S. Ardo and G. J. Meyer, Chem. Soc. Rev., 2009, 38, 115–164.
rhodanine-3-acetic acid anchoring groups, an undissociated 6 A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo and H. Pettersson,
monodentate adsorption mode was found to be of almost Chem. Rev., 2010, 110, 6595–6663.
comparable stability. The two different cyanoacrylic and 7 N. A. Anderson and T. Lian, Coord. Chem. Rev., 2004, 248, 1231–1246.
8 K. Schwarzburg, R. Ernstorfer, S. Felber and F. Willig, Coord.
rhodanine-3-acetic acid anchoring groups were found to induce
Chem. Rev., 2004, 248, 1259–1270.
a totally different dye packing on the TiO2 surface, with dye 9 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,
molecules standing up or lying almost flat compared to the TiO2 E. Müller, P. Liska, N. Vlachopoulos and M. Grätzel, J. Am.
surface, respectively. These structural differences were related to Chem. Soc., 1993, 115, 6382–6390.
10 M. K. Nazeeruddin, F. De Angelis, S. Fantacci, A. Selloni,
the different electron injection/recombination with oxidized dye G. Viscardi, P. Liska, S. Ito, T. Bhesso and M. Grätzel, J. Am.
properties which were experimentally assessed for the two Chem. Soc., 2005, 127, 16835–16847.
classes of dyes. Furthermore, considering the cyanoacrylic 11 M. K. Nazeeruddin, P. Pechy and M. Grätzel, Chem. Commun.,
anchoring mode, a stronger coupling between the dye LUMO 1997, 1705–1706.
12 M. K. Nazeeruddin, P. Pechy, T. Renouard, S. M. Zakeeruddin,
and the TiO2 manifold of unoccupied states was calculated for R. Humphry-Baker, P. Comte, P. Liska, L. Cevey, E. Costa,
the bridged bidentate than for the monodentate adsorption V. Shklover, L. Spiccia, G. B. Deacon, C. A. Bignozzi and
mode, suggesting a possibly faster electron injection from the M. Grätzel, J. Am. Chem. Soc., 2001, 123, 1613–1624.
13 L. Han, A. Islam, H. Chen, M. Chandrasekharam,
dyes anchored in the bridged bidentate mode. B. Chiranjeevi, S. Zhang, X. Yang and M. Yanagida, Energy
To contribute to the understanding of dye–electrolyte inter- Environ. Sci., 2012, 5, 6057–6060.
actions, we investigated the adsorption mode of various prototype 14 P. Wang, S. M. Zakeeruddin, I. Exnar and M. Grätzel, Chem.
organic dyes on TiO2, along with I2 binding to the various electron- Commun., 2002, 2972–2973.
15 C. Y. Chen, S. J. Wu, C. G. Wu, J. G. Chen and K. C. Ho, Angew.
donating sites in the molecule. We found that oxygen-containing Chem., Int. Ed., 2006, 45, 5822–5825.
groups in the dyes (either EDOT or carbonyl groups) may 16 F. Gao, Y. Wang, D. Shi, J. Zhang, M. Wang, X. Jing,
constitute sites for I2 binding, thus accounting for the accelerated R. Humphry-Baker, P. Wang, S. M. Zakeeruddin and
M. Grätzel, J. Am. Chem. Soc., 2008, 130, 10720–10728.
recombination rate measured for the dye bearing such groups. By 17 T. Bhesso, E. Yoneda, J.-H. Yum, M. Guglielmi, I. Tavernelli,
performing combined Car–Parrinello molecular dynamics and H. Imai, U. Rothlisberger, M. K. Nazeeruddin and M. Grätzel,
structural geometry optimizations for two coumarin dyes adsorbed J. Am. Chem. Soc., 2009, 131, 5930–5934.
onto TiO2, differing by the length of the p-bridge separating the 18 P. G. Bomben, B. D. Koivisto and C. P. Berlinguette, Inorg.
Chem., 2010, 49, 4960–4971.
donor and acceptor moieties, we were able to relate the decreasing 19 A. Mishra, M. K. R. Fischer and P. Bäuerle, Angew. Chem., Int.
distance of the preferred carbonylic oxygen from the titania to an Ed., 2009, 48, 2474–2499.
increased I2 concentration in proximity of the oxide surface, which 20 M. Pastore, E. Mosconi, S. Fantacci and F. De Angelis, Curr.
might account for the differences in the electron lifetime values Org. Synth., 2012, 9, 215–232.
21 W. Zeng, Y. Cao, Y. Bai, Y. Wang, Y. Shi, M. Zhang, F. Wang,
measured for DSCs based on the shorter dye. C. Pan and P. Wang, Chem. Mater., 2010, 22, 1915–1925.
In conclusion, we believe that computer simulations based 22 J.-H. Yum, E. Baranoff, F. Kessler, T. Moehl, S. Ahmad,
on a first principles description of dye adsorption modes onto T. Bessho, A. Marchioro, E. Ghadiri, J.-E. Moser, C. Yi,
M. K. Nazeeruddin and M. Grätzel, Nat. Commun., 2012, 3, 631.
semiconductor oxides, lying at the heart of DSCs, represent a 23 S. L. Wu, H. P. Lu, H. T. Yu, S. H. Chuang, C. L. Chiu,
powerful tool for unravelling the atomistic feature of the C. W. Lee, E. W. G. Diau and C. Y. Yeh, Energy Environ. Sci.,
crucial DSCs interface. This in turn may allow for a detailed 2010, 3, 949–955.
understanding of the factors affecting the efficiency of the basic 24 Y. C. Chang, C. L. Wang, T. Y. Pan, S. H. Hong, C. M. Lan,
H. H. Kuo, C. F. Lo, H. Y. Hsu, C. Y. Lin and E. W. G. Diau,
charge generations/recombination processes, possibly leading Chem. Commun., 2011, 47, 8910–8912.
to improved DSCs photovoltaic performances. 25 A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran,
The interplay between theory/simulation and experiments M. K. Nazeeruddin, E. W. G. Diau, C. Y. Yeh, S. M.
Zakeeruddin and M. Grätzel, Science, 2011, 334, 629–634.
appears key, in our view, to further DSCs progress, both
26 S. Ito, P. Chen, P. Comte, M. K. Nazeeruddin, P. Liska, P. Pechy
concerning the design of new dye sensitizers and the under- and M. Grätzel, Prog. Photovoltaics, 2007, 15, 603–612.
standing of the sensitizers’ interaction with the semiconductor 27 K. Shankar, G. K. Mor, H. E. Prakasam, S. Yoriya, M. Paulose,
and with the solution environment and/or the electrolyte upon O. K. Varghese and C. A. Grimes, Nanotechnology, 2007, 18, 065707.
28 M. Saito and S. Fujihara, Energy Environ. Sci., 2008, 1, 280–283.
being adsorbed onto the semiconductor. 29 K. Keis, J. Lindgren, S. E. Lindquist and A. Hagfeldt, Langmuir,
2000, 16, 4688–4694.
30 S. Ferrere, A. Zaban and B. A. Gregg, J. Phys. Chem. B, 1997,
Acknowledgements 101, 4490–4493.
31 A. Kay and M. Grätzel, Chem. Mater., 2002, 14, 2930–2935.
We thank FP7-NMP-2009, project 246124 ‘‘SANS’’, FP7- 32 P. Persson, M. J. Lundqvist, R. Ernstofer, W. A. Goddard III and
ENERGY-2010, project 261820 ‘‘ESCORT’’, Istituto Italiano F. Willig, J. Chem. Theory Comput., 2006, 2, 441–451.

15972 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012
View Article Online

33 M. J. Lundqvist, M. Nilsing, S. Lunell, B. Åkemark and 70 N. Martsinovich and A. Troisi, J. Phys. Chem. C, 2011, 115, 11781.
P. Persson, Inorg. Chem., 2006, 110, 20513–20525. 71 F. De Angelis, Chem. Phys. Lett., 2010, 493, 323.
34 F. Ambrosio, N. Martsinovich and A. Troisi, J. Phys. Chem., 72 F. Labat and C. Adamo, J. Phys. Chem. C, 2007, 111, 15034.
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

2012, 116, 2622–2629. 73 R. Luschtinetz, S. Gemming and G. Seifert, Eur. Phys. J. Plus,
35 J. Wiberg, T. Marinado, D. P. Hagberg, L. Sun, A. Hagfeldt and 2011, 126, 1.
B. Albinsson, J. Phys. Chem. C, 2009, 113, 3881–3886. 74 S. Manzhos, H. Segawa and K. Yamashita, Phys. Chem. Chem.
36 M. Pastore and F. De Angelis, ACS Nano, 2010, 4, 556–562. Phys., 2012, 14, 1749.
37 M. Pastore and F. De Angelis, Phys. Chem. Chem. Phys., 2012, 75 P. Persson, R. Bergstrom and S. Lunell, J. Phys. Chem. B, 2000,
14, 920–928. 104, 10348.
38 M. Grätzel, J. Photochem. Photobiol. A, 2004, 168–235. 76 E. Mosconi, A. Selloni and F. De Angelis, J. Phys. Chem. C,
39 F. Odobel, E. Blart, M. Lagree, M. Vilieras, H. Baujtita, N. El 2011, 115, 25219.
Murr, S. Caramori and C. A. Bignozzi, J. Mat. Chem., 2003, 13, 77 A. Vittadini, A. Selloni, F. P. Rotzinger and M. Grätzel, J. Phys.
502–510. Chem. B, 2000, 104, 1300.
40 A. Abbotto, N. Manfredi, C. Marinzi, F. De Angelis, E. Mosconi, 78 G. te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca
J.-H. Yum, Z. Xianxi, M. K. Nazeeruddin and M. Grätzel, Guerra, S. J. A. van Gisbergen, J. G. Snijders and T. Ziegler,
Energy Environ. Sci., 2009, 2, 1094. J. Comput. Chem., 2001, 22, 931.
41 R. Argazzi, C. A. Bignozzi, M. Yang, G. Hasselmann and 79 J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996,
G. Meyer, Nano Lett., 2002, 2, 625–628. 77, 3865.
42 R. Katoh, M. Kasuya, A. Furube, N. Fuke, N. Koide and 80 A. D. Becke, Phys. Rev. A, 1988, 38, 309830100.
L. Han, Sol. Energy Mater. Sol. Cells, 2009, 93, 698. 81 V. Barone, M. Cossi and J. Tomasi, J. Chem. Phys., 1997,
43 K. Hara, Y. Dan-oh, C. Kasada, Y. Ohga, A. Shinpo, S. Suga, 107, 3210.
K. Sayama and H. Arakawa, Langmuir, 2004, 20, 4205–4210. 82 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
44 A. Kay and M. Grätzel, J. Phys. Chem., 1993, 97, 6272. M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven,
45 Y. Liu, A. Hagfeldt, X.-R. Xiao and S. E. Lindquist, Sol. Energy K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi,
Mater. Sol. Cells, 1998, 55, 267. V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega,
46 Z. Hui, Y. Xiong, L. Heng, L. Yuan and W. Yu-Xiang, Chim. G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota,
Phys. Lett., 2007, 24, 3272. R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
47 P. Falaras, Sol. Energy Mater. Sol. Cells, 1998, 53, 163–175. O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox,
48 K. S. Finnie, J. R. Bartlett and J. L. Woolfrey, Langmuir, 1998, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo,
14, 2744–2749. J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev,
49 E. M. J. Johansson, T. Edvinsson, M. Odelius, D. P. Hagberg, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski,
L. Sun, A. Hagfeldt, H. Siegbahn and H. Rensmo, J. Phys. Chem. P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador,
C, 2007, 111, 8580–8586. J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich,
50 T. Marinado, D. Hagberg, M. Hedlund, T. Edvinsson, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick,
E. Johansson, G. Boschloo, H. Rensmo, T. Brinck, L. Sun and A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz,
A. Hagfeldt, Phys. Chem. Chem. Phys., 2009, 11, 133–141. Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov,
51 M. Hahlin, E. Johansson, S. Plogmaker, M. Odelius, L. Sun, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin,
H. Siegbahn and H. Rensmo, Phys. Chem. Chem. Phys., 2010, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng,
12, 1507. A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson,
52 K. M. Karlsson, X. Jiang, S. K. Eriksson, E. Gabrielsson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 03,
H. Rensmo, A. Hagfeldt and L. Sun, Chem.–Eur. J., 2011, Revision C.02, Gaussian, Inc., Wallingford CT, 2004.
17, 6415. 83 R. Car and M. Parrinello, Phys. Rev. Lett., 1985, 55, 2471.
53 M. J. Griffith, M. James, G. Triani, P. Wagner, G. G. Wallace 84 P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car,
and D. L. Officer, Langmuir, 2011, 27, 12944–12950. C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni,
54 M. Nara, H. Torii and M. Tasumi, J. Phys. Chem., 1996, I. Dabo, A. Dal Corso, S. De Gironcoli, S. Fabris, G. Frates,
100, 19812. R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj,
55 G. B. Deacon and R. J. Phillips, Coord. Chem. Rev., 1980, 33, 227. M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri,
56 S. Fantacci and F. De Angelis, Coord. Chem. Rev., 2011, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto,
255, 2704. C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen,
57 N. Martsinovich and A. Troisi, Energy Environ. Sci., 2011, A. Smogunov, P. Umaril and R. M. Wentzcovitch, J. Phys.:
4, 4473. Condens. Matter, 2009, 21, 395502.
58 K. Srinivas, K. Yesudas, K. Bhanuprakash, V. J. Rao and 85 Z.-F. Pei and V. Ponec, Appl. Surf. Sci., 1996, 103, 171.
L. Giribabu, J. Phys. Chem. C, 2009, 113, 20117. 86 M. A. Hasan, M. I. Zaki and L. Pasupulety, Appl. Catal., A, 2003,
59 F. De Angelis, S. Fantacci and R. Gebauer, J. Phys. Chem. Lett., 243, 81.
2011, 2, 813. 87 A. S. Foster and R. M. Nieminen, J. Chem. Phys., 2004,
60 M. Pastore and F. De Angelis, J. Phys. Chem. Lett., 2011, 2, 1261. 121, 9039.
61 F. Ambrosio, N. Martsinovich and A. Troisi, J. Phys. Chem. 88 K. L. Miller, C. B. Musgrave, J. L. Falconer and J. W. Medlin,
Lett., 2012, 3, 1531. J. Phys. Chem. C, 2011, 115, 2738.
62 F. De Angelis, S. Fantacci, A. Selloni, M. K. Nazeeruddin and 89 K. L. Miller, J. L. Falconer and J. W. Medlin, J. Catal., 2011,
M. Grätzel, J. Am. Chem. Soc., 2007, 129, 14156. 278, 321.
63 F. De Angelis, S. Fantacci, A. Selloni, M. Grätzel and 90 D. P. Hagberg, T. Edvinsson, T. Marinado, G. Boschloo,
M. K. Nazeeruddin, Nano Lett., 2007, 7, 3189. A. Hagfeldt and L. C. Sun, Chem. Commun., 2006, 2245.
64 F. Schiffmann, J. VandeVondele, J. Hutter, R. Wirz, A. Urakawa 91 D. P. Hagberg, J.-H. Yum, H. Lee, F. De Angelis, T. Marinado,
and A. Baiker, J. Phys. Chem. C, 2010, 114, 8398. K. M. Karlsson, R. Humphry-Baker, L. Sun, A. Hagfeldt,
65 P. Chen, J. H. Yum, F. D. Angelis, E. Mosconi, S. Fantacci, M. Grätzel and M. K. Nazeeruddin, J. Am. Chem. Soc., 2008,
S.-J. Moon, R. H. Baker, J. Ko, M. K. Nazeeruddin and 130, 6259.
M. Grätzel, Nano Lett., 2009, 9, 2487. 92 J. Wiberg, T. Marinado, D. P. Hagberg, L. Sun, A. Hagfeldt and
66 F. De Angelis, S. Fantacci, A. Selloni, M. K. Nazeeruddin and B. Albinsson, J. Phys. Chem. C, 2009, 113, 3881.
M. Grätzel, J. Phys. Chem. C, 2010, 114, 6054. 93 D. P. Hagberg, T. Marinado, K. M. Karlsson, K. Nonomura,
67 F. De Angelis, S. Fantacci, E. Mosconi, M. K. Nazeeruddin and P. Qin, G. Boschloo, T. Brinck, A. Hagfeldt and L. Sun, J. Org.
M. Grätzel, J. Phys. Chem. C, 2011, 115, 8825. Chem., 2007, 72, 9550.
68 D. Rocca, R. Gebauer, F. De Angelis, M. K. Nazeeruddin and 94 W. H. Howie, F. Claeyssens, H. Miura and L. M. Peter, J. Am.
S. Baroni, Chem. Phys. Lett., 2009, 475, 49. Chem. Soc., 2008, 130, 1367.
69 N. Martsinovich, D. R. Jones and A. Troisi, J. Phys. Chem. C, 95 U. B. Cappel, S. M. Feldt, J. Schoneboom, A. Hagfeldt and
2010, 114, 22659. G. Boschloo, J. Am. Chem. Soc., 2010, 132, 9096.

This journal is c the Owner Societies 2012 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 15973
View Article Online

96 H. Long, D. Zhou, M. Zhang, C. Peng, S. Uchida and P. Wang, 105 Y. Bai, J. Zhang, D. Zhou, Y. Wang, M. Zhang and P. Wang,
J. Phys. Chem. C, 2011, 115, 14408. J. Am. Chem. Soc., 2011, 133, 11442.
97 J. P. Muscat and D. M. Newns, Prog. Surf. Sci., 1978, 9, 1. 106 M. Tuikka, P. Hirva, K. Rissanen, J. Korppi-Tommola and
Published on 11 October 2012. Downloaded by Centro de Investigacion en Materiales Avanzados SC on 07/04/2015 16:49:05.

98 G. Boschloo and A. Hagfeldt, Acc. Chem. Res., 2009, 42, M. Haukka, Chem. Commun., 2011, 47, 4499.
1819. 107 T. Privalov, G. Boschloo, A. Hagfeldt, P. H. Svensson and
99 S. Ito, H. Miura, S. Uchida, M. Takata, K. Sumioka, P. Liska, L. Kloo, J. Phys. Chem. C, 2009, 113, 783.
P. Comte, P. Pechy and M. Grätzel, Chem. Commun., 2008, 5194. 108 C.-H. Hu, A. M Asaduzzaman and G. Schreckenbach, J. Phys.
100 N. Koumura, Z.-S. Wang, S. Mori, M. Miyashita, E. Suzuki and Chem. C, 2010, 114, 15165.
K. Hara, J. Am. Chem. Soc., 2006, 128, 14256. 109 F. Schiffmann, J. VandeVondele, J. Hutter, A. Urakawa, R. Wirz
101 A. Burke, S. Ito, H. J. Snaith, U. Bach, J. Kwiatkowski and and A. Baiker, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 4830.
M. Grätzel, Nano Lett., 2008, 8, 977. 110 J. Nyhlen, G. Boschloo, A. Hagfeldt, L. Kloo and T. Privalov,
102 B. C. O’Regan, K. Walley, M. Juozapavicius, A. Y. Anderson, Chem. Phys. Chem., 2010, 11, 1858–1862.
F. Matar, T. Ghaddar, S. M. Zakeeruddin, C. Klein and 111 H. Kusama, H. Sugihara and K. Sayama, J. Phys. Chem. C, 2011,
J. R. Durrant, J. Am. Chem. Soc., 2009, 131, 3541. 115, 2544–2552.
103 M. Miyashita, K. Sunahara, K. Nishikawa, Y. Uemura, 112 M. G. Lobello, S. Fantacci and F. De Angelis, J. Phys. Chem. C,
N. Koumura, K. Hara, A. Mori, T. Abe, E. Suzuki and 2011, 115, 18863.
S. Mori, J. Am. Chem. Soc., 2008, 130, 17874. 113 M. Pastore, E. Mosconi and F. De Angelis, J. Phys. Chem. C,
104 M. Planells, L. Pelleja, J. N. Clifford, M. Pastore, F. De Angelis, 2012, 116, 5965.
Ń. Lopez, S. R. Marder and E. Palomares, Energy Environ. Sci., 114 W. R. Duncan, C. F. Craig and O. V. Prezhdo, J. Am. Chem.
2011, 4, 1820. Soc., 2007, 129, 8528.

15974 Phys. Chem. Chem. Phys., 2012, 14, 15963–15974 This journal is c the Owner Societies 2012

You might also like